20.01.2013 Views

Catalytic Synthesis and Characterization of Biodegradable ...

Catalytic Synthesis and Characterization of Biodegradable ...

Catalytic Synthesis and Characterization of Biodegradable ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Catalytic</strong> <strong>Synthesis</strong> <strong>and</strong> <strong>Characterization</strong> <strong>of</strong><br />

<strong>Biodegradable</strong> Polyesters <strong>and</strong><br />

Their Radical Block Copolymers<br />

生分解性ポリエステルおよびそのラジカル<br />

ブロック共重合体の触媒的合成と特性<br />

A Thesis<br />

Presented to<br />

Waseda University<br />

July 2010<br />

Xiuli Zhuang<br />

庄 秀丽


Promoter: Pr<strong>of</strong>. Dr. Hiroyuki Nishide<br />

Referees: Pr<strong>of</strong>. Dr. Kuniki Kino<br />

Pr<strong>of</strong>. Dr. Timothy E. Long<br />

Assoc. Pr<strong>of</strong>. Dr. Kenichi Oyaizu


Preface<br />

<strong>Biodegradable</strong> polyesters have attracted great attentions to be developed as the environment<br />

friendly materials <strong>and</strong> biomedical materials due to their excellent biodegradable <strong>and</strong> biocompatible<br />

properties. Two main tasks, i.e. controlled synthesis <strong>and</strong> functionalization <strong>of</strong> polyesters, have<br />

emerged in order to obtain the polyesters with tunable properties for biomedical applications. Great<br />

advances have been achieved by virtue <strong>of</strong> the recent progresses on the controlled living<br />

polymerizations <strong>and</strong> polyesters functionalities. However, efforts still remained on the synthesis <strong>of</strong><br />

polyesters with more controllable structures <strong>and</strong> smart properties.<br />

In this thesis, the author has provided two strategies to synthesize biodegradable polyesters by<br />

using a series <strong>of</strong> newly synthesized cobalt-Schiff-base complexes. Controlled synthesis <strong>of</strong><br />

polylactide (PLA)-analogues polyesters are expected based on the studies on the catalyst’s activity<br />

<strong>and</strong> polymerization mechanism. Moreover, functionality <strong>of</strong> PLA by incorporating with stable<br />

TEMPO radical-substituted polymers has also been synthesized <strong>and</strong> investigated for the biomedical<br />

applications in tissue engineering <strong>and</strong> bioimaing. Chapter 1 reviews on biodegradable polyesters,<br />

electro-active polymers <strong>and</strong> radical polymers, plus their applications in biomedical fields. Chapter 2<br />

describes the synthesis <strong>of</strong> a series <strong>of</strong> cobalt-Schiff-base catalysts for alternative copolymerization <strong>of</strong><br />

racemic propylene oxide <strong>and</strong> carbon dioxide. Chapter 3 describes the application <strong>of</strong><br />

cobalt-Schiff-base catalysts in ring-opening polymerization <strong>of</strong> lactic acid-O-NCA (LCA) monomer<br />

to synthesize PLA. Chapter 4 describes the synthesis <strong>and</strong> bio-valuations <strong>of</strong> TEMPO-contained<br />

PLA-PTAm r<strong>and</strong>om copolymers. Chapter 5 deals with the controlled synthesis <strong>of</strong> block copolymers<br />

bearing radical polymer segments by RAFT polymerization method <strong>and</strong> their biomedical<br />

applications. Chapter 6 deals with the synthesis <strong>of</strong> amphphilic di- <strong>and</strong> tri-block copolymers<br />

containing stable TEMPO radical segment <strong>and</strong> their untiliting as EPR bioimaging probes. Chapter 7<br />

concludes this thesis <strong>and</strong> proposes the perspective <strong>of</strong> the controlled synthesis, functionalization <strong>and</strong><br />

biomedical applications <strong>of</strong> polyesters.<br />

III<br />

Xiuli Zhuang


Preface<br />

<strong>Catalytic</strong> <strong>Synthesis</strong> <strong>and</strong> <strong>Characterization</strong> <strong>of</strong><br />

<strong>Biodegradable</strong> Polyesters <strong>and</strong><br />

Their Radical Block Copolymers<br />

生分解性ポリエステルおよびそのラジカル<br />

ブロック共重合体の触媒的合成と特性<br />

Contents<br />

Chapter 1 Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

1.1 Introduction …..2<br />

1.2 Copolymer <strong>of</strong> Epoxide <strong>and</strong> Carbon Dioxide …..2<br />

1.3 Polylactide …11<br />

1.4 Others <strong>Biodegradable</strong> Polyesters …15<br />

1.5 Electroactive Polymers …18<br />

1.6 Radical Polymers …39<br />

References …49<br />

Chapter 2 <strong>Synthesis</strong> <strong>and</strong> Electrochemistry <strong>of</strong> Schiff Base Cobalt(III) Complexes<br />

<strong>and</strong> Their <strong>Catalytic</strong> Activity for Copolymerization <strong>of</strong> Epoxide <strong>and</strong> Carbon Dioxide<br />

2.1 Introduction …60<br />

2.2 <strong>Synthesis</strong> <strong>and</strong> <strong>Characterization</strong> <strong>of</strong> L-Co III -dnp Complexes …60<br />

2.3 Copolymerization <strong>of</strong> CO2 <strong>and</strong> rac-PO …63<br />

2.4 Electrochemistry <strong>of</strong> Cobalt(III) Complexes …66<br />

2.5 Conclusions …68<br />

2.6 Experimental Section …68<br />

References …71<br />

IV


Chapter 3 Polymerization <strong>of</strong> Lactic O-Carboxylic Anhydride using Organometallic<br />

Catalysts<br />

3.1 Introduction …76<br />

3.2 Polymerization <strong>of</strong> LacOCA using Co(III) Complexes with<br />

Schiff Base Lig<strong>and</strong>s …77<br />

3.3 Polymerization <strong>of</strong> LacOCA using Tin(II) or Al(III) Based Complexes …79<br />

3.4 <strong>Characterization</strong> <strong>of</strong> the Obtained PLA …80<br />

3.5 Conclusions …82<br />

3.6 Experimental Section …82<br />

References …83<br />

Chapter 4 <strong>Synthesis</strong> <strong>of</strong> Lactide-Grafted Poly(TEMPO-acrylamide) <strong>and</strong> its<br />

Electrochemical Property <strong>and</strong> Biocompatibility<br />

4.1 Introduction …86<br />

4.2 <strong>Synthesis</strong> <strong>of</strong> PTAm-g-PLA …87<br />

4.3 Electrochemical Properties …90<br />

4.4 Thermal Properties …92<br />

4.5 Phase Separation Behavior <strong>of</strong> PTAm-g-PLA …92<br />

4.6 Cytotoxicity <strong>of</strong> PTAm-g-PLA …93<br />

4.7 Cell Adhesion <strong>and</strong> Spreading …94<br />

4.8 Conclusions …95<br />

4.9 Experimental Section …95<br />

References …98<br />

Chapter 5 <strong>Synthesis</strong> <strong>of</strong> Triblock Copolymers <strong>of</strong> TEMPO-Acrylamide <strong>and</strong> Lactate by<br />

RAFT Polymerization<br />

5.1 Inroduction ..102<br />

5.2 <strong>Synthesis</strong> <strong>of</strong> Triblock Copolymer ..103<br />

5.3 Thermal Properties ..105<br />

V


5.4 Cytotoxicity <strong>and</strong> Biocompatibility <strong>of</strong> Copolymers ..106<br />

5.5 Electrospun <strong>of</strong> Copolymers ..107<br />

5.6 Conclusions ..108<br />

5.7 Experimental Section ..108<br />

References ..110<br />

Chapter 6 <strong>Synthesis</strong> <strong>of</strong> Amphiphilic Triblock Copolymers <strong>of</strong> Acrylamide, Lactate,<br />

<strong>and</strong> Ethyleneoxide<br />

6.1 Inroduction ..114<br />

6.2 <strong>Synthesis</strong> <strong>of</strong> MPEG-CPAD <strong>and</strong> MPEG-b-PLA-CPAD ..115<br />

6.3 <strong>Synthesis</strong> <strong>of</strong> MPEG-b-PTAm <strong>and</strong> MPEG-b-PLA-b-PTAm ..116<br />

6.4 Micelles preparation <strong>and</strong> characterization ..119<br />

6.5 EPR study <strong>of</strong> the micelles in PBS containing ascorbic acid ..120<br />

6.6 Cytotoxicity assay ..121<br />

6.7 Conclusions ..122<br />

6.8 Experimental section ..123<br />

References ..126<br />

Chapter 7 Conclusion <strong>and</strong> Future Prospects<br />

7.1. Conclusion ..130<br />

7.2. Future Prospects ..132<br />

List <strong>of</strong> publications<br />

Acknowledgements<br />

VI


Chapter 1<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

1.1 Introduction<br />

1.2 Copolymer <strong>of</strong> Epoxide <strong>and</strong> Carbon Dioxide<br />

1.3 Polylactide<br />

1.4 Others <strong>Biodegradable</strong> Polyesters<br />

1.5 Electroactive Polymers<br />

1.6 Radical Polymers<br />

References


Chapter 1<br />

1.1 Introduction<br />

<strong>Biodegradable</strong> polyesters are one <strong>of</strong> the main types <strong>of</strong> biomaterials. <strong>Biodegradable</strong><br />

polyesters like polyglycolide, polylactides, poly-ɛ-caprolactone, polycarbonates <strong>and</strong> their<br />

copolymers have found a wide range <strong>of</strong> applications, such as sutures, bone fracture fixation<br />

devices, drug controlled release carriers, tissue engineering scaffolds <strong>and</strong> green plastics as<br />

wrapping materials, disposal containers <strong>and</strong> fibers because <strong>of</strong> their biodegradability <strong>and</strong><br />

biocompatibility. 1 In recent years, developments in tissue engineering, regenerative medicine,<br />

gene therapy, <strong>and</strong> controlled drug delivery have promoted the need <strong>of</strong> new properties <strong>of</strong> both<br />

biologically derived <strong>and</strong> synthetic biodegradable polyesters with biodegradability. 1<br />

<strong>Biodegradable</strong> polyesters with diverse special properties are needed for in vivo<br />

applications because <strong>of</strong> the diversity <strong>and</strong> complexity <strong>of</strong> in vivo environments. Although<br />

biologically derived biodegradable polymers possess good biocompatibility, synthetic<br />

polyesters have been becoming better alternatives for biomedical applications because <strong>of</strong> the<br />

following reasons: (1) chemical modifications for biologically derived biodegradable<br />

polymers are difficult; <strong>and</strong> (2) chemical modifications likely cause the denaturation <strong>of</strong> the<br />

bulk properties <strong>of</strong> the biologically derived biodegradable polymers. In contrary, numerous<br />

properties can be obtained for synthetic polyesters <strong>and</strong> further modifications are easy to be<br />

carried out for as-prepared synthetic biomaterials.<br />

1.2 Copolymer <strong>of</strong> Epoxide <strong>and</strong> Carbon Dioxide<br />

1.2.1 Introduction<br />

Because carbon dioxide (CO2) is an abundant, inexpensive, <strong>and</strong> nontoxic biorenewable<br />

resource, it is an attractive raw material for incorporation into important industrial processes.<br />

Eminent on the list <strong>of</strong> processes that are technologically viable is the use <strong>of</strong> CO2 as both a<br />

monomer <strong>and</strong> a solvent in the manufacturing <strong>of</strong> biodegradable copolymers, most notably,<br />

polycarbonates. This process is illustrated in Scheme 1.2.1 for the copolymerization <strong>of</strong><br />

cyclohexene oxide <strong>and</strong> CO2 to afford poly(cyclohexene carbonate). As indicated in Scheme<br />

1.2.1, in general, this process is accompanied by the production <strong>of</strong> varying quantities <strong>of</strong><br />

five-membered cyclic carbonates. There appears to be some confusion among members <strong>of</strong> the<br />

scientific community who object to referring to the use <strong>of</strong> CO2 as a raw material for<br />

generating useful chemicals as “green chemistry”. It has been apparent to most <strong>of</strong> us that the<br />

‐ 2 ‐


- 3 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

small quantity <strong>of</strong> CO2 consumed in these processes is likely always to be a very small fraction<br />

<strong>of</strong> the total CO2 produced from fossil fuel combustion <strong>and</strong> other sources. Although this<br />

consumption would have a effect on global warming, its use is considered “green chemistry”<br />

in the context <strong>of</strong> providing more environmentally benign routes to producing chemicals<br />

otherwise made utilizing reagents detrimental to the environment.<br />

Scheme 1.2.1<br />

Both the monomeric <strong>and</strong> the polymeric products provided from the coupling <strong>of</strong> CO2 <strong>and</strong><br />

epoxides have important industrial applications. Polycarbonates possess outst<strong>and</strong>ing<br />

properties, which include strength, lightness, durability, high transparency, heat resistance,<br />

<strong>and</strong> good electrical insulation; in addition, they are easily processed <strong>and</strong> colored. 2 Hence,<br />

these materials have wide-scale uses in electronics, optical media, glazing <strong>and</strong> sheeting, the<br />

automotive industry, the medical <strong>and</strong> healthcare industry, <strong>and</strong> many other consumable goods.<br />

On the other h<strong>and</strong>, cyclic carbonates find numerous applicabilities, most importantly as high<br />

boiling <strong>and</strong> flash point solvents with low odor/toxicity in degreasing, paint stripping, <strong>and</strong><br />

cleaning processes. 3 These monomeric compounds are also extensively employed as reactive<br />

intermediates.<br />

In 1969, Inoue <strong>and</strong> co-workers made the remarkable discovery that a mixture <strong>of</strong> ZnEt2 <strong>and</strong><br />

H2O was active for catalyzing the alternating copolymerization <strong>of</strong> propyleneoxide (PO) <strong>and</strong><br />

CO2, marking the advent <strong>of</strong> epoxide–CO2 coupling chemistry. 4, 5 An optimum 1:1 ratio <strong>of</strong><br />

ZnEt2/H2O gave the best yields <strong>of</strong> methanol-insoluble PPC with an activity <strong>of</strong> 0.12 h -1 (mol <strong>of</strong><br />

PO converted to polymer per mol Zn per h) at 80 o C <strong>and</strong> 20–50 atm CO2. On the basis <strong>of</strong><br />

elemental analysis, the copolymer contained 88% carbonate linkages. Notably, a 1:1 mixture<br />

<strong>of</strong> ZnEt2 <strong>and</strong> MeOH did not generate an active catalytic species for polycarbonate synthesis.<br />

Following this initial lead, Inoue investigated the use <strong>of</strong> dihydric sources, including<br />

resorcinol, 6, 7 dicarboxylic acids, 8 <strong>and</strong> primary amines, 9 in mixtures with ZnEt2 for PO–CO2<br />

copolymerization. These systems showed TOFs <strong>of</strong> 0.17, 0.43, <strong>and</strong> 0.06 h -1 , respectively.


Chapter 1<br />

The next major breakthrough in the area <strong>of</strong> propylene oxide <strong>and</strong> CO2 copolymerization<br />

came with the uncovering <strong>of</strong> air-stable dicarboxylic acid derivatives <strong>of</strong> zinc by Soga <strong>and</strong><br />

co-workers. 10 These catalysts were prepared by the reaction <strong>of</strong> zinc hydroxide or zinc oxide<br />

with dicarboxylic acids in toluene. Although the zinc glutarate analogue was found to be<br />

catalytically most active, producing significant quantities <strong>of</strong> polypropylene carbonate from<br />

propylene oxide <strong>and</strong> CO2, the percent <strong>of</strong> active zinc sites on this heterogeneous catalyst was<br />

quite low (


- 5 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

in the presence <strong>of</strong> a quaternary organic salt or triphenylphosphine. At ambient temperature<br />

<strong>and</strong> a relatively high catalyst loading, reaction rates were slow, generally requiring 12-23 days<br />

<strong>and</strong> thereby affording rather low molecular weight polymers. Nevertheless, copolymers from<br />

ethylene oxide, propylene oxide, or cyclohexene oxide <strong>and</strong> carbon dioxide with very narrow<br />

molecular weight distributions (1.06-1.14) were isolated. Approximately a decade later,<br />

Kruper <strong>and</strong> Dellar investigated the use <strong>of</strong> (tpp)CrX complexes in the presence <strong>of</strong> 4-10 equiv<br />

<strong>of</strong> nitrogen donors, such as N-methylimidazole (N-MeIm) or (4-dimethylamino)pyridine<br />

(DMAP), for the coupling <strong>of</strong> a wide variety <strong>of</strong> epoxides <strong>and</strong> carbon dioxide to provide cyclic<br />

carbonates. 15 For example, propylene oxide readily afforded propylene carbonate at 50 bar <strong>of</strong><br />

CO2 pressures with a TOF <strong>of</strong> 158 h -1 at 80 . Conversely, cyclohexene oxide <strong>and</strong> CO2<br />

yielded predominantly copolymer under similar reaction conditions. In more recent studies,<br />

Holmes <strong>and</strong> co-workers have reported the copolymerization <strong>of</strong> cyclohexene oxide <strong>and</strong> CO2 in<br />

the presence <strong>of</strong> a fluorinated (tpp)CrCl catalyst (Figure 1.2.2) along with a cocatalyst such as<br />

DMAP. 16 TOFs greater than 150 h -1 were observed for reactions performed at 110 in scCO2<br />

(225 bar) where the catalyst solubility is greatly enhanced over its (tpp)CrCl analogue, with<br />

the resulting copolymers having a high percentage <strong>of</strong> carbonate linkages (>97%) <strong>and</strong> narrow<br />

polydispersities (PDIs). The rapid advances currently being experienced in the<br />

copolymerization <strong>of</strong> cyclohexene oxide <strong>and</strong> carbon dioxide were fueled by Darensbourg’s<br />

development <strong>of</strong> a series <strong>of</strong> discrete zinc phenoxide derivatives as catalysts in 1995. 17 A<br />

typical bis(phenoxide)Zn(THF)2 complex (where THF = tetrahydr<strong>of</strong>uran) used in these<br />

studies is illustrated in Figure 1.2.3, where the steric bulk <strong>of</strong> the phenoxide lig<strong>and</strong>s limits<br />

aggregate formation.<br />

Figure 1.2.2 Fluorinated porphyrin derivative <strong>of</strong> chromium(III) chloride.<br />

These catalysts were extremely effective for homopolymerizing epoxides to polyether;


Chapter 1<br />

hence, the CO2 content <strong>of</strong> the copolymer provided by this method was generally about 90%<br />

even though the reactions were carried out at high CO2 pressures. 18 The copolymerization<br />

processes were performed in the absence <strong>of</strong> an organic solvent at 55 bar CO2 pressure <strong>and</strong> 80<br />

. Bulk polymerization reactions were run over a 24 h period, providing minimum values <strong>of</strong><br />

TOFs <strong>of</strong> about 10 h -1 with catalytic activity varying slightly as a function <strong>of</strong> the substituents<br />

on the phenoxide lig<strong>and</strong>s. Although the coupling reaction <strong>of</strong> propylene oxide <strong>and</strong> CO2<br />

provided predominantly propylene carbonate at 80 , at 40 , the reaction selectivity<br />

switched mainly to copolymer formation. This temperature dependence <strong>of</strong> product selectivity<br />

is a general phenomenon observed for most catalyst systems investigated. Terpolymers with<br />

up to 20% propylene oxide content were produced from reaction mixtures <strong>of</strong> propylene<br />

oxide/cyclohexene oxide <strong>and</strong> carbon dioxide. More moisture-tolerant dimeric zinc derivatives<br />

(Figure 1.2.4) containing the less sterically encumbering 2,6-difluorophenoxide were also<br />

shown to be quite active for the copolymerization process. 19<br />

Figure 1.2.3 Monomeric zinc-bis(phenoxide) complex for the copolymerization <strong>of</strong> cyclohexene<br />

oxide <strong>and</strong> CO2. The phenoxide lig<strong>and</strong> is 2,6-diisopropylphenoxide.<br />

Figure 1.2.4 Dimeric zinc-bis(phenoxide) complex for the copolymerization <strong>of</strong> cyclohexene<br />

oxide <strong>and</strong> CO2. The phenoxide lig<strong>and</strong> is 2,6-difluorophenoxide.<br />

‐ 6 ‐


1.2.2 Aluminum Catalysts<br />

- 7 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

In 1978, Inoue developed the first single-site catalysts for epoxide–CO2 copolymerization<br />

based on a tetraphenylporphyrin (tpp) lig<strong>and</strong> framework, 1a–d. 20 [(tpp)AlCl] (1a) <strong>and</strong><br />

[(tpp)AlOMe] (1b) (Figure 1.2.5) were found to be living initiators for the<br />

homopolymerization <strong>of</strong> PO <strong>and</strong> lactones, including lactide, b-butyrolactone, <strong>and</strong><br />

ɛ-caprolactone, as well as for the copolymerization <strong>of</strong> CO2 <strong>and</strong> epoxides <strong>and</strong> <strong>of</strong> PO <strong>and</strong><br />

phthalic anhydrides. 21-23 1a <strong>and</strong> 1b reacted with PO to form poly(propylene oxide) (PPO) in a<br />

living polymerization with PDIs <strong>of</strong> 1.07–1.15. The chloride initiator ring-opened the least<br />

hindered C-O bond <strong>and</strong> generated a regioregular PPO. In addition, 1b copolymerized PO <strong>and</strong><br />

CO2 at 20 o C <strong>and</strong> 8 atm CO2, giving PPC (Mn=3900 gmol -1 ; Mw/Mn=1.15) with 40%<br />

carbonate linkages over the course <strong>of</strong> 19 days. 23 Although molecular weights were low <strong>and</strong><br />

reaction times were long, this reaction marked the first example <strong>of</strong> monodisperse<br />

polycarbonates having a narrow PDI. The low molecular weights <strong>of</strong> polymers produced<br />

by{(tpp)Al} catalysts suggest chain transfer, which supports Inoue’s proposal <strong>of</strong> an<br />

“immortal” type polymerization. 20 An immortal polymerization allows for multiple chains to<br />

propagate from one metal center, whereas a living polymerization grows only one chain per<br />

metal center. Protic sources facilitate chain swapping such that there are more polymer chains<br />

than active catalytic sites. Free chains are dormant, but continue to grow polymer when<br />

exchanged onto the active site. If the chain swapping is more rapid than propagation, polymer<br />

chains with narrow PDIs are produced.<br />

Figure 1.2.5 Aluminum <strong>and</strong> manganese porphyrins for the homopolymerization <strong>of</strong> epoxides <strong>and</strong><br />

copolymerization <strong>of</strong> epoxides <strong>and</strong> CO2 (R=alkyl, oligomer <strong>of</strong> PPO).<br />

Recently, a salicylaldimine (salen)–aluminum complex, 2a, was found to be highly active


Chapter 1<br />

for the cyclization <strong>of</strong> EO <strong>and</strong> CO2 to ethylene carbonate (Figure 1.2.6). 24, 25 Lewis bases<br />

orquaternary ammonium salts, including pyridine, MeIm, <strong>and</strong> nBu4NX (X=Cl, Br, I), were<br />

utilized as cocatalysts, enhancing rates by up to a factor <strong>of</strong> five. At 110 o C <strong>and</strong> in scCO2 (ca.<br />

150 atm CO2), a 1:1 mixture <strong>of</strong> 2a/nBu4NBr catalyzed the conversion <strong>of</strong> EO to EC with a<br />

TOF <strong>of</strong> 2220 h -1 . Salen-chromium (2b) <strong>and</strong> -cobalt (2c) analogs also promoted cyclic-species<br />

formation showing rates <strong>of</strong> 2140 <strong>and</strong> 1320 h -1 , respectively. Darensbourg <strong>and</strong> co-workers<br />

have reported AlCl4 - -based complexes that exhibit TOFs up to 50 h -1 for the synthesis <strong>of</strong><br />

propylene carbonate from PO <strong>and</strong> CO2. 26<br />

Figure 1.2.6 Salen catalyst systems (cocatalyst = tetrabutylammonium halide, pyridine, or MeIm)<br />

for the synthesis <strong>of</strong> ethylene carbonate.<br />

Aluminum complexes are indeed active for the copolymerization <strong>of</strong> epoxides <strong>and</strong> CO2;<br />

however, they are plagued by low activities <strong>and</strong> yield polycarbonates with high percentages <strong>of</strong><br />

ether linkages. It appears that without additives, current aluminum catalysts do not cleanly<br />

generate alternating copolymer. Nevertheless, the “immortal” polymerization <strong>of</strong> [(tpp)AlX]<br />

compounds shows promise for the synthesis <strong>of</strong> a wealth <strong>of</strong> unique copolymers with varying<br />

levels <strong>of</strong> carbonate linkages provided the activities can be improved.<br />

1.2.3 Chromium Salen Catalysts<br />

Kruper <strong>and</strong> Dellar discovered that [(tpp)CrX] (Figure 1.2.7) in mixtures with 4–10<br />

equivalents <strong>of</strong> a Lewis-basic amine cocatalyst [such as MeIm or (4-dimethylamino)pyridine<br />

(DMAP)] are moderately active for the cyclization <strong>of</strong> epoxides <strong>and</strong> CO2. 27 A wide range <strong>of</strong><br />

epoxides, including PO, trans-2-butene oxide, epichlorohydrin, CHO, <strong>and</strong> cyclopentene oxide<br />

(CPO), were rapidly converted to the corresponding cyclic carbonates. For instance, one <strong>of</strong><br />

the (tpp)CrX <strong>and</strong> DMAP catalyzed CHC formation at 50 atm CO2 <strong>and</strong> 95 o C, exhibiting<br />

activities <strong>of</strong> 103 h -1 . In this case, PCHC was isolated as the major product. Following<br />

‐ 8 ‐


- 9 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

thermolysis, a 95:5 ratio <strong>of</strong> trans <strong>and</strong> cis CHC was observed, suggesting the possibility <strong>of</strong> dual<br />

mechanisms.<br />

Figure 1.2.7 Chromium–porphyrin complexes for the coupling <strong>of</strong> epoxides <strong>and</strong> CO2.<br />

Jacobsen <strong>and</strong> co-workers found [(salen)CrCl] complexes to be highly active in the<br />

asymmetric ring opening <strong>of</strong> epoxides. 28 This elegant work has since led to many crucial<br />

discoveries in the coupling <strong>of</strong> epoxides with CO2; in fact the first report <strong>of</strong><br />

(salen)chromiummediated epoxide–CO2 polymerization appeared in a 2000 patent by<br />

Jacobsen <strong>and</strong> co-workers. 29 Nguyen <strong>and</strong> Paddock reported highly active [(salen)CrCl]/DMAP<br />

based systems (Figure 1.2.8), for the cyclo-addition <strong>of</strong> CO2 <strong>and</strong> a variety <strong>of</strong> terminal aliphatic<br />

epoxides, including PO, epichlorohydrin, butadiene monoepoxide, <strong>and</strong> styrene oxide (SO). 30<br />

Figure 1.2.8 Salen–chromium <strong>and</strong> salen–cobalt complexes for the coupling <strong>of</strong> epoxides <strong>and</strong> CO2.<br />

Darensbourg <strong>and</strong> co-workers began efforts to uncover well-defined transition metal<br />

coordination complexes as catalysts for the coupling <strong>of</strong> CO2 <strong>and</strong> epoxides to selectively


Chapter 1<br />

provide polycarbonates. (Figure 1.2.9). 28 They use (Salen)M III X as a catalyst for the<br />

copolymerization <strong>of</strong> CO2 <strong>and</strong> cyclohexene oxide in the presence <strong>of</strong> N-MeIm. At 80 <strong>and</strong><br />

58.5 bar CO2 pressure, (Salen)M III X alone copolymerized cyclohexene oxide <strong>and</strong> carbon<br />

dioxide to poly(cyclohexylenecarbonate), void <strong>of</strong> polyether linkages, with a TOF <strong>of</strong> 10.4 h-1.<br />

1.2.4 Lanthanide-Based Catalysts<br />

Figure 1.2.9. (Salen)M III X catalyst for epoxides ring opening.<br />

Yttrium, aluminum, rare-earth metals, <strong>and</strong> combinations <strong>of</strong> multiple metal reagents have<br />

shown activity for the copolymerization <strong>of</strong> epoxides <strong>and</strong> CO2. For example, PO–CO2<br />

copolymerization was effected by a rare-earth metal system comprised <strong>of</strong> yttrium<br />

tris[bis(2-ethylhexyl)phosphate], AliBu3, <strong>and</strong> glycerol. The PPC produced contained only<br />

10–30% carbonate linkages, although molecular weights <strong>of</strong> up to 476000 gmol -1 were<br />

achieved. 31 This system also exhibited activity for the alternating copolymerization <strong>of</strong> CO2<br />

with epichlorohydrin 32 <strong>and</strong> glycidyl ether monomers. 33 Other rare-earth metal systems<br />

consisted <strong>of</strong> yttrium carboxylates [Y(CO2CF3)3 or Y(CO2RC6H4)3 where R is H, OH, Me, or<br />

NO2], ZnEt2, <strong>and</strong> glycerine. 34-36 The alternating copolymerization <strong>of</strong> CO2 <strong>and</strong> PO yielded PPC<br />

with up to 98.5% carbonate linkages, turnover frequencies up to 2.5 h -1 , <strong>and</strong> molecular<br />

weights reaching 100 000 gmol -1 . CHO–CO2 copolymerization produced PCHC with 100%<br />

carbonate linkages, molecular weights <strong>of</strong> 19000–330000 gmol -1 , <strong>and</strong> Mw/Mn’s <strong>of</strong> 3.5–12.5. It<br />

must be noted that control experiments indicated that ZnEt2, but not Y(CO2CF3)3, was<br />

essential for polymerization. Finally, ternary catalysts composed <strong>of</strong> Nd(CO2CCl3)3, ZnEt2,<br />

<strong>and</strong> glycerol were also reported to copolymerize PO <strong>and</strong> CO2. 37<br />

‐ 10 ‐


1.3 Polylactide<br />

1.3.1 Introduction<br />

- 11 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

Polylactide (PLA) is biodegradable, thermoplastic, aliphatic polyester derived from lactic<br />

acid, which is made from renewable resources, such as corn starch or sugarcanes. 38 PLA has<br />

similar mechanical properties to polyethylene terephthalate, but has a significantly lower<br />

maximum continuous use temperature. Poly(L-lactide) (PLLA) is the most important PLA. It<br />

is a semicrystalline polymer <strong>and</strong> has a crystallinity around 37%, a glass transition temperature<br />

~67 <strong>and</strong> a melting temperature ~180 . PLA can be processed like all other thermoplastic<br />

polymers with extrusion, injection molding, blow molding, or fiber spinning processes into<br />

various products. The products can be recycled after use either by remelting <strong>and</strong> processing<br />

the material a second time or they can be hydrolyzed into lactic acid, the basic chemical. The<br />

last possibility is to compost the polylactide to introduce it into the natural life cycle <strong>of</strong> all<br />

biomass, where it degrades into CO2 <strong>and</strong> water. PLA can be recycled following the traditional<br />

ways, composted like all other organic matter, <strong>and</strong> it will do no harm if burned in an<br />

incineration plant or introduced into a classical waste management system. PLA becomes <strong>of</strong><br />

commercial interest in recent years, in light <strong>of</strong> its biodegradability.<br />

1.3.2 <strong>Synthesis</strong> <strong>of</strong> PLA<br />

Figure 1.3.1 synthesis <strong>of</strong> PLA by the polycondensation <strong>of</strong> lactic acid.<br />

PLA can be prepared by direct polycondensation <strong>of</strong> lactic acid (Figure 1.3.1), which is<br />

produced from the bacterial fermentation <strong>of</strong> corn starch or cane sugar. 39 However, it is usually<br />

difficult to obtain high molecular weight polymer by polycondensation. Because each<br />

polymerization reaction generates one molecule <strong>of</strong> water, the presence <strong>of</strong> which degrades the<br />

forming polymer chain <strong>and</strong> results in the generation <strong>of</strong> low molecular weight PLA. So water<br />

has to be removed during the polycondensation in order to generate PLA with high molecular<br />

weight. Yamaguchi et al. has reported the synthesis <strong>of</strong> high molecular weight PLA by using<br />

molecular sieve to dry water during direct polycondensation. 40 However long reaction times<br />

<strong>and</strong> high temperatures are required in lactic acid polycondensation. 41


Chapter 1<br />

Polylactide <strong>and</strong> poly(lactic acid) are same chemical product. PLA is used as abbreviation<br />

for both poly(lactic acid) <strong>and</strong> polylactide. Lactide is the cyclic dimer <strong>of</strong> lactic acid, which is<br />

produced by means <strong>of</strong> a combined process <strong>of</strong> oligomerization <strong>and</strong> cyclization <strong>of</strong> lactic acid.<br />

Ring-opening polymerization (ROP) <strong>of</strong> lactide (Figure 1.3.2) is a preferred route to prepare<br />

PLA because <strong>of</strong> the higher controllability <strong>of</strong> the polymerization. 42 Indeed, high<br />

molecular-weight PLA can be obtained by ROP <strong>of</strong> lactide. 43 Because one lactide molecule<br />

has the two chiral carbon atoms, three stereoisomers <strong>of</strong> lactide can be generated (L, L-, D, D-,<br />

<strong>and</strong> meso-lactide) (Figure 1.3.3). The 1:1 mixture <strong>of</strong> L, L-, <strong>and</strong> D, D-lactide is<br />

racemic-lactide (rac-LA). Since lactide monomer is chiral, the control <strong>of</strong> PLA<br />

stereochemistry is easily achieved by the polymerization <strong>of</strong> L, L-, D, D- <strong>and</strong> meso-lactide<br />

stereochemical forms. Modulation <strong>of</strong> the polymer stereochemistry leads to PLA with<br />

dramatically different properties. For example, PLLA is a semicrystalline polymer (Tg at 67<br />

, melting transition at 180 ), while poly(rac-lactide) is an amorphous material (Tg at 58<br />

). The degree <strong>of</strong> crystallinity <strong>of</strong> PLA decreases dramatically as the L,L- content decreases<br />

over narrow compositional range from 100 to 92%. The degree <strong>of</strong> crystallinity <strong>of</strong> PLA with<br />

85% L,L- content was amorphous. 44<br />

Figure 1.3.2 <strong>Synthesis</strong> <strong>of</strong> PLA by the ROP <strong>of</strong> lactide.<br />

O<br />

O<br />

O<br />

O<br />

L,L-lactide<br />

O<br />

O<br />

O<br />

‐ 12 ‐<br />

O<br />

O<br />

O<br />

O<br />

O<br />

D,D-lactide meso-lactide<br />

Figure 1.3.3 Three stereoisomers <strong>of</strong> lactide.<br />

The driving force for the ROP <strong>of</strong> lactide is the ring strain <strong>of</strong> the monomer. As a<br />

polymerizable six-membered ring, the ring strain <strong>of</strong> lactide is modest. Therefore, high


- 13 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

termperatures are usually required for the ROP <strong>of</strong> lactides. 45 Consequently, a polymerizable<br />

activated monomer <strong>of</strong> the lactide equivalent is highly desirable.<br />

5-Methyl-1,3-dioxolane-2,4-dione is a five-membered O-carboxyanhydride (LacOCA)<br />

derived from the lactic acid. It can be readily obtained by the reaction <strong>of</strong> lactate salt with<br />

diphosgene (Figure 1.3.4). 46 Bourissou et al. reported that the ROP <strong>of</strong> LacOCA proceeded in<br />

the presence <strong>of</strong> an organocatalyst, dimethylaminopyridine (DMAP) 47, 48 or lipase. 49 The ROP<br />

<strong>of</strong> LacOCA catalyzed by DMAP is well controllable. PLA with high molecular weight <strong>and</strong><br />

narrow polydispersity index has been obtained under mild polymerization conditions. This<br />

indicates that the ROP <strong>of</strong> LacOCA is an alternative route to prepare PLA. 47<br />

Figure 1.3.4 <strong>Synthesis</strong> <strong>of</strong> LacOCA.<br />

1.3.3 Ring-opening polymerization <strong>of</strong> lactide <strong>and</strong> LacOCA<br />

Organometallic complexes are usually used as the catalyst <strong>of</strong> the ROP <strong>of</strong> lactide. For<br />

instance, stannous octoate, 45 alkylaluminum, aluminum alkoxides, 50-55 zinc alkyl, 56 calcium<br />

alkoxides, 57, 58 <strong>and</strong> strontium alkoxide 59 have been used as the catalyst/initiator for the ROP<br />

<strong>of</strong> lactides. Polymerization <strong>of</strong> lactide is usually assumed to proceed through a<br />

coordination-insertion mechanism. At first a complex between initiator <strong>and</strong> monomer is<br />

formed, followed by a rearrangement <strong>of</strong> the covalent bonds. The monomer is included in<br />

between the metal-oxygen bond <strong>of</strong> the initiator, cleaving the acyloxygen bond <strong>of</strong> the cyclic<br />

monomer, so that the metal is incorporated with an alkoxide bond into the growing chain<br />

(Figure 1.3.5). Among these complexes, stannous octoate [Sn(Oct)2] is most frequently used<br />

catalyst for the ROP <strong>of</strong> lactide because <strong>of</strong> its low toxicity <strong>and</strong> high thermal stability. The ROP<br />

reaction is carried out at ~120 ), because at the condition, transesterification reactions are<br />

virtually nonexistent, <strong>and</strong> retention <strong>of</strong> stereochemical purity during the conversion <strong>of</strong><br />

monomer to polymer is exceptional (99% ). 60 Even though the mechanism <strong>of</strong> ROP <strong>of</strong> lactide<br />

with Sn(Oct)2 is not yet clearly established, it is widely accepted that the ring-opening<br />

polymerization is actually initiated from compounds containing hydroxyl groups such as


Chapter 1<br />

water <strong>and</strong> alcohols, which are either present in the lactide feed or can be added by dem<strong>and</strong>.<br />

Although high molecular weight PLA can be obtained through the ROP <strong>of</strong> lactide, side<br />

reactions, such as intermolecular <strong>and</strong> intramolecular transesterification reactions exist <strong>and</strong><br />

perturb the chain propagation, broaden the molecular weight distribution, <strong>and</strong> yield cyclic<br />

oligomers. Therefore, catalyst that has lower side reactions is desired. Indeed, aluminum<br />

based single-site complexes have shown excellent controllability for the solution<br />

52, 54, 61-67<br />

polymerization <strong>of</strong> lactide.<br />

Figure 1.3.5 ROP <strong>of</strong> lactide through a coordination-insertion mechanism.<br />

DMAP 47 <strong>and</strong> lipase 49 can be used as the catalysts for the ROP <strong>of</strong> LacOCA. The<br />

ring-opening reactions <strong>of</strong> LacOCA with alcohols are predicted to occur through the activation<br />

<strong>of</strong> the alcohol <strong>and</strong> with both the traditional stepwise mechanisms, which involve tetrahedral<br />

intermediates (Figure 1.3.6). Furthermore, DMAP is proposed to act as a bifunctional catalyst<br />

through its basic nitrogen center <strong>and</strong> an acidic ortho-hydrogen atom. 48 The ROP <strong>of</strong> LacOCA<br />

using DMAP as catalyst gives access to PLA <strong>of</strong> controlled molecular weights <strong>and</strong> low<br />

polydispersities under mild conditions (typically within a few minutes at room temperature<br />

with LacOCA). 47 Both lipase PS <strong>and</strong> Novozym 435 promote the ring-opening<br />

polymerization <strong>of</strong> LacOCA (Figure 1.3.7). Accordingly, PLA <strong>of</strong> relatively high molecular<br />

weights <strong>and</strong> low polydispersities are obtained in high yields within a few hours at 80 .<br />

Slight preference for L-lacOCA over D-lacOCA is observed, <strong>and</strong> with Novozym 435, the<br />

molecular weight <strong>of</strong> the obtained PLA can be controlled by varying the lipase loading.<br />

‐ 14 ‐


1.3.4 Application <strong>of</strong> PLA<br />

Figure 1.3.6 DMAP-catalyzed ROP <strong>of</strong> LacOCA.<br />

Figure 1.3.7 Liphase-catalyzed ROP <strong>of</strong> LacOCA.<br />

- 15 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

Because PLA is biocompatible <strong>and</strong> bioresorbable, it can be used in a number <strong>of</strong><br />

biomedical applications, such as suture, implants, fracture fixation, drug delivery, <strong>and</strong> tissue<br />

engineering. 68-70 Because PLA is biodegradable, it can also be employed in the preparation <strong>of</strong><br />

bioplastic, useful for producing loose-fill packaging, compost bags, food packaging,<br />

disposable tableware, upholstery, disposable garments, awnings, feminine hygiene products,<br />

<strong>and</strong> nappies. As the depletion <strong>of</strong> petrochemical feedstocks draws near, PLA becomes an<br />

environmentally sustainable alternative to petrochemically-derived products because <strong>of</strong> the<br />

biodegradable characteristics <strong>and</strong> the renewable nature <strong>of</strong> its feedstock. 71<br />

1.4 Others <strong>Biodegradable</strong> Polyesters<br />

1.4.1 Poly(ε-caprolactone)<br />

Poly(ε-caprolactone) (PCL) is an aliphatic polyester composed <strong>of</strong> hexanoate repeat units.<br />

It has a low melting point <strong>of</strong> around 60 <strong>and</strong> a glass transition temperature <strong>of</strong> about -60 .<br />

The physical, thermal <strong>and</strong> mechanical properties <strong>of</strong> PCL depend on its molecular weight <strong>and</strong><br />

its degree <strong>of</strong> crystallinity. It is a semicrystalline polymer with a degree <strong>of</strong> crystallinity which<br />

can reach 69%. PCL shows the rare property <strong>of</strong> being miscible with many other polymers,<br />

such as poly(vinyl chloride), poly(styrene–acrylonitrile), poly(acrylonitrile butadiene styrene),<br />

poly(bisphenol-A), polycarbonates, nitrocellulose <strong>and</strong> cellulose butyrate. It is also<br />

mechanically compatible with many other polymers, such as polyethylene, polypropylene,<br />

natural rubber, poly(vinyl acetate), <strong>and</strong> poly(ethylene–propylene) rubber. 72 PCL biodegrades<br />

within several months to several years, depending on the molecular weight, the degree <strong>of</strong>


Chapter 1<br />

crystallinity, <strong>and</strong> the conditions <strong>of</strong> degradation. 73-75 PCL is degraded by hydrolysis <strong>of</strong> its ester<br />

linkages in physiological conditions. Therefore PCL has been used in many fields such as<br />

scaffolds in tissue engineering, in long-term drug delivery systems, in microelectronics, as<br />

adhesives, <strong>and</strong> in packaging. 76-81 PCL has been approved by Food <strong>and</strong> Drug Administration<br />

(FDA) to be used in the human body as a drug delivery device, suture, or adhesion barrier.<br />

PCL can be synthesized by the polycondensation <strong>of</strong> hydroxycarboxylic acids. 82 But high<br />

molecular weight PCL is usually made by the ROP <strong>of</strong> ε-caprolactone (ε-CL) that is<br />

industrially produced from the oxidation <strong>of</strong> cyclohexanone by peracetic acid (Figure 1.4.1).<br />

Three different catalytic systems, including metal-based, enzymatic, <strong>and</strong> organic systems, can<br />

59, 83-85<br />

be used to catalyze the polymerization <strong>of</strong> ε-CL.<br />

1.4.2 Polyglycolide<br />

Figure 1.4.1 Preparation <strong>of</strong> poly(ε-caprolactone)<br />

Polyglycolide or polyglycolic acid (PGA) is a biodegradable, thermoplastic polymer <strong>and</strong><br />

the simplest linear, aliphatic polyester. PGA has a glass transition temperature between 35-40<br />

<strong>and</strong> its melting point is reported to be in the range <strong>of</strong> 225-230 . The degree <strong>of</strong><br />

crystallinity <strong>of</strong> PGA is around 45-55%. 86 High molecular weight PGA is insoluble in almost<br />

all common organic solvents. However PGA is soluble in highly fluorinated solvents like<br />

hexafluoroisopropanol <strong>and</strong> hexafluoroacetone sesquihydrate. PGA was used to develop the<br />

first synthetic absorbable suture with the tradename <strong>of</strong> Dexon by the Davis & Geck<br />

subsidiary <strong>of</strong> the American Cyanamid Corporation in 1962. 87 It is naturally degraded in the<br />

body by hydrolysis <strong>and</strong> is absorbed as water-soluble monomers, completed between 60 <strong>and</strong><br />

90 days. PGA <strong>and</strong> its copolymers with lactic acid, ε-caprolactone, or trimethylene carbonate,<br />

are widely used as a material for the synthesis <strong>of</strong> absorbable sutures <strong>and</strong> are being evaluated<br />

in the biomedical field, such as implantable medical devices, tissue engineering or controlled<br />

drug delivery. 88 PGA can be prepared starting from glycolic acid by polycondensation or the<br />

ROP <strong>of</strong> glycolide. But the polycondensation yields a low molecular weight product. The most<br />

‐ 16 ‐


- 17 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

common synthesis route to produce a high molecular weight PGA is ROP <strong>of</strong> "glycolide", the<br />

cyclic diester <strong>of</strong> glycolic acid (Figure 1.4.2). Glycolide can be prepared by heating low MW<br />

PGA under reduced pressure, collecting the diester by means <strong>of</strong> distillation. ROP <strong>of</strong> glycolide<br />

can be catalyzed using different catalysts, including antimony, zinc, tin, aluminum, calcium,<br />

<strong>and</strong> lanthanide compounds. Stannous octoate is the most commonly used catalyst, since it is<br />

high efficient <strong>and</strong> approved by the FDA as a food stabilizer. 89<br />

Figure 1.4.2 Ring-opening polymerization <strong>of</strong> glycolide to produce polyglycolide.<br />

1.4.3 Poly(lactide-co-glycolide)<br />

. Using the PGA <strong>and</strong> PLA properties as a starting point, it is possible to copolymerize the<br />

two monomers to extend the range <strong>of</strong> homopolymer properties. Copolymers <strong>of</strong> glycolide with<br />

lactide (PLGA) have been developed for medical device, tissue engineering <strong>and</strong> drug delivery<br />

applications. 90-92 It is important to note that there is not a linear relationship between the<br />

copolymer composition <strong>and</strong> the mechanical <strong>and</strong> degradation properties <strong>of</strong> the materials. For<br />

example, a copolymer <strong>of</strong> 50% glycolide <strong>and</strong> 50% D,L-lactide degrades faster than either<br />

homopolymer. 93 PLGA with 25-70% glycolide are amorphous due to the disruption <strong>of</strong> the<br />

regularity <strong>of</strong> the polymer chain by the other monomer. PLGA <strong>of</strong> 90% glycolide <strong>and</strong> 10%<br />

lactide was developed by Ethicon as an absorbable suture material under the trade name<br />

Vicryl. It absorbs within 3-4 months but has a slightly longer strength-retention time. 93 PLGA<br />

is produced by the copolymerization <strong>of</strong> lactide <strong>and</strong> glycolide under ring-opening<br />

polymerization catalyst (Figure 1.4.3).<br />

Figure 1.4.3 <strong>Synthesis</strong> <strong>of</strong> poly(lactide-co-glycolide).


Chapter 1<br />

1.4.4 Polyhydroxybutyrate<br />

Polyhydroxybutyrate (PHB) is a biodegradable thermoplastic polymer produced by<br />

micro-organisms. 94 The polymer is primarily a product <strong>of</strong> carbon assimilation <strong>and</strong> is<br />

employed by micro-organisms as a form <strong>of</strong> energy storage molecule to be metabolized when<br />

other common energy sources are not available. The poly(3-hydroxybutyrate) (P3HB) (Figure<br />

1.4.4) form <strong>of</strong> PHB is probably the most common type <strong>of</strong> polyhydroxyalkanoate. PHB has a<br />

melting point <strong>of</strong> 175 <strong>and</strong> glass transition temperature <strong>of</strong> 15 . The tensile strength is 40<br />

MPa, which is close to that <strong>of</strong> polypropylene. 95 PHB is water insoluble <strong>and</strong> relatively<br />

resistant to hydrolytic degradation, which differentiates PHB from most other currently<br />

available biodegradable plastics, such as PLA, PCL, <strong>and</strong> PGA, which are either water soluble<br />

or moisture sensitive. However ICI had developed the material to pilot plant stage in the<br />

1980s, but interest faded when it became clear that the cost <strong>of</strong> material was too high.<br />

1.5 Electroactive Polymers<br />

Figure 1.4.4 Chemical structure <strong>of</strong> poly(3-hydroxybutyrate).<br />

Recent advances in electroactive polymers have aroused more <strong>and</strong> more interests in their<br />

application as biomedical materials. These researches based on the theory that a multitude <strong>of</strong><br />

cell functions, such as attachment, proliferation, migration, <strong>and</strong> differentiation could be<br />

modulated through electrical stimulation 96 , which have been demonstrated for a long time.<br />

Presently the electroactive polymers used are all conducting polymers, which have electrical<br />

<strong>and</strong> optical properties similar to inorganic semiconductors or metals. 97 Moreover, they also<br />

exhibit the unique properties as conventional polymers, including ease <strong>of</strong> synthesis <strong>and</strong><br />

processing. These attractive properties have given these polymers a wide range <strong>of</strong><br />

applications in the biological field, such as the matrix for tissue engineering or drug carrier<br />

for controlled release. 98, 99 Although there have been many conducting polymers discovered<br />

<strong>and</strong> characterized at the past 30 years, a small number <strong>of</strong> them have been developed for the<br />

‐ 18 ‐


- 19 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

biomedical applications, including poly(pyrrole)s, polyanilines <strong>and</strong> polythiophenes. Thus,<br />

developing new type <strong>of</strong> electroactive polymers for biomedical application is still necessary<br />

both in theory <strong>and</strong> in practice.<br />

1.5.1 Introduction <strong>of</strong> Electroactive Polymers<br />

1.5.1.1 Poly(acetylene)<br />

Poly(acetylene) is the simplest conjugated polymer <strong>and</strong> historically the discovery <strong>of</strong> the<br />

metallike conductivity <strong>of</strong> poly(acetylene) doping triggered the explosion <strong>of</strong> the research to<br />

electroactive polymers (Figure 1.5.1A). Poly(acetylene) can exist in two isomeric forms:<br />

cis-transoid <strong>and</strong> trans-transoid, commonly called cis- <strong>and</strong> trans-poly(acetylene), respectively.<br />

The latter form being thermodynamically stable since cis to trans isomerization is irreversible.<br />

Polyacetylene is one <strong>of</strong> the most promising materials for applications in optoelectronics. 100<br />

The conductivity <strong>of</strong> this polymer after doping is equal to that <strong>of</strong> copper, <strong>and</strong> some forms <strong>of</strong><br />

polyacetylene have record values <strong>of</strong> non-linear third-order optical susceptibility. Because it<br />

contains high concentrations <strong>of</strong> unsaturated sites which can be easily attacked by ozone <strong>and</strong><br />

UV light, polyacetylene is very unstable to the environment. 101 The low stability <strong>of</strong><br />

polyacetylene has been the major obstacle in the way <strong>of</strong> practical applications <strong>of</strong> this polymer.<br />

An effective approach to obtain stable poly(acetylene) without weakening its conductivity<br />

was blend poly(acetylene) with processable plastics.<br />

Figure 1.5.1 Structure <strong>of</strong> poly(acetylene) (A), polypyrrole (B) <strong>and</strong> poly(thiophene) (C).


Chapter 1<br />

1.5.1.2 Polypyrrole (PPy)<br />

Among conducting polymers, polypyrrole is especially promising for commercial<br />

applications because <strong>of</strong> its good environmental stability, facile synthesis, <strong>and</strong> higher<br />

conductivity than many other conducting polymers (Figure 1.5.1B). PPy can be easily<br />

prepared by either a chemical or electrochemical polymerization <strong>of</strong> pyrrole. However, the<br />

synthetically conductive PPy is insoluble <strong>and</strong> infusible. To improve the processibility, many<br />

approaches have been developed. For example, several kinds <strong>of</strong> soluble PPy have been<br />

synthesized, such as poly (3-alkylpyrrole) with an alkyl group equal to or greater than a butyl<br />

group. 102 Poly (3-alkylpyrrole) is easily soluble in common solvents or soluble in water when<br />

the substituents bear hydrophilic groups, such as –SOH3. However, poly (N-substituted<br />

pyrroles) has much lower conductivity due to greatly suppressed conjugation along the<br />

103, 104<br />

polymer chains by the substituents on nitrogen.<br />

1.5.1.3 Poly(thiophene) (PT)<br />

Like polypyrrole, poly(thiophene) is also an important poly(heterocyc1es) (Figure 1.5.1C).<br />

From a theoretical viewpoint, PT has been <strong>of</strong>ten considered as a model for the study <strong>of</strong><br />

charge transport in conductive polymers with a nondegenerate ground state, while on the<br />

other h<strong>and</strong>, the high environmental stability <strong>of</strong> both its doped <strong>and</strong> undoped states has led to<br />

its wide applications. 105 PT is essentially prepared by means <strong>of</strong> two main routes: the chemical<br />

<strong>and</strong> the electrochemical syntheses. Although it is likely that chemical syntheses are the most<br />

adequate methods to prepare oligomers with defined structure, until now, the most<br />

extensively conjugated <strong>and</strong> most conductive PT have been prepared by electrochemical<br />

polymerization. Unsubstituted PT is conductive after doping, but is intractable <strong>and</strong> soluble<br />

only in solutions like mixtures <strong>of</strong> arsenic trifluoride <strong>and</strong> arsenic pentafluoride. 106 However,<br />

examples <strong>of</strong> organic-soluble PT were reported when it was substituted. Nickel-catalyzed<br />

Grignard cross-coupling was used to synthesize two soluble PTs, poly(3-butylthiophene) <strong>and</strong><br />

poly(3-methylthiophene-co-3'-octylthiophene), which could be cast into films <strong>and</strong> doped with<br />

iodine to reach conductivities <strong>of</strong> 4 to 6 S/cm. 107 Hotta et al. synthesized<br />

poly(3-butylthiophene) <strong>and</strong> poly(3-hexylthiophene) electrochemically 108 , <strong>and</strong> characterized<br />

the polymers in solution <strong>and</strong> cast into films. 109 The soluble PATs demonstrated both<br />

thermochromism <strong>and</strong> solvatochromism in chlor<strong>of</strong>orm <strong>and</strong> 2,5-dimethyltetrahydr<strong>of</strong>uran. 110<br />

‐ 20 ‐


1.5.1.4 Polyaniline (PANi)<br />

- 21 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

Polyaniline is an intrinsically conducting polymer consisting <strong>of</strong> phenyldiamine <strong>and</strong><br />

quinodiimine units, the general formula can be shown as Figure 1.5.2. Polyaniline is widely<br />

studied due to its ease <strong>of</strong> synthesis, environmental stability, <strong>and</strong> simple doping/dedoping<br />

chemistry. The synthesis approaches <strong>of</strong> PANi were various but the most common method was<br />

based on mixing aqueous solutions <strong>of</strong> aniline hydrochloride <strong>and</strong> ammonium peroxydisulfate<br />

at room temperature, followed by the separation <strong>of</strong> PANi hydrochloride precipitate by<br />

filtration <strong>and</strong> drying. 111 Electrochemical synthesis is also a common alternative for preparing<br />

PANi, particularly because this synthetic procedure is relatively straightforward. It was found<br />

that the structure <strong>of</strong> PANi <strong>and</strong> its physical <strong>and</strong> chemical properties strongly dependent on<br />

their preparation <strong>and</strong> doping methods. Therefore, to explore new synthesis <strong>and</strong> doping<br />

method is necessary to exp<strong>and</strong> the application extent <strong>of</strong> PANi. In recent years, a number <strong>of</strong><br />

novel polymerization methods were developed, such as emulsion polymerization,<br />

microemulsion polymerization <strong>and</strong> template polymerization. 112<br />

Figure 1.5.2 Structure <strong>of</strong> polyaniline.<br />

1.5.2 Electroactive Polymers for Biomedical Applications<br />

Electroactive polymers have been widely applied in the microelectronics industry,<br />

including battery technology, light emitting, diodes photovoltaic devices, <strong>and</strong> electrochromic<br />

displays. 113 Based on the concept that cellular activities can be influenced by electricity,<br />

electroactive polymers were found to be unique in modulating <strong>of</strong> the cell adhesion, migration,<br />

DNA synthesis, <strong>and</strong> protein secretion. 114-118 Most <strong>of</strong> the researches focused on the tissues or<br />

cells which respond to electrical impulses, including nerve, muscle, bone, <strong>and</strong> cardiac cells.<br />

Besides the ability <strong>of</strong> transferring charge from a biochemical reaction, most <strong>of</strong> electroactive<br />

polymers such as PPy <strong>and</strong> PT are biocompatible, which is an important property <strong>of</strong><br />

biomaterials. These unique characteristics made them useful in many biomedical applications,<br />

such as tissue engineering materials, biosensors, drug delivery devices <strong>and</strong> bioactuators.


Chapter 1<br />

1.5.2.1 Tissue engineering applications<br />

The electroactive polymer, being electrical stimulus responsive, has become the focus <strong>of</strong><br />

research after the demonstration that electrical signals can regulate cell attachment,<br />

proliferation <strong>and</strong> differentiation. 96 Lots <strong>of</strong> efforts had been made to use the electroactive<br />

polymers to regulate the proliferation <strong>and</strong> differentiation <strong>of</strong> cells. And they had been widely<br />

used in biological systems due to the key advantage <strong>of</strong> facile control <strong>of</strong> the intensity <strong>and</strong><br />

duration <strong>of</strong> stimulation.<br />

PPy was one <strong>of</strong> the first electroactive polymers studied for its effect on cells. And it has<br />

been demonstrated to support adhesion <strong>and</strong> growth <strong>of</strong> a number <strong>of</strong> different cell types. The<br />

investigation <strong>of</strong> biocompatibility <strong>of</strong> PPy was just the first step. Its most attractive property is<br />

electroactivity.<br />

Figure 1.5.3 Photomicrograph <strong>of</strong> endothelial cells cultured on fibronectin-coated PPy doped with<br />

p-toluene sulfonate (PPyTS) for 4 h: (A) PPy in native oxidized state <strong>and</strong> (B) PPyTs reduced by<br />

application <strong>of</strong> -0.5V for 4 h(Magnify: ×700). 116<br />

Some reports have shown the possibility <strong>of</strong> using PPy to modulate cellular response via<br />

electrical stimulation. Wong et al. assessed the suitability <strong>of</strong> electroactive polymers for<br />

sustaining cell growth <strong>and</strong> controlling cell function. 116 In their study, aortic endothelial cells<br />

were cultured on fibronectin (FN)-coated PPy films which were in different oxidation state<br />

(Figure 1.5.3). It was found that neutral state PPy caused cell rounding <strong>and</strong> a concomitant<br />

drop (~98%) in DNA synthesis, <strong>and</strong> the oxidation <strong>of</strong> PPy resulted in cell spreading.<br />

Interestingly, cell viability (>90%) <strong>and</strong> adhesion were very good on both oxidized <strong>and</strong> neutral<br />

PPy despite the change in cellular morphology. Schmidt et al. have also made great<br />

‐ 22 ‐


- 23 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

contribution to the application <strong>of</strong> PPy in the biomedical field. They reported that electrical<br />

stimulation <strong>of</strong> PPy in its oxidized form can also be used to modulate cell function. 119 In their<br />

experiment, rat pheochromocytoma cell line (PC-12) cells on poly(styrene sulfonate)<br />

(PSS)-doped PPy films were found to exhibit a ~91% increase in median neurite length in the<br />

presence <strong>of</strong> a direct current lasted for 2 h (Figure 1.5.4). These studies demonstrate that the<br />

electrical stimulus can significantly enhance PC-12 neurite outgrowth <strong>and</strong> spreading.<br />

Moreover, the reason why PPy could enhance the cell extension was studied <strong>and</strong> they<br />

concluded that the electrical stimulation increases the adsorption <strong>of</strong> serum proteins, which<br />

help to improve the growth <strong>and</strong> proliferation <strong>of</strong> cells. Subsequently, Lakard et al., 120 George<br />

et al., 121 <strong>and</strong> several other groups investigated adhesion <strong>and</strong> proliferation <strong>of</strong> cell cultured<br />

different cell lines on PPy. 122-124<br />

Figure 1.5.4 PC-12 cell differentiation on poly(styrene sulfonate)-doped PPy films. (A) Cells<br />

grown for 48 h but not subjected to electrical stimulation are shown for comparison. (B) PC-12<br />

cells were grown on PPy for 24 h in the presence <strong>of</strong> nerve growth factor, then exposed to<br />

electrical stimulation (100 mV) across the polymer film. Images were acquired 24 h after<br />

stimulation. Scale bar = 100 μm. 119<br />

The limitation <strong>of</strong> PPy as a favorable biomaterial is that it lacks biological activity. In order<br />

to endow it bioactivity, many biomolecules, such as adenosine 5'-triphosphate (ATP) 125-127<br />

<strong>and</strong> nerve growth factor (NGF) 128 were entrap in PPy. The introduction <strong>of</strong> biologically active<br />

molecules endows the material new properties. Thus, the resulted materials can modulate the<br />

growth <strong>of</strong> different cell types simply by varying the biomolecules. For example, Collier et al.


Chapter 1<br />

synthesized hyaluronic acid (HA)-doped PPy <strong>and</strong> explored its tissue engineering<br />

applications. 129 In vivo studies showed that the introducing <strong>of</strong> HA into PPy films improved the<br />

biocompatibility <strong>of</strong> PPy <strong>and</strong> promoted the vascularization.<br />

It should be noted that both conductivity <strong>and</strong> the film surface are <strong>of</strong>ten greatly changed<br />

when biologically active molecules are used as dopants. 129, 130 For example, the morphology<br />

<strong>and</strong> conductivity <strong>of</strong> polypyrrole-based films with different dopants were systematically<br />

studied. It was concluded that larger biomolecule dopants can significantly decrease<br />

conductivity <strong>and</strong> affect more obviously on morphology compared to small dopants. However,<br />

the exact effects <strong>of</strong> particular dopants are not very predictable (Figure 1.5.5).<br />

Figure 1.5.5 Morphology <strong>of</strong> thick PPy films as assessed using scanning electron microscopy<br />

(SEM) (top) <strong>and</strong> corresponding cyclic voltammograms (CVs), indicating electrical activity <strong>of</strong> the<br />

materials (bottom). (A) PPyCl. (B) PPy doped with poly(vinyl sulfate). (C) PPy doped with<br />

dermatan sulfate. (D) PPy doped with collagen (inset: thin film <strong>of</strong> collagen-doped PPy). A narrow<br />

‐ 24 ‐


- 25 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

CV spectrum correlates to decreased electroactivity. Scale bars are 100 μm. 130<br />

PPy can not be further covalently modified due to there are no functional pendent or<br />

terminal groups on its chain. This limitation can be overcome by doping with polymers<br />

having functional groups. A good example <strong>of</strong> this technique is conducted by Song et al. In<br />

their report, PPy was doped with poly(glutamic acid) (PGlu), which has pendent carboxylic<br />

acid groups could further react with amino <strong>of</strong> polylysine <strong>and</strong>/or laminin. Dorsal root ganglion<br />

(DRG) neurons were observed to preferentially adhere to <strong>and</strong> extend neurites on the PPy<br />

surfaces modified with polylysine <strong>and</strong>/or laminin (Figure 1.5.6). 130<br />

Figure 1.5.6 Phase contrast <strong>and</strong> fluorescence images <strong>of</strong> DRGs adhered to the surface <strong>of</strong> pPy-X,<br />

where X = 1. pLys, 2. Lmn, <strong>and</strong> 3. pLys-Lmn. Cell nuclei were labeled with DAPI (blue<br />

florescence). All images were taken at 10× magnification <strong>and</strong> the 200 μm scale bar applies to all<br />

images shown. 130<br />

PPy is not biodegradable in nature <strong>and</strong> many works were carried out to blend it with<br />

biodegradable polymeric materials to obtain partially absorbable materials. For example, PPy<br />

nanoparticle was blended with PLA to fabricate a composite which is both biodegradable <strong>and</strong><br />

conductive. 117 Altering the content <strong>of</strong> PPy in the composites will result in the changes <strong>of</strong> both


Chapter 1<br />

parameters. Moreover, the morphology <strong>of</strong> the composites will also be changed with different<br />

PPy content (Figure 1.5.7). The fibroblasts exhibited improved attaching <strong>and</strong> growth on the<br />

PPy nanoparticle-PLA composites membranes under the simulated <strong>of</strong> current compared to<br />

that cultured without simulation. The composites provided a new way to design electroactive,<br />

biocompatible <strong>and</strong> biodegradable materials with simple blending approach.<br />

Figure 1.5.7 SEM images <strong>of</strong> the surface <strong>of</strong> PPy nanoparticle-PLA composite membranes with<br />

various PPy content: (A) 5%, (B) 7%, (C) 9%, <strong>and</strong> (D) 17%. 117<br />

Many cells, especially nerve cell, have filaceous parts which will exhibit different<br />

topography. And the topography <strong>of</strong> culture substrate will influence the growth <strong>and</strong><br />

differentiation <strong>of</strong> these cells. Some studies have explored the effect <strong>of</strong> surface topography <strong>of</strong><br />

electroactive materials on cell culture. For instance, patterning <strong>of</strong> 1 <strong>and</strong> 2 μm wide PPyPSS<br />

microchannels can be fabricated using electron beam lithography. 131 Then the embryonic<br />

hippocampal neurons were seeded on this matrix <strong>and</strong> found to polarize (i.e., define an axon)<br />

faster, exhibiting a two-fold increase in the number <strong>of</strong> cells with axons compared to cells<br />

cultured on unmodified PPyPSS (Figure 1.5.8). For instance, patterning <strong>of</strong> 1 <strong>and</strong> 2 μm wide<br />

PPyPSS microchannels can be fabricated using electron beam lithography. 131 Then the<br />

embryonic hippocampal neurons were seeded on this matrix <strong>and</strong> found to polarize (i.e., define<br />

an axon) faster, exhibiting a two-fold increase in the number <strong>of</strong> cells with axons compared to<br />

cells cultured on unmodified PPyPSS (Figure 1.5.8).<br />

‐ 26 ‐


- 27 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

Figure 1.5.8 Patterning <strong>of</strong> PPy to create microchannels for contact guidance <strong>of</strong> neurons.<br />

Phase-contrast (left) <strong>and</strong> fluorescence (right) photomicrographs <strong>of</strong> hippocampal neurons on PPy.<br />

(A, B) Cells cultured on 2 mm wide <strong>and</strong> 200nm deep PPy microchannels. (C, D) Cells cultured on<br />

unmodified PPy. The green labeling (Alexa 488) corresponds to Tau-1 (axonal marker)<br />

immunostaining. Cells polarized (i.e., established a single axon) more readily on microchannels<br />

than on unmodified PPy. Scale bar = 20 μm. Images are at the same magnification. 131<br />

PANi was another conductive polymer that had been explored for applications as novel<br />

intelligent scaffolds for cardiac <strong>and</strong>/or neuronal tissue engineering. As compared to PPy, the<br />

investigation <strong>of</strong> PANi as biomaterials developed more slowly. However, it is attracting more<br />

<strong>and</strong> more attentions due to the pr<strong>of</strong>ound underst<strong>and</strong>ing on its electrical properties. For<br />

example, Mattioli-Belmonte et al. first demonstrated that PANi was biocompatible in vitro<br />

<strong>and</strong> in long term animal studies in vivo. 132 Later, Wei et al. reported that PANi films<br />

functionalized with the bioactive lamimin-derived adhesion peptide YIGSR (Tyr-Ile-Gly<br />

Ser-Arg) exhibited significant enhanced PC-12 cell attachment <strong>and</strong> differentiation. 133


Chapter 1<br />

Figure 1.5.9 (A) SEM image <strong>of</strong> PANi-gelatin blend fibers with ratio 45:55. Original<br />

magnification is 5000×. (B) Morphology <strong>of</strong> H9c2 myoblast cells at 20 h post-seeding on 45:55<br />

PANi-gelatin blend fiber. Staining is for nuclei-bisbenzimide <strong>and</strong> actin cytoskeletonphalloidin;<br />

fibers aut<strong>of</strong>luoresce. Original magnification is 400×. (C) SEM images <strong>of</strong> H9c2 cells cultured on<br />

45:55 PANI-gelatin blend fibers. 134<br />

Despite the progress <strong>of</strong> PNAi in tissue engineering application, there was still a lot <strong>of</strong> work<br />

to do to improve their poor solubility, the poor polymer-cell interaction <strong>and</strong> biodegradability.<br />

Therefore, it is necessary to design novel electroactive polymers with good solubility,<br />

biocompatibility <strong>and</strong> biodegradability for the tissue engineering application. For example,<br />

polyanline was blended with natural polymers such as collagen <strong>and</strong> gelatin or covalently<br />

grafted with oligopeptides such as Ty-Ile-Gly-Ser-Arg (YIGSR) to improve its polymer-cell<br />

133, 134<br />

interaction (Figure 1.5.9).<br />

Figure 1.5.10 (A) Phase contrast images <strong>of</strong> PC-12 cell morphology <strong>of</strong> (a) TCP, (b) TCP with<br />

NGF, (c) ATQD-RGD,<strong>and</strong> (d) ATQD-RGD with NGF on day 10 (B)Neurite length distribution<br />

chart for ATQD-RGD substrates with <strong>and</strong> without NGF.<br />

‐ 28 ‐


- 29 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

The solubility <strong>of</strong> polyaniline could be enhanced by covalently grafting side groups or<br />

polymers, such as poly(ethylene glycol), on the backbone <strong>of</strong> polyaniline. 135 Wei et al. 136<br />

demonstrated that the electroactive silsesquioxane precursor,<br />

N-(4-aminophenyl)-N'-(4'-(3-triethoxysilyl-propyl-ureido) phenyl-1,4- quinonenediimine)<br />

(ATQD), containing aniline trimer covalently modified by oligopeptide could be a kind <strong>of</strong><br />

promising biomaterial for tissue engineering. The bioactive material, ATQD-RGD, supported<br />

PC-12 cell adhesion <strong>and</strong> proliferation, <strong>and</strong> stimulated spontaneous neuritogenesis in PC-12<br />

cells in the absence <strong>of</strong> neurotrophic growth factors (NGF), as shown in Figure 1.5.10.<br />

Figure 1.5.11 (A) Representational fluorescence micrographs <strong>of</strong> PC-12 cell for the substrates (a)<br />

TCPS (-) without electrical stimulation, (b) TCPS (+) exposed to electrical stimulation, (c) EM<br />

PLAAP (-) doped with CSA without electrical stimulation, (d) EM PLAAP (+) doped with CSA<br />

exposed to electrical stimulation on day 4. (B) The mean neurite length <strong>of</strong> PC-12 cells cultured<br />

on the substrates <strong>of</strong> EM PLAAP (-), TCPS (+), <strong>and</strong> EM PLAAP (+) on day 4. 136<br />

Based on the above works, Chen’s group did many works to exp<strong>and</strong> the architecture <strong>of</strong> this<br />

kind <strong>of</strong> materials. They chose aniline oligomers (especially aniline pentamer with dicarboxyl<br />

end group (AP)) as electroactivity resource, which incorporated with degradable polymers,<br />

such as polylactide (PLA), poly(ε-caprolactone) (PCL) <strong>and</strong> natural biopolymer chitosan, to<br />

prepare new biodegradable electroactive biomaterials. 137-139 The introducing <strong>of</strong> PLA<br />

endowed the material with good electroactivity, solubility <strong>and</strong> biodegradability similar to<br />

pure PLA. In vitro cell evaluation showed that the electroactive copolymers could indeed<br />

promote the attachment <strong>and</strong> growth <strong>of</strong> rat C6 glioma cells. Moreover, in the comparison


Chapter 1<br />

experiments with <strong>and</strong> without applying electrical potentials, the dopped electroactive<br />

copolymers had the ability <strong>of</strong> improving the differentiation <strong>of</strong> PC-12 cells, as shown in Figure<br />

1.5.11. They also prepared a new kind <strong>of</strong> water-soluble electroactive polymer, aniline<br />

pentamer cross-linked chitosan. These new polymers showed good electroactivity even in<br />

aqueous solution. The MTT assay, cell adhesion test, <strong>and</strong> degradation assessment in the<br />

presence <strong>of</strong> enzyme confirmed that these polymers had good biocompatibility <strong>and</strong><br />

biodegradability. The electroactive polymers can obviously improve the neuronal<br />

differentiation <strong>of</strong> PC-12 cells even without the extra electrical stimulation, as shown in<br />

Figure1.5.12.<br />

Figure 1.5.12 Visualization <strong>of</strong> PC-12 neurite outgrowth by micrographs for the substrates (A)<br />

without electroactivity (chitosan), (B) with electroactivity (aniline pentamer cross-linked<br />

chitosan) on day 5. 139<br />

A B<br />

As the study progressing, Chen et al. found that the oligomers without high conductivity<br />

<strong>and</strong> the polymers containing oligomers also showed improvement in the C6 cell<br />

differentiation in the absence <strong>of</strong> electrical stimulation, as shown in Figure 1.5.13. In the<br />

culture medium, the only difference from the electroactive polymers may be the exchange <strong>of</strong><br />

the ion between the medium with polymer, <strong>and</strong> between the polymer with cells, which means<br />

the electroactivity changed the ion exchange between the cells <strong>and</strong> medium.<br />

‐ 30 ‐


- 31 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

Figure 1.5.13 Visualization <strong>of</strong> C-6 outgrowth by micrographs in the culture medium (a), (b) in<br />

the presence <strong>of</strong> aniline pentamer cross-linked chitosan, (c) <strong>and</strong> (d) without aniline pentamer<br />

cross-linking chitosan.<br />

Besides the extensively investigated PPy <strong>and</strong> PANi, several electroactive polymers have<br />

been also studied, such as polythiophene (PT) <strong>and</strong> others. For example, a series <strong>of</strong> polymers<br />

composed <strong>of</strong> a PT backbone with oligosiloxane grafted to the b position <strong>of</strong> the thiophene<br />

rings was prepared. 140 All the polymers exhibited different conductivities after doped with<br />

iodine. However, the iodine is toxic to cells <strong>and</strong> the authors only investigated the<br />

biocompatibility <strong>of</strong> undoped polymers. The carbon-carbon bond <strong>of</strong> PT made it inherently<br />

nonbiodegradable. To overcome this drawback, some reports concentrated on the synthesis <strong>of</strong><br />

biodegradable elcetroactive polymers containing heterocycles unit. For example, ester linkers<br />

were introduced to connect the conductive heterocycles (Figure. 1.5.14A). 96 The resulted<br />

polymer is conductive, <strong>and</strong> more importantly, it is biodegradable in the presence <strong>of</strong> esterases<br />

which made it be applicable in vivo. The conductivity <strong>of</strong> the polymer is mainly attributed to<br />

inter <strong>and</strong> intra “electron hopping” formed by the overlap <strong>of</strong> conjugated regions. However, the<br />

doping agent is also iodine. Despite this drawback, the undoped polymer was biocompatible<br />

in vitro. Nerve cells adhered to polymer film <strong>and</strong> readily expressed their nerve-like phenotype<br />

by extending neurites after one day (Figure 1.5.14B, left). After 8 days, significant cell<br />

proliferation was observed (Figure 1.5.14B, right). These data demonstrate that in addition to<br />

being non-toxic to cells in culture, this biodegradable electroactive polymer can also support<br />

cell attachment <strong>and</strong> proliferation. Subsequently, in vivo biocompatibility <strong>of</strong> this polymer was


Chapter 1<br />

characterized by subcutaneous implantation into rats for 14 <strong>and</strong> 29 days using FDA-approved<br />

PLGA as a control (Figure 1.5.14C). The results showed that inflammation around<br />

electroactive polymer is mild <strong>and</strong> similar to that seen with PLGA at 29 days. Thus, there<br />

appears to be no detectable toxic effect <strong>of</strong> the base material or its degraded products in vivo.<br />

Figure 1.5.14 (A) Chemical structures <strong>of</strong> biocompatible electroactive polymer. (B) Human<br />

neuroblastoma cells cultured in vitro on electroactive polymer films after 1 day (left) <strong>and</strong> s 8 days<br />

(right). Both images are at the same magnification. Scale bar = 100 μm. (C) Histological tissue<br />

sections (stained with hematoxylin <strong>and</strong> eosin) <strong>of</strong> Electroactive polymer (left) <strong>and</strong> PLGA (right)<br />

demonstrated comparably low inflammatory responses at 29 days. Both images are at the same<br />

magnification. Scale bar = 50 μm. 96<br />

1.5.2.2 Biosensor Applications<br />

Most <strong>of</strong> the electroactive polymers used for biosensors are conjugated polymers (CPs),<br />

which can couple analyte receptor interactions, as well as nonspecific interactions, into<br />

‐ 32 ‐


- 33 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

observable responses. A key advantage <strong>of</strong> CP-based sensors is the potential <strong>of</strong> the CP to<br />

exhibit collective properties that are sensitive to very minor chemical or physical changes.<br />

The most important aspect <strong>of</strong> biosensor is how to combine the electrical component (i.e., CP)<br />

with the biological recognition components. An effective <strong>and</strong> widely used approach was to<br />

immobilize bioactive molecules in or on CPs. 141-143<br />

Physical adsorption is the simplest way to introduce bioactive molecules on CPs for<br />

biosensors. For example, glucose oxidase was directly adsorbed onto PPy for a biosensor<br />

which can detect glucose from concentrations <strong>of</strong> 2.5 to 30 mM using dimethylferrocene as an<br />

electron transfer mediator. 144 However, this method is not reliable because the amount <strong>of</strong> the<br />

absorbed compound is not repeatable <strong>and</strong> this immobilization is not stable. 145 An alternative<br />

way is to entrap the biomolecule into the polymers substrate. For example, poly(3,<br />

4-ethylenedioxythiophene) (PEDOT) exhibited a 5% shrinkage in thickness when it is rinsed<br />

with ethanol, which can be utilized to incorporate horseradish peroxidase. 146 Besides enzymes,<br />

this method has also been used for immobilization <strong>of</strong> antibodies <strong>and</strong> DNA. 147-149 Just like<br />

the absorption method, this approach also has some limitations such as denature <strong>of</strong> the<br />

proteins <strong>and</strong> the low accessibility <strong>of</strong> analytes to the sensing element. Moreover, entrapment<br />

methods need a high concentration <strong>of</strong> the biomolecule (~0.2-3.5 mg/mL), which is not<br />

suitable for industry application due to the higher cost.<br />

Compared to entrapment <strong>and</strong> absorption, affinity binding seemed to be more stable<br />

bonding techniques. The avidin-biotin complex was the most studied affinity binding pair due<br />

to the extremely specific <strong>and</strong> high-affinity interactions between biotin <strong>and</strong> the glycoprotein<br />

avidin (Ka = 1 × 10 15 mol -1 L). 150 This method requires the synthesis <strong>of</strong> biotinylated CPs<br />

firstly. For example, the pyrrole monomers was copolymerized with biotinylated hydrophilic<br />

pyrrole monomers with different PEG lengths as spacer arms to create the bonding site. 151<br />

And the amount <strong>of</strong> anchored avidin can be tuned by the biotin density or the length <strong>of</strong> the<br />

PEG spacer in the copolymer. There are also other affinity binding complexes that have been<br />

studied for the immobilization <strong>of</strong> DNA onto CP surfaces. For example, a unique intercalator<br />

(i.e., a molecule that inserts itself into double-str<strong>and</strong>ed DNA)-based immobilization technique<br />

has been developed. 152 Target DNA str<strong>and</strong>s are detected when they form DNA duplexes with<br />

labeled DNA probes <strong>and</strong> are subsequently immobilized by the affinity binding <strong>of</strong> the<br />

intercalator, which is covalently bound to PPy. And more importantly, the detection limits can<br />

lower to 1 pg/mL.


Chapter 1<br />

Covalent modification was another effective way to prepare stable detective component.<br />

Conducting copolymers containing covalently substituted monomers have been considered as<br />

an effective way to immobilize biomolecules. For example, copolymers <strong>of</strong> PPy <strong>and</strong><br />

oligonucleotides bearing a pyrrole group have been reported for DNA sensors. 153, 154 An<br />

easier method is to prepare copolymers bearing functional groups such as amine or carboxylic<br />

acid groups, which can further react with biomolecules. Minett et al. designed an<br />

immunosensors for detecting Listeria monocytogenes by covalently binding the Listeria<br />

monoclonal antibody to a copolymer <strong>of</strong> carboxylic acid-functionalized PPy <strong>and</strong> regular<br />

PPy. 155, 156 In another example, N-Hydroxysuccinimide was introduced as a pendent group <strong>of</strong><br />

the PT or PPy (Figure 1.5.15). 152, 157-159 This activated ester group can easily react with<br />

amines in proteins <strong>and</strong> other biomolecules under mild conditions to prepare detection<br />

materials.<br />

Figure 1.5.15 Immobilization <strong>of</strong> DNA on PPy derivatives with N-hydroxysuccinimide groups. 158<br />

Another conjugation method is the postpolymerization modification <strong>of</strong> CPs. For example,<br />

poly(acrylic acid) was grafted onto the unmodified PPy substrates by a photo-grafting method.<br />

The bonded carboxylic groups can be used for the subsequent immobilization <strong>of</strong> glucose<br />

oxidase. This method only altered the surface property <strong>of</strong> the substrate <strong>and</strong> did not change its<br />

bulk, which will not affect the conductivities <strong>of</strong> materials. 160<br />

Molecular imprinting technique is a novel method to precisely recognize molecules. A<br />

recent report provided a new approach for a biosensor. 161 The authors added caffeine during<br />

the electropolymerizing <strong>of</strong> PPy <strong>and</strong> subsequently the caffeine molecules were removed to<br />

‐ 34 ‐


- 35 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

form their imprints, which are unique three-dimensional sites only selectively recognizing<br />

caffeine. Another novel biosensor technique is an indirect way. Le Floch et al. coated the<br />

aptamers covalently immobilized gold electrodes with ferrocene-modified PT. In this study,<br />

the bioactive aptamer is not immobilized directly to the CP; however, the CP still can act as<br />

an electrical transducer. 162<br />

1.5.3 Other Applications<br />

In addition to the tissue engineering, biosensors <strong>and</strong> probes, the electroactive polymers also<br />

have biological applications in other fields, such as drug delivery carrier, actuators, <strong>and</strong><br />

antioxidants. The eleectroactive polymers exhibited a reversible redox property, which<br />

triggers adsorption or expulsion <strong>of</strong> ions leading to the volume change <strong>of</strong> materials. 163 Most <strong>of</strong><br />

their applications in drug delivery <strong>and</strong> actuator are based on this electricity-responsive<br />

property. When they were used as drug delivery carriers, the incorporated biologically active<br />

proteins <strong>and</strong> drugs such as NGF, 128 dexamethasone, 98, 164 <strong>and</strong> heparin 165 can be controlled<br />

release through electrical stimulation. For example, Hodgson et al. reported an approach to<br />

entrap <strong>and</strong> electrically controlled release bovine serum albumin (BSA) <strong>and</strong> NGF from PPy<br />

doped with polyelectrolytes (e.g., dextran sulfate). 128<br />

Figure. 1.5.16 (A) Postpolymerization <strong>of</strong> PPy to incorporate aldehyde groups. (B) Covalent<br />

immobilization <strong>of</strong> PVA hydrogels containing heparin on PPy substrates. 165


Chapter 1<br />

The introduction <strong>of</strong> polyelectrolytes led to the materials being hydrophilic enough to swell<br />

in water, which allows the easy diffusion <strong>of</strong> entrapped protein. When the PPy was reduced by<br />

negative potential, the anions were quickly expulsed in less than one minute. The in vitro test<br />

demonstrated that the released NGF still retained activity. Generally the polymeric carriers<br />

will affect the entrapped protein’s folding <strong>and</strong> activity due to the strong hydrophobic nature<br />

<strong>of</strong> the polymer. However, this approach employed polyelectrolytes to increase the<br />

hydrophilicity <strong>of</strong> the materials, which had solved this problem in some extent.<br />

In another investigation, authors incorporated streptavidin into the PPy films doped with<br />

both biotin <strong>and</strong> dodecylbenzenesulfonate. Then the biotinylated NGF was immobilized to this<br />

film. When the film was electrically stimulated, NGF was exclusively released. The activity<br />

<strong>of</strong> the released protein was also confirmed in vitro. This method can be exp<strong>and</strong>ed to most <strong>of</strong><br />

drugs that could be delivered through biotin-streptavidin strategy.<br />

Hydrogels have been studied for a long time as drug carrier. The release behaviors <strong>of</strong><br />

“smart” hydrogels can be controlled by many stimuli, including pH, temperature,<br />

biomacromolecues <strong>and</strong> electrical stimulation. For example, PVA hydrogels were covalently<br />

immobilized onto aldehyde groups modified PPy films (Figure. 1.5.16). The heparin was<br />

loaded into the hydrogels <strong>and</strong> could be released through the diffusion without electrically<br />

stimulation. 165 When PPy was electrically stimulated, the release <strong>of</strong> heparin from the<br />

hydrogel was accelerated. Many factors were considered by the authors that could affect the<br />

results, such as changes in temperature, electrophoresis, changes in pH as a result <strong>of</strong><br />

electrolysis, <strong>and</strong> the polyanionic nature <strong>of</strong> heparin.<br />

Besides the film, electroactive polymers can be fabricated into other shapes for drug<br />

delivery. For example, PEDOT nanotubes were synthesized on the electrospun PLGA fibers<br />

(~100 nm diameter) template (Figure. 1.5.17). 164 Then the PLGA loaded with dexamethasone<br />

was removed to produce PEDOT nanotubes encapsulating small molecules. The drug release<br />

<strong>of</strong> PEDOT nanotubes drug delivery system can be controlled by the electrical stimulation,<br />

probably as a consequence <strong>of</strong> expansion/reduction <strong>of</strong> polymer cavities induced by the<br />

expulsion <strong>of</strong> anions.<br />

‐ 36 ‐


- 37 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

Figure. 1.5.17 (A) Electrospun PLGA fibers are loaded with drug (dexamethasone). (B)<br />

Degradation <strong>of</strong> PLGA fibers release the drug. (C) PEDOT is electropolymerized around PLGA<br />

fibers loaded with drug. (D) After degradation <strong>of</strong> PLGA fibers, PEDOT nanotubes are loaded<br />

with drug. (E) PEDOT nanotubes do not release drug in a neutral electrical condition. (F) PEDOT<br />

nanotubes release the drug upon external electrical stimulation with a positive voltage, which<br />

produces contraction <strong>of</strong> the nanotubes <strong>and</strong> the subsequent expulsion <strong>of</strong> the drug. (G) SEM image.<br />

Electropolymerized PEDOT nanotubes on the electrode site <strong>of</strong> an acute neural microelectrode<br />

after removing the PLGA core fibers. (H) Higher magnification <strong>of</strong> a single PEDOT nanotube on a<br />

neural microelectrode array. 164<br />

The volume change <strong>of</strong> electroactive polymers responding to electrical stimulation has also<br />

been explored to develop actuators as “artificial muscle” devices. For example, Otero et al.<br />

placed two layers <strong>of</strong> PPy in a triple layer arrangement separated by a non-conductive material,<br />

as illustrated in Figure. 1.5.18. 166, 167 When current applied across the two conductive films,<br />

one <strong>of</strong> the films is oxidized <strong>and</strong> the other is reduced. The oxidized film exhibited an inflow <strong>of</strong><br />

dopant ions <strong>and</strong> an associated expansion, whereas the reduced film expels ions <strong>and</strong> shrinks. 166,<br />

168<br />

The combined effect is changed into a mechanical force that bends the films, which has<br />

been compared to the mechanisms in natural muscles. In order to increase the<br />

electromechanical actuation <strong>of</strong> the materials, carbon nanotubes (CNTs) were blended with


Chapter 1<br />

PANI fibers. 169 These devices have promising biomedical applications as<br />

electrical-responsive actuators, such as micropumps <strong>and</strong> valves for labs-on-a-chip, 170<br />

steerable catheters for minimally invasive surgery, 171 blood vessel connectors (Figure.<br />

1.5.19), 172 <strong>and</strong> microvalves for urinary incontinence.<br />

Figure. 1.5.18 (A) When current is applied, the left PPy film acts as anode <strong>and</strong> swells by the<br />

entry <strong>of</strong> the hydrated counter ions. Simultaneously, the right film acts as cathode <strong>and</strong> contracts by<br />

the expulsion <strong>of</strong> the counter ions. These volume changes <strong>and</strong> the constant length <strong>of</strong> the<br />

non-conducting film promote the movement <strong>of</strong> the triple layer device. (B) By changing the<br />

direction <strong>of</strong> current the movement takes place in the opposite direction. 167<br />

Scavenging <strong>of</strong> free radicals is a protective response in biological system, which may afford<br />

against various diseases, such as cardio-vascular diseases <strong>and</strong> cancer. The potential role <strong>of</strong><br />

PANi 129 as antioxidants in rubber mixes has already been considered, <strong>and</strong> the PANi was<br />

demonstrated to be efficient in slowing down the rate <strong>of</strong> oxidation. And recently the<br />

electroactive polymers were also used in biological applications. 173 These polymers such as<br />

PANi <strong>and</strong> PPy in their neutral form can eliminate 2 ~ 4 free radicals per monomer unit. The<br />

ability <strong>of</strong> these materials to scavenge free radicals is attributed to their oxidation activity <strong>and</strong><br />

has no relationship with electrical stimulation. These materials have potential biomedical<br />

applications in tissues experiencing high oxidative stress.<br />

‐ 38 ‐


- 39 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

Figure. 1.5.19 PPy blood vessel connector insertion surgery. (A, B) A PPy-Au bilayer device in<br />

the reduced state is curled into a roll that is inserted half-way into one <strong>of</strong> the ends <strong>of</strong> the severed<br />

blood vessel. (C, D) The other half <strong>of</strong> the device is inserted inside the other end <strong>of</strong> the vessel <strong>and</strong><br />

the bilayer exp<strong>and</strong>s a few minutes after PPy is oxidized. The PPy tube holds the two parts <strong>of</strong> the<br />

blood vessel together during healing, which replaces sutures. The connector is non-thrombogenic<br />

<strong>and</strong> very thin to avoid restricting the space inside the blood vessel. 172<br />

Although many kinds <strong>of</strong> electroactive polymers (most <strong>of</strong> them are CPs) have been explored<br />

for biomedical application, little attentions have been paid on the biologically active radical<br />

polymers. Therefore, in this thesis we prepared novel TEMPO-contained electroacitve<br />

polymers <strong>and</strong> investigated the feasibility <strong>of</strong> them in biomedical application. This exploration<br />

was creative <strong>and</strong> important for developing novel electroacitve biomaterials.<br />

1.6 Radicals Polymers<br />

As a stable radical, 2,2,6,6-Tetramethylpiperidine-1-oxyl (TEMPO) has applications<br />

throughout chemistry <strong>and</strong> biochemistry. For example, it was widely used as a radical trapping,<br />

as a spinning-labeled reagent in biological systems, as a reagent in organic synthesis, <strong>and</strong> as a<br />

mediator in controlled free radical polymerization. 174-177 TEMPO is prepared by the oxidation<br />

<strong>of</strong> 2,2,6,6-tetramethylpiperidine (TEMP). It was well known that the nitroxide free radicals<br />

could participate in one-electron redox reaction in acidic medium to yield relatively stable


Chapter 1<br />

diamagnetic products, hydroxypiperidines structure (NOH), <strong>and</strong> oxo-piperidinium cations<br />

(NO + ). Figure 1.6.1 shows such one-electron redox reactions <strong>of</strong> free nitric oxide derivatives.<br />

Figure 1.6.1 The one-electron redox reactions <strong>of</strong> nitroxide free radicals. 178<br />

Based on the redox properties <strong>of</strong> the nitroxyl radicals, polymers bearing stable radicals in<br />

the side chains, also known as radical polymers, have been extensively studied as redox<br />

resins which catalyze the oxidative <strong>and</strong>/or reductive reactions <strong>of</strong> organic compounds. For<br />

example, poly(acrylate)-combined TEMPOs were synthesized <strong>and</strong> studied as a catalytic<br />

reagent for the oxidation <strong>of</strong> alcohols into aldehydes <strong>and</strong> ketones. 179 The organic radical-based<br />

or metal-free redox reagents have been recently reexamined from the perspective <strong>of</strong> green or<br />

environmentally compatible chemical reaction processes.<br />

In addition to this chemistry application, the robust electron exchange between unpaired<br />

electrons in the polymer side chains suggests these radical polymers to be used for organic<br />

electronic or magnetic materials. However, it is not until recently that the applications have<br />

been intensively investigated for use in organic secondary battery or memory devices by our<br />

group. The large population <strong>of</strong> the radical redox sites along the polymer chain allows the<br />

effective redox gradient-deriven electron transport, which leads to organic batteries with large<br />

energy density. On the other h<strong>and</strong>, polymers with nitroxyl radicals also attracts significant<br />

interest to be as biomedical materials since the stable nitroxyl radials have been discovered to<br />

have similar functions as nitric oxide (NO•) which was presence in many bioprocess such as<br />

inflammation, blood clotting, blood pressure, neurotransmission, cardiovascular disorders <strong>and</strong><br />

antimicrobial property. 180-182 Thus, emphasis <strong>of</strong> this part <strong>of</strong> introduction would focus on the<br />

radical polymers for organic electronic devices <strong>and</strong> biomedical applications.<br />

‐ 40 ‐


1.6.1 Radical Polymers for Organic Electronic Devices<br />

1.6.1.1 Organic Radical Batteries<br />

- 41 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

Rechargeable secondary batteries are widely used in portable equipments. Li-ion battery<br />

represents the currently most popular usage. However, the application <strong>of</strong> metal-based<br />

electrodes has come up against some inherent disadvantages, such as limited raw-material<br />

resource <strong>and</strong> tedious waste process. Therefore, organic-based electrode-active materials have<br />

received more <strong>and</strong> more attentions. Over the past few years, we have focused on the<br />

development <strong>of</strong> polymer bearing densely populated unpaired electrons, such as nitroxides,<br />

phenoxyl <strong>and</strong> galvinoxyl, in the pendant per repeating unit <strong>and</strong> utility <strong>of</strong> them as<br />

electro-active or charge-storage material in a rechargeable device. 178, 183-190 The organic<br />

178, 183, 188<br />

radical battery composed <strong>of</strong> the radical polymer electrodes has several advantages:<br />

(1) a high charging <strong>and</strong> -discharging capacity (>100 mAh/g), ascribed to the stoichiometric<br />

redox <strong>of</strong> the radical moieties, (2) a high-charging <strong>and</strong> -discharging rate performance resulting<br />

from the rapid electron-transfer process <strong>of</strong> the radical species <strong>and</strong> from the amorphous state <strong>of</strong><br />

the radical polymers, <strong>and</strong> (3) a long cycle life, <strong>of</strong>ten exceeding 1000 cycles, derived from the<br />

chemical stability <strong>of</strong> the radicals <strong>and</strong> from the amorphous electrode structure.<br />

A variety <strong>of</strong> polymer backbones have been employed by our group to bear the pendant<br />

radicals, such as poly(meth)arylates, polystyrene derivatives, polymer(vinyl ether)s,<br />

polyethers <strong>and</strong> poly(norbornene)s, <strong>and</strong> were depicted in Figure 1.6.2 <strong>and</strong> Figure 1.6.3. Radical<br />

polymers, most <strong>of</strong> time, were synthesized by conventional radical polymerizations, <strong>and</strong> an<br />

additional oxidation processes must be needed in this case. The unpaired electron density <strong>of</strong><br />

the polymers was then measured by SQUID, revealing the presence <strong>of</strong> the radicals in each<br />

repeating unit. The versatile <strong>of</strong> the polymers backbones also have another advantage, i.e. the<br />

polymers allow to be conveniently placed on the surface <strong>of</strong> a current collector by a<br />

solution-based wet process such spin-coating method. When the polymer is placed on the<br />

surface <strong>of</strong> the current collector <strong>and</strong> equilibrated in electrolyte solution, charge propagation<br />

within the polymer layer is sufficiently accomplished, leading to high-density charge storage<br />

because the redox sites are so populated that electron self-exchange reactions are completed<br />

within a finite distance <strong>of</strong> the polymer layer. 189


Chapter 1<br />

Figure 1.6.2 Nitroxide radical polymers based on various backbones. 188<br />

Figure 1.6.3 n-Type radical polymers. 188<br />

Typical radical polymers such poly(2,2,6,6-tetramethypiperidinyloxyl-4-yl-methyarylate)<br />

(PTAM) is a “p-type” redox active materials in organic battery. In order to obtain totally<br />

organic radical battery, we have recently extensively explored the n-type redox-active radical<br />

polymers based on the molecular design (Figure 1.6.4). A representative combination is the<br />

PTAM as the cathode <strong>and</strong> the poly(galvinoxylstyrene) as the anode. 190 The radical density <strong>of</strong><br />

each polymer was >0.9 unpaired electrons per monomer unit. Solutions <strong>of</strong> the radical<br />

polymers were coated as thin films on an ITO-PET substrate as the current collectors. The<br />

coated radical polymers were suitably modified via a cross-linking reaction to impede their<br />

dissolution into the electrolyte solution which causes self-discharging <strong>of</strong> the battery. A<br />

microporous separator film containing the electrolyte solution, such as ethylene carbonate<br />

‐ 42 ‐


- 43 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

containing tetrabutylammonium chloride, was s<strong>and</strong>wiched between two radical polymer films<br />

coated on the current collectors, to fabricate the all-organic “radical battery”.<br />

Figure 1.6.4 Totally organic radical battery. 188<br />

Though the organic-based batteries composed <strong>of</strong> radical polymers pose considerable<br />

advantages over the Li-ion batteries, the flammable or ignition risk <strong>of</strong> the organic electrolytes<br />

in the devices emerged as the fundamental safety issue in the practical use. To overcome this<br />

problem, radical polymer poly(2,2,6,6-tetramethylpiperidinyloxyl-4-yl vinylether) (PTVE)<br />

with a hydrophilic polyvinylether backbone was synthesized <strong>and</strong> demonstrated<br />

charging-discharging operation in aqueous electrolyte. 186, 187 However, the PTVE based<br />

devices were limited by their long cycle operation due to the low polymer weight <strong>and</strong><br />

insufficient hydrophilicity <strong>of</strong> the polymers. Thus, an alternative type <strong>of</strong> radical polymer,<br />

poly(2,2,6,6-tetramethylpiperidinyloxyl-4-yl acrylamide), was designed <strong>and</strong> synthesized as an<br />

electrode-active polymer for organic rechargeable device containing aqueous electrolyte. 191<br />

The device demonstrated a 1.2 V output voltage, exceeded 2000 charging-discharging cycles,<br />

<strong>and</strong> a high charging rate performance within 1 min.<br />

1.6.1.2 Organic Radical Memories<br />

Besides the organic-based batteries using radical polymers, the development <strong>of</strong> a similar<br />

charge-storage configuration has also been anticipated for a dry system in the absence <strong>of</strong> the<br />

electrolyte by s<strong>and</strong>wiching a dielectric material within the radical polymers. 192 The<br />

metal–insulator–metal diode-type structure <strong>of</strong> the radical memory (see Figure 1.6.5) is<br />

composed <strong>of</strong> the thin layers <strong>of</strong> a p-type radical polymer such as PTMA, poly(vinylidene<br />

difluoride) (PVDF) as the dielectric material, <strong>and</strong> an n-type radical polymer such as PGSt.<br />

These layers have been conveniently spin-coated onto an indium tin oxide (ITO)/glass<br />

electrode without intermixing. When an increasing voltage <strong>of</strong> 0 to –5V is applied, the state <strong>of</strong>


Chapter 1<br />

the device is switched to a low resistance near –4.5 V as the threshold voltage. The low<br />

resistance state is maintained during the reverse sweep <strong>of</strong> the voltage from –5 to +1.4 V. The<br />

low-resistance or ON state was observed for repeated sweeps, regardless <strong>of</strong> the sweep<br />

direction. When a bias <strong>of</strong> –1.5 V was applied, the device sharply switched to the<br />

high-resistance or OFF state again. In the hysteretic curve, the ON–OFF ratio amounted to<br />

four orders <strong>of</strong> magnitude. The retention cycles <strong>of</strong> the ON <strong>and</strong> OFF states under open-circuit<br />

conditions persisted for more than 10 4 times. Furthermore, each state survived for a month<br />

after the corresponding once-time-only pulse as well as consecutive pulses.<br />

The charge injection at the radical polymer/electrode interface <strong>and</strong> transport in the bulk <strong>of</strong><br />

the radical polymer layer are dominated by the Schottky barrier <strong>and</strong> the Poole–Frenkel<br />

mechanism. In the ON state, charges are trapped at the radical polymer/PVDF interfaces, <strong>and</strong><br />

induce accumulation <strong>of</strong> the opposite charge at the radical polymer/electrode interfaces <strong>and</strong><br />

reduction the Schottky barriers, allowing charges to transfer across the radical<br />

polymer/electrode interfaces even at low voltages. 192 The SOMO levels <strong>of</strong> the p-type PTMA<br />

<strong>and</strong> the n-type PGSt are near 5.4 <strong>and</strong> 4.7 eV, respectively.<br />

Figure 1.6.5 p-type <strong>and</strong> n-type radical polymer-based memory architecture, <strong>and</strong> charge injection,<br />

transporting, <strong>and</strong> trapping configuration.<br />

The p-type layer accepts holes, <strong>and</strong> electrons are injected into the n-type layer. The<br />

injected holes <strong>and</strong> electrons are transported by the hopping mechanism, <strong>and</strong> are stored at the<br />

radical polymer/PVDF interfaces. Under open-circuit conditions, the trapped holes <strong>and</strong><br />

electrons were nonvolatile due to charge trapping at the radical polymer/PVDF interfaces <strong>and</strong><br />

to blocking <strong>of</strong> the charge transfer in the polymer layers having sufficient resistances. The<br />

asymmetric configuration <strong>of</strong> the electrodes with the different work functions allows<br />

‐ 44 ‐


initialization <strong>of</strong> the device by applying an inverse voltage.<br />

1.6.2 Radical Polymers for Biomedical Applications<br />

1.6.2.1 Nitric Oxide Release Systems<br />

- 45 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

Nitric oxide (NO•) is a small but highly reactive free radical with exp<strong>and</strong>ing known<br />

biological activities. Its essential roles involve the regulatory agent for normal physiological<br />

activities <strong>and</strong> cytotoxic species in diseases <strong>and</strong> their treatments. There have been numerous<br />

examples <strong>of</strong> application <strong>of</strong> the nitric oxide on disease treatments, such as antiviral compounds,<br />

cancer treatment, anti-inflammatory drugs <strong>and</strong> other nitric oxide-related diseases treatment.<br />

However, the excess introduction <strong>of</strong> nitric oxide into body may induce significant adverse<br />

side effects like microvascular leakage, tissue damage in cystic fibrosis, septic shock, B-cell<br />

destruction, <strong>and</strong> possible mutagenic risk. 193 Therefore, it is very important to develop smart<br />

delivery vehicles for control release <strong>of</strong> the nitric oxide.<br />

<strong>Biodegradable</strong> polymers are known to be biocompatible <strong>and</strong> degradable in physiological<br />

conditions which have been widely utilized as drug delivery vehicles in the forms <strong>of</strong> matrices<br />

or nanoparticles. Accordingly, pioneered works by C. C. Chu <strong>and</strong> co-workers have reported to<br />

incorporate <strong>of</strong> stable nitric oxide radicals to either the chain end or backbone <strong>of</strong> the<br />

biodegradable polymers. 193-196 In such contributions, nitric oxide derivative, tampamine<br />

nitroxyl radical (4-amino-2,2,6,6-tetramethylpiperidine-1-oxy, TAM) have been chemically<br />

coupled with various biodegradable polymers such as polyglycolide (PGA), poly(acrylic<br />

acid/lactide/ε-caprolactone) (PBLCA) <strong>and</strong> poly(ester amide)s (PEAs). The biological<br />

activities <strong>of</strong> these TAM-incorporated polymers <strong>and</strong> the release kinetic <strong>of</strong> the TAM from the<br />

polymer matrices have also been evaluated. For example, the level <strong>of</strong> the retardation <strong>of</strong><br />

smooth muscle cell (SMC) <strong>of</strong> the TAM-PGA was conducted in vitro cell culture (Figure<br />

1.6.6). The TAM-PGA was found to show pr<strong>of</strong>ound retardation <strong>of</strong> the proliferation <strong>of</strong> SMC as<br />

similar with the TAM free nitroxyl radicals at concentration <strong>of</strong> 1 μg/mL. Thus, it appeared<br />

that the long PGA segments had no evidently interference on the biological functions <strong>of</strong><br />

nitroxyl radicals incorporated. However, the incorporation <strong>of</strong> TAM into the polymer chain<br />

end have limitation in the TAM content in the polymer, improved strategies have also been<br />

proposed by the authors via conjugating the TAM to the backbone <strong>of</strong> the biodegradable<br />

polymers such as PBLCA <strong>and</strong> PEAs. 193, 194 Up to 8.32% (wt %) <strong>of</strong> the TAM content in


Chapter 1<br />

PBCLA could be achieved. Though the TAM content was improved, the release <strong>of</strong> TAM from<br />

the materials seen to be less effective even in the presence <strong>of</strong> hydrolysis lipase. Only less than<br />

10 % <strong>of</strong> the TAM was released after 6 days. In order to solve this issue, a more versatile<br />

method has been investigated recently. The TAM nitroxyl radicals were loaded in the<br />

biodegradable PEAs nan<strong>of</strong>iber membranes instead <strong>of</strong> chemical conjugation. 195 The payload<br />

<strong>of</strong> the TAM in the nan<strong>of</strong>iber could be tunable by pre-loading different amount <strong>of</strong> TAM <strong>and</strong><br />

the release kinetic study revealed that the encapsulated TAM would be completely released<br />

from the membranes within 5 days.<br />

Figure 1.6.6 The effect <strong>of</strong> TAM-PGA on the proliferation <strong>of</strong> human smooth muscle. 196<br />

1.6.2.2 Magnetic Resonance Imaging (MRI)<br />

MRI allows noninvasive imaging <strong>of</strong> deep internal structure <strong>of</strong> the body <strong>and</strong>, is<br />

widespread use in clinical diagnostic <strong>of</strong> physiologic <strong>and</strong> pathologic changes <strong>of</strong> disease. For<br />

MRI, the paramagnetic species are needed to serve as the contrast agents to enhance the<br />

image contrast. The most commercially used MR contrast agents are the gadolinium chelates,<br />

such as gadolinium diethylenetriaminepentaacetic aicd (Gd-DTPA). However, these agents<br />

are non-selective distribution in tissues <strong>and</strong> have short resistant time in tumors. Nitroxyl<br />

radicals are only well-known for their electronic paramagnetic resonance (EPR) properties<br />

until 1984, the T1 contrast property was found. Afterward, MRI applications <strong>of</strong> the nitroxyl<br />

radicals have been extensively investigated owing to their chemical flexibility, feasible<br />

preparation <strong>and</strong> low toxicity as comparison with conventional gadolinium chelates.<br />

Unfortunately, there are two fetal flaws would emerge when applied the nitroxyl radical MRI<br />

reagents in vivo. One is the relatively low MR relaxivities, while the other is the<br />

‐ 46 ‐


- 47 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

paramagnetic nitroxyl radical may be reduced to diamagnetic hydroxylamine with the loss <strong>of</strong><br />

T1 MR relaxation in the internal reduction environment.<br />

Figure 1.6.7 illustration <strong>of</strong> HPS-TEMPO.<br />

To address the abovementioned issues, several strategies have been reported, including<br />

conjugation <strong>of</strong> nitroxyl radical compounds with dendritic/hyperbranched polymers or<br />

copolymerization with hydrophilic monomers. As an example, Winalski et al. 197 reported to<br />

find that the fourth-generation polypropylenimide-(DAB) <strong>and</strong> third-generation<br />

polyamidoamide-(PAMAM) dendrimer-linked nitroxides had greater relaxivity than<br />

Gd-DTPA <strong>and</strong> the selective enhance contrast were observed in healthy articular cartilage. The<br />

relaxivities <strong>of</strong> the compounds were found to increase as the increase <strong>of</strong> nitroxides that<br />

incorporated in the dendrimers. This result was also observed by Gh<strong>and</strong>ehari et al. <strong>and</strong> Koga<br />

et al.. Koga <strong>and</strong> co-workers have synthesized a series <strong>of</strong> high water-soluble hyperbranched<br />

poly(styrene) (HPS) carrying stable TEMPO radicals (Figure 1.6.7). 198 The relaxivity 14<br />

mM -1 s -1 , which is comparable with commercial Gd-DTPA agent, could be obtained when<br />

high density <strong>of</strong> TEMPO radicals has been conjugated with the polymers. Studies by<br />

Gh<strong>and</strong>ehari et al. 199 have been reported that N-(2-hydroxylpropyl)methacryamide (HPMA)<br />

copolymer-linked nitroxides were synthesized for potential MRI application. The HPMA<br />

monomer was firstly copolymerized with reactive MAGGONp monomer <strong>and</strong> then chemical<br />

link with the nitroxides was performed via reactive MAGGONp residues. The relaxicities <strong>of</strong><br />

the copolymers bearing nitroxides were found to be linear dependent on the nitroxides<br />

content in the copolymers which is consistent with the previous report by Winalski et al..<br />

Another strategy has also proposed by Gussoni <strong>and</strong> co-workers. 200 They have synthesized<br />

pH-sensitive poly(amidoamine)s containing TEMPO radicals in the backbone. Though the<br />

relaxicity was measured to be not too high, the polymers were found to be tendency to<br />

accumulate in the tumor for a sufficient time.


Chapter 1<br />

1.6.2.3 Electronic Paramagnetic Resonance (EPR) imaging<br />

Electron paramagnetic resonance imaging (EPRI) is one <strong>of</strong> the recent functional imaging<br />

modalities that can provide valuable in vivo physiological information, such as tissue redox<br />

status, pO2, pH, <strong>and</strong> microviscosity, based on variation <strong>of</strong> EPR spectral characteristics, i.e.,<br />

intensity, linewidth, hyperfine splitting, <strong>and</strong> spectral shape <strong>of</strong> free radical probes. 201 EPR<br />

imaging (EPRI) can obtain 1D–3D spatial distribution <strong>of</strong> such spectral components using<br />

several combinations <strong>of</strong> magnetic field gradients. With the addition <strong>of</strong> appropriate<br />

paramagnetic probes, the sensitivity <strong>of</strong> EPR on a molar basis is about 700 times greater than<br />

that <strong>of</strong> NMR. For EPRI applications, the spins probes should be (1) chemically stable <strong>and</strong><br />

water soluble in biological media; (2) have simple EPR spectrum at ambient temperature; (3)<br />

have pharmacological half-time <strong>of</strong> at least 10 min to permit the imaging; (4) be<br />

non-toxicity. 202 The most used species are nitroxides, such as TEMPO radicals, due to their<br />

excellent EPR effect. However, the small molecular species pose some limitations, such as<br />

non-selective accumulation in normal tissues, quickly renal clearance, <strong>and</strong> rapid reduce under<br />

the reduction environment in vivo. 203 To solve these issues, Nagasaki <strong>and</strong> co-workers have<br />

designed core–shell-type nanoparticles carrying stable radicals in the core (Figure 1.6.8). 204,<br />

205<br />

The nanoparticles showed intense EPR signals <strong>and</strong> could resist to reduction environment<br />

even at the presence <strong>of</strong> 3.5 mM ascorbic acid which suggested the promising use <strong>of</strong> these<br />

nanoparticles to be use as the EPR probes in vivo. Further studies <strong>of</strong> such nanoparticles<br />

applied in vivo have revealed that the RNPs had more sufficient long-term blood circulation<br />

compared to the TEMPOL free radicals <strong>and</strong> were <strong>of</strong> extremely low toxicity due to the<br />

confinement <strong>of</strong> the TEMPO moieties in the nanoparticle core. Especially, the RNPs were pH<br />

sensitive, the L-b<strong>and</strong> EPR signal could be observed only at the pH value below 6. Therefore,<br />

the pH-sensitive RNPs demonstrated the promising use as the EPR probes for low pH<br />

circumstances in vivo.<br />

Figure 1.6.8 Illustration <strong>of</strong> the radical-containing nanoparticle. 204<br />

‐ 48 ‐


1.6.2.4 Others<br />

- 49 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

Stimuli-responsive polymers have step critical improvement to be developed as the<br />

biomedical materials. There are various factors that have been discovered to induce<br />

responsive <strong>of</strong> the polymers such pH, temperature, light, electric/magnetic field, biological<br />

events <strong>and</strong> so on. 206 Due to one electron redox properties <strong>of</strong> the nitroxyl radicals, polymers<br />

bearing these compounds are expected to have the redox sensitivity that may find potential<br />

use as the biomedical materials. Yoshida <strong>and</strong> Tanaka have proposed the oxidation-induced 207<br />

<strong>and</strong> reduction-induced 208 micellization <strong>of</strong> a diblock copolymer containing stable nitroxyl<br />

radicals. Light scatting technology <strong>and</strong> UV-Vis absorbance were used to characterization <strong>of</strong><br />

the reversible micellization induced by the redox systems which have been illustrated in the<br />

Figure 1.6.1. However, the stimuli-responsive systems the authors have studied were<br />

performed in organic solvent which is far beyond the applications as biomaterials. Therefore,<br />

further efforts should be needed to investigate the reversible self-assembly systems in the<br />

aqueous or physiological medium <strong>and</strong> thus applications in biomedical field could be found.<br />

Reference<br />

1. LS. Nair, L. C, Prog. Polym. Sci. 2007, 32, 762.<br />

2. D. J. Brunelle, M. R. Korn, Proceedings <strong>of</strong> Symposium <strong>of</strong> the American Chemical Society,<br />

March 2003, Washington, DC 2005, 281.<br />

3. J. H. Clements, Ind. Eng. Chem. Res 2003, 42, 663.<br />

4. S. Inoue, H. K., T. Tsuruta, J. Polym. Sci., Part B: Polym. Phys. 1969, 7, 287.<br />

5. S. Inoue, H. K., T. Tsuruta, Makromol. Chem 1969, 130, 210.<br />

6. M. Kobayashi, S. I., T. Tsuruta, Macromolecules 1971, 4, 658.<br />

7. M. Kobayashi, Y. L. T., T. Tsuruta, S. Inoue, Makromol.Chem 1973, 169, 69.<br />

8. M. Kobayashi, S. I., T. Tsuruta, J. Polym. Sci. Polym. Chem. Ed. 1973, 11, 2383.<br />

9. S. Inoue, M. K., H. Koinuma, T. Tsuruta, Makromol.Chem 1972, 155, 61.<br />

10. K. I. Soga, E.; Hattori, I., Polym. J. 1981, 13, 407.<br />

11. A. Rokicki, U.S. Patent 4,943,677, 1990. Products <strong>and</strong> Chemicals, Inc.; Arco Chemicals<br />

Co. 1990.<br />

12. D. J. S. Darensbourg, N. W.; Katsurao, T., J. Mol. Catal. A 1995, 104, L1.<br />

13. D. J. H. Darensbourg, M. W.; Reibenspies, J. H., Polyhedron 1996, 15, 2341.


Chapter 1<br />

14. T. I. Aida, M.; Inoue, S., Macromolecules 1986, 19, 8.<br />

15. W. J. D. Kruper, D. V., J. Org. Chem 1995, 60, 725.<br />

16. S. C. Mang, A. I.; M. E. Colclough, N. Chauhan, A. B. Holmes, Macromolecules 2000,<br />

33, 303.<br />

17. D. J. H. Darensbourg, M. W.; Reibenspies, J. H., Macromolecules 1995, 28, 7577.<br />

18. C. W. Koning, J. R.; R. Parton, B. Plum, Steeman, P.; Darensbourg, D. J.; Holtcamp, M.<br />

W.; Reibenspies, J. H., Polymer 2001, 42, 3995.<br />

19. D. J. W. Darensbourg, J. R.; J. C. Yarbrough, J. Am. Chem. Soc. 2000, 122, 12487.<br />

20. S. Inoue, J. Polym. Sci., Part A: Polym. Chem. 2000, 38, 2861.<br />

21. N. Takeda, S. I., Makromol. Chem 1978, 179, 1377.<br />

22. T. Aida, S. I., Macromolecules 1981, 14, 1162.<br />

23. T. Aida, S. I., J. Am. Chem. Soc. 1983, 105, 1304.<br />

24. X. B. Lu, X. J. F., R. He,, Appl. Catal. A 2002, 234, 25.<br />

25. X. B. Lu, R. H., C. X. Bai,, J. Mol. Catal. A 2002, 186, 1.<br />

26. D. J. Darensbourg, E. L. M., M.W. Holtcamp, K. K.Klausmeyer, J. H. Reibenspies,,<br />

Inorg. Chem 1996, 35, 2682.<br />

27. W. J. Kruper, D. V. D., J. Org. Chem 1995, 60, 725.<br />

28. E. N.Jacobsen, Acc. Chem. Res 2000, 33, 421.<br />

29. E. N. Jacobsen, M. T., J. F. Larrow,, 2000, PCT Int. Appl.WO 00/09463.<br />

30. R. L. Paddock, S. T. N., J. Am. Chem. Soc. 2001, 123, 11498.<br />

31. X. H. Chen, Z. Q. S., Y. F. Zhang, Macromolecules 1991, 24, 5305.<br />

32. Z. Q. Shen, X. H. C., Y. F. Zhang, Macromol. Chem. Phys 1994, 195, 2003.<br />

33. J. T. Guo, X. Y. W., Y. S. Xu, J. W. Sun, J. Appl. Polym. Sci. 2003, 87, 2356.<br />

34. C. S. Tan, T. J. H., Macromolecules 1997, 30, 3147.<br />

35. T. J. Hsu, C. S. T., Polymer 2001, 42, 5143.<br />

36. Z. Quan, X. H. W., X. J. Zhao, F. S.Wang,, Polymer 2003, 44, 5605.<br />

37. B. Y. Liu, X. J. Z., X. H.Wang, F. S.Wang,, J. Polym. Sci., Part A: Polym. Chem. 2001,<br />

39, 2751.<br />

38. R. Auras, B. Harte, S. Selke, Macromol. Biosci. 2004, 4, 835.<br />

39. J. Zhang, P. Krishnamachari, J. Z. Lou, A. Shahbazi, In Proceedings <strong>of</strong> the 2007 National<br />

Conference on Environmental Science <strong>and</strong> Technology, Uzochukwu, G. A., Ed. Springer:<br />

New York, 2009; pp 3-8.<br />

‐ 50 ‐


- 51 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

40. M. Ajioka, K. Enomoto, K. Suzuki, A. Yamaguchi, Bull. Chem. Soc. Jpn. 1995, 68, 2125.<br />

41. S. Jacobsen, P. H. Degee, H. G. Fritz, P. H. Dubois, R. Jerome, Polym. Eng. Sci. 1999, 39,<br />

1311.<br />

42. H. R. Kricheldorf, M. Berl, N. Scharnagl, Macromolecules 1988, 21, 286.<br />

43. G. Schwach, J. Coudane, R. Engel, M. Vert, J. Polym. Sci., Part A: Polym. Chem. 1997,<br />

35, 3431.<br />

44. M. S. Reeve, S. P. McCarthy, M. J. Downey, R. A. Gross, Macromolecules 1994, 27, 825.<br />

45. G. Schwach, J. Coudane, R. Engel, M. Vert, Biomaterials 2002, 23, 993.<br />

46. L. Tang, L. Deng, J. Am. Chem. Soc. 2002, 124, 2870.<br />

47. O. T. Du Boullay, E. Marchal, B. Martin-Vaca, F. P. Cossio, D. Bourissou, J. Am. Chem.<br />

Soc. 2006, 128, 16442.<br />

48. C. Bonduelle, B. Martin-Vaca, F. P. Cossio, D. Bourissou, Chem.-Eur. J. 2008, 14, 5304.<br />

49. C. Bonduelle, B. Martin-Vaca, D. Bourissou, Biomacromolecules 2009, 10, 3069.<br />

50. X. Pang, H. Z. Du, X. S. Chen, X. H. Wang, X. B. Jing, Chem. Eur. J. 2008, 14, 3126.<br />

51. H. Z. Du, X. Pang, H. Y. Yu, X. L. Zhuang, X. S. Chen, D. M. Cui, X. H. Wang, X. B.<br />

Jing, Macromolecules 2007, 40, 1904.<br />

52. Z. H. Tang, V. C. Gibson, Eur. Polym. J. 2007, 43, 150.<br />

53. X. Pang, H. Z. Du, X. S. Chen, X. L. Zhuang, D. M. Cui, X. B. Jing, J. Polym. Sci., Part<br />

A: Polym. Chem. 2005, 43, 6605.<br />

54. Z. H. Tang, Y. K. Yang, X. Pang, J. L. Hu, X. S. Chen, N. H. Hu, X. B. Jing, J. Appl.<br />

Polym. Sci. 2005, 98, 102<br />

55. A. Kowalski, A. Duda, S. Penczek, Macromolecules 1998, 31, 2114.<br />

56. B. M. Chamberlain, M. Cheng, D. R. Moore, T. M. Ovitt, E. B. Lobkovsky, G. W. Coates,<br />

J. Am. Chem. Soc. 2001, 123, 3229.<br />

57. L. H. Piao, Z. L. Dai, M. X. Deng, X. S. Chen, X. B. Jing, Polymer 2003, 44, 2025.<br />

58. Z. Y. Zhong, P. J. Dijkstra, C. Birg, M. Westerhausen, J. Feijen, Macromolecules 2001,<br />

34, 3863-3868.<br />

59. Z. H. Tang, X. S. Chen, Q. Z. Hang, X. C. Bian, L. X. Yang, L. H. Piao, X. B. Jing, J.<br />

Polym. Sci., Part A: Polym. Chem. 2003, 41, 1934.<br />

60. H. R. Kricheldorf, A. Serra, Polym. Bull. 1985, 14, 497.<br />

61. Z. Y. Zhong, P. J. Dijkstra, J. Feijen, Angew. Chem.Int. Edit. 2002, 41, 4510.<br />

62. Z. H. Tang, X. S. Chen, Y. K. Yang, X. Pang, J. R. Sun, X. F. Zhang, X. B. Jing, J. Polym.


Chapter 1<br />

Sci., Part A: Polym. Chem. 2004, 42, 5974.<br />

63. Z. H. Tang, X. S. Chen, X. Pang, Y. K. Yang, X. F. Zhang, X. B. Jing,<br />

Biomacromolecules 2004, 5, 965.<br />

64. N. Nomura, A. Akita, R. Ishii, M. Mizuno, J. Am. Chem. Soc. 2010, 132, 1750.<br />

65. M. J. Stanford, A. P. Dove, Chem. Soc. Rev. 2010, 39, 486.<br />

66. H. Z. Du, A. H. Velders, P. J. Dijkstra, J. R. Sun, Z. Y. Zhong, X. S. Chen, J. Feijen,<br />

Chem.Eur. J. 2009, 15, 9836.<br />

67. D. Pappalardo, L. Annunziata, C. Pellecchia, Macromolecules 2009, 42, 6056.<br />

68. K. J. L. Burg, S. Porter, J. F. Kellam, Biomaterials 2000, 21, 2347.<br />

69. J. P. Penning, H. Dijkstra, A. J. Pennings, Polymer 1993, 34, 942.<br />

70. R. Bodmeier, K. H. Oh, H. Chen, Int. J. Pharm. 1989, 51, 1.<br />

71. H. Ohara, U. Doi, M. Otsuka, H. Okuyama, S. Okada, Seibutsu-Kogaku Kaishi-J. Soc.<br />

Ferment. Bioeng. 2001, 79, 142.<br />

72. M. Labet, W. Thielemans, Chem. Soc. Rev. 2009, 38, 3484.<br />

73. I. Barakat, P. Dubois, R. Jerome, P. Teyssie, Macromolecules 1991, 24, 6542.<br />

74. D. Bratton, M. Brown, S. M. Howdle, Macromolecules 2005, 38, 1190.<br />

75. V. Bergeot, T. Tassaing, M. Besnard, F. Cansell, A. F. Mingotaud, J. Supercrit. Fluids<br />

2004, 28, 249.<br />

76. S. W. Lee, S. J. Heo, J. Y. Jang, J. Y. Jeong, S. H. Kim, S. A. Park, E. S. Jeon, J. W. Shin,<br />

Tissue Eng. Regen. Med. 2010, 7, 9.<br />

77. S. Martinez-Diaz, N. Garcia-Giralt, M. Lebourg, J. A. Gomez-Tejedor, G.Vila, E. Caceres,<br />

P. Benito, M. M. Pradas, X. Nogues, J. L. G. Ribelles, J. C. Monllau, Am. J. Sports Med.<br />

2010, 38, 509.<br />

78. P. F. McDonald, J. G. Lyons, L. M. Geever, C. L. Higginbotham, J. Mater. Sci. 2010, 45,<br />

1284-1292.<br />

79. M. A. Del Nobile, A. Conte, G. G. Buonocore, A. L. Incoronato, A. Massaro, O. Panza, J.<br />

Food Eng. 2009, 93, 1.<br />

80. P. Ferreira, J. F. J. Coelho, M. H. Gil, Int. J. Pharm. 2008, 352, 172.<br />

81. Y. Aoyagi, C. K. Leong, D. D. L. Chung, J. Electron. Mater. 2006, 35, 416.<br />

82. C. Braud, R. Devarieux, A. Atlan, C. Ducos, M. Vert, J. Chromatogr. B 1998, 706, 73.<br />

83. H. Dong, S. G. Cao, Z. Q. Li, S. P. Han, D. L. You, J. C. Shen, J. Polym. Sci., Part A:<br />

Polym. Chem. 1999, 37, 1265.<br />

‐ 52 ‐


- 53 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

84. H. Dong, H. D. Wang, S. G. Cao, J. C. Shen, Biotechnol. Lett. 1998, 20, 905.<br />

85. Y. Shibasaki, H. Sanada, M. Yokoi, F. S<strong>and</strong>a, T. Endo, Macromolecules 2000, 33, 4316.<br />

86. J. C. Middleton, A. J. Tipton, Medical Plastics <strong>and</strong> Biomaterials Mar 1998, 30.<br />

87. D. K. Gilding, A. M. Reed, Polymer 1979, 20, 1459.<br />

88. S. L. Ishaug-Riley, L. E. Okun, G. Prado, M. A. Applegate, A. Ratcliffe, Biomaterials<br />

1999, 20, 2245.<br />

89. B. J. O'Keefe, M. A. Hillmyer, W. B. Tolman, J. Chem. Soc.-Dalton Trans. 2001, 15,<br />

2215.<br />

90. J. Zhou, G. Romero, E. Rojas, S. Moya, L. Ma, C. Y. Gao, Macromol. Chem. Phys. 2010,<br />

211, 404.<br />

91. W. D. Dai, N. Kawazoe, X. T. Lin, J. Dong, G. P. Chen, Biomaterials 2010 31, 2141.<br />

92. U. Bhardwaj, R. Sura, F. Papadimitrakopoulos, D. J. Burgess, Int. J. Pharm. 2010, 384,<br />

78.<br />

93. R. A. Miller, J. M. Brady, D. E. Cutright, J. Biomed. Mater. Res. 1977, 11, (5), 711-719.<br />

94. C. Simon-Colin, G. Raguenes, J. Cozien, J. G. Guezennec, J. Appl. Microbiol. 2008, 104,<br />

1425-1432.<br />

95. S. M. F. Pereira, R. S. Rodriguez, J. G. C. Gomez, Materia 2008, 13, 1.<br />

96. T. J. Rivers, T. W. Hudson, C. E. Schmidt, Adv. Funct. Mater. 2002, 12, 33.<br />

97. A. J. Heeger, Synth. Met. 2001, 125, 23-42.<br />

98. R. Wadhwa, C. F. Lagenaur, X. T. Cui, J. Controlled Release 2006, 110, 531.<br />

99. E. G. Fine, R. F. Valentini, R. Bellamkonda, P. Aebischer, Biomaterials 1991, 12, 775.<br />

100. H. Shirakawa, Synth. Met. 2001, 125, 3.<br />

101. K. I. Lee, H. Jopson, Polym. Bull. 1983, 10, 105.<br />

102. J. Ruhe, T. A. Ezquerra, G. Wegner, Synth. Met. 1989, 28, C177.<br />

103. D. Stanke, M. L. Hallensleben, L. Toppare, Synth. Met. 1995, 72, 89.<br />

104. J. Hlavaty, V. Papez, L. Kavan, P. Krtil, Synth. Met. 1994, 66, 165.<br />

105. J. Roncali, Chem. Rev. 1992, 92, 711.<br />

106. E. W. Tsai, S. Basak, J. Ruiz, J. R. Reynolds, K. Rajeshwar, J. Electrochem. Soc. 1988,<br />

135, C392.<br />

107. R. L, Elsenbaumer, K. Y. Jen, R. Oboodi, Synth. Met. 1986, 15, 169.<br />

108. S. Hotta, Synth. Met. 1987, 22, 103.<br />

109. M. J. Winokur, D. Spiegel, Y. Kim, S. Hotta, A. J. Heeger, Synth. Met. 1989, 28, C419.


Chapter 1<br />

110. C. A. S<strong>and</strong>stedt, R. D. Rieke, C. J. Eckhardt, Chem. Mater. 1995, 7, 1057.<br />

111. E. M. Genies, A. Boyle, M. Lapkowski, C. Tsintavis, Synth. Met. 1990, 36, 139.<br />

112. S. Bhadra, D. Khastgir, N. K. Singha, J. H. Lee, Prog. Polym. Sci. 2009, 34, 783.<br />

113. K. Gurunathan, A. V. Murugan, R. Marimuthu, U. P. Mulik, D. P. Amalnerkar,<br />

Mater.Chem.Phys. 1999, 61, 173.<br />

114. N. C. Foulds, C. R. Lowe, J. Chem. Soc. Far. Trans. I 1986, 82, 1259.<br />

115. M. Umana, J. Waller, Anal. Chem. 1986, 58, 2979.<br />

116. J. Y. Wong, R. Langer, D. E. Ingber, Proc. Natl. Acad. Sci. U. S. A. 1994, 91, 3201.<br />

117. G. X. Shi, M. Rouabhia, Z. X. Wang, L. H. Dao, Z. Zhang, Biomaterials 2004, 25, 2477.<br />

118. N. K. Guimard, N. Gomez, C. E. Schmidt, Prog. Poly. Sci. 2007, 32, 876.<br />

119. C. E. Schmidt, V. R. Shastri, J. P. Vacanti, R. Langer, Pro. Natl. Acad. Sci. U. S. A.1997,<br />

94, 8948.<br />

120. S. Lakard, G. Herlem, N. Valles-Villareal, G. Michel, A. Propper, T. Gharbi, B. Fahys,<br />

Biosens. Bioelectron. 2005, 20, 1946.<br />

121. P. M. George, A. W. Lyckman, D. A. LaVan, A. Hegde, Y. Leung, R. Avasare, C. Testa, P.<br />

M. Alex<strong>and</strong>er, R. Langer, M. Sur, Biomaterials 2005, 26, 3511.<br />

122. Y. Wan, H. Wu, D. J. Wen, Macromol. Biosci. 2004, 4, 882.<br />

123. H. Castano, E. A. O'Rear, P. S. McFetridge, V. I. Sikavitsas, Macromol. Biosci. 2004, 4,<br />

785.<br />

124. X. P. Jiang, Y. Marois, A. Traore, D. Tessier, L. H. Dao, R. Guidoin, Z. Zhang, Tissue<br />

Engineering 2002, 8, 635.<br />

125. A. Boyle, E. Genies, M. Fouletier, J. Electroanal. Chem. 1990, 279, 179.<br />

126. M. Pyo, G. Maeder, R. T. Kennedy, J. R. Reynolds, J. Electroanal.l Chem. 1994, 368,<br />

329.<br />

127. M. Pyo, J. R. Reynolds, Chem. Mater. 1996, 8, 128.<br />

128. A. J. Hodgson, M. J. John, T. Campbell, A. Georgevich, S. Woodhouse, T. Aoki, N.<br />

Ogata, G. G. Wallace, In Smart Materials Technologies <strong>and</strong> Biomimetics-Smart<br />

Structures <strong>and</strong> Materials 1996, Crowson, A., Ed. 1996, 2716, 164.<br />

129. J. H. Collier, J. P. Camp, T. W. Hudson, C. E. Schmidt, J. Biomed. Mater. Res. 2000, 50,<br />

574.<br />

130. D. D. Ateh, P. Vadgama, H. A. Navsaria, Tissue Engineering 2006, 12, 645.<br />

131. N. Gomez, J. Y. Lee, J. D. Nickels, C. E. Schmidt, Adv. Funct. Mater. 2007, 17, 1645.<br />

‐ 54 ‐


- 55 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

132. M. Mattioli-Belmonte, G. Giavaresi, G. Biagini, L. Virgili, M. Giacomini, M. Fini, F.<br />

Giantomassi, D. Natali, P. Torricelli, R. Giardino, Int. J. Artif. Organs 2003, 26, 1077.<br />

133. Y. Wei, P. I. Lelkes, A. G. MacDiarmid, E. Guterman, S. P. K. Cheng,; P. Bidez,<br />

Contemporary Topics in Advanced Polymer Science <strong>and</strong> Technology. Peking University<br />

Press: Beijing, 2004.<br />

134. M. Y. Li, Y. Guo, Y. Wei, A. G. MacDiarmid, P. I. Lelkes, Biomaterials 2006, 27, 2705.<br />

135. P. Wang, K. L. Tan, F. Zhang, E. T. Kang, K. G. Neoh, Chem. Mater. 2001, 13, 581.<br />

136. Y. Guo, M. Y. Li, A. Mylonakis, J. J. Han, A. G. MacDiarmid, X. S. Chen, P. I. Lelkes, Y.<br />

Wei, Biomacromolecules 2007, 8, 3025.<br />

137. L. H. Huang, J. Hu, L. Lang, X. Wang, P. B. Zhang, X. B. Jing, X. H. Wang, X. S. Chen,<br />

P. I. Lelkes, A. G. MacDiarmid, Y. Wei, Biomaterials 2007, 28, 1741.<br />

138. L. H. Huang, X. L. Zhuang, J. Hu, L. Lang, P. B. Zhang, Y. S. Wang, X. S. Chen, Y. Wei,<br />

X. B. Jing, Biomacromolecules 2008, 9, 850.<br />

139. J. Hu, L. H. Huang, X. L. Zhuang, P. B. Zhang, L. Lang, X. S. Chen, Y. Wei, X. B. Jing,<br />

Biomacromolecules 2008, 9, 2637.<br />

140. M. Waugaman, B. Sannigrahi, P. McGeady, I. M. Khan, Eur. Polym. J. 2003, 39, 1405.<br />

141. M. Gerard, A. Chaubey, B. D. Malhotra, Biosens. Bioelectron. 2002, 17, 345.<br />

142. E. Bakker, M. Telting-Diaz, Anal. Chem. 2002, 74, 2781.<br />

143. S. Cosnier, Electroanalysis 2005, 17, 1701.<br />

144. E. Tamiya, I. Karube, S. Hattori, M. Suzuki, K. Yokoyama, Sens. Actuators. 1989, 18,<br />

297.<br />

145. T. Ahuja, I. A. Mir, D. Kumar, Rajesh, Biomaterials 2007, 28, 791.<br />

146. J. Chen, B. Winther-Jensen, C. Lynam, O. Ngamna, S, Moulton, W. M. Zhang, G. G.<br />

Wallace, Electrochem. Solid. State. Lett. 2006, 9, H68.<br />

147. M. I. Rodriguez, E. C. Alocilja, Ieee. Sens. J. 2005, 5, 733.<br />

148. M. Mir, I. Katakis, Anal. Bioanal. Chem. 2005, 381, 1033.<br />

149. X. H. Jiang, X. Q. Lin, Anal. Chim. Acta. 2005, 537, 145.<br />

150. M. Wilchek, E. A. Bayer, Anal. Biochem. 1988, 171, 1.<br />

151. L. M. T. Rodrigez, M. Billon, A. Roget, G. Bidan, J.Electroanal. Chem. 2002, 523, 70.<br />

152. S. Cosnier, R. E. Ionescu, S. Herrmann, L. Bouffier, M. Demeunynck, R. S. Marks, Anal.<br />

Chem. 2006, 78, 7054.<br />

153. T. Livache, A. Roget, E. Dejean, C. Barthet, G. Bidan, R. Teoule, Nucleic. Acids. Res.


Chapter 1<br />

1994, 22, 2915.<br />

154. N. Lassalle, A. Roget, T. Livache, P. Mailley, E. Vieil, Talanta 2001, 55, 993.<br />

155. A. I. Minett, J. N. Barisci, G. G. Wallace, React. Funct. Polym. 2002, 53, 217.<br />

156. A. I. Minett, J. N. Barisci, G. G. Wallace, Anal. Chim. Acta. 2003, 475, 37.<br />

157. M. Hiller, C. Kranz, J. Huber, P. Bauerle, W. Schuhmann, Adv. Mater. 1996, 8, 219.<br />

158. H. Korri-Youssoufi, B. Makrouf, Anal. Chim. Acta. 2002, 469, 85.<br />

159. M. Bera-Aberem, H. A. Ho, M. Leclerc, Tetrahedron 2004, 60, 11169.<br />

160. J. W. Lee, F. Serna, J. Nickels, C. E. Schmidt, Biomacromolecules 2006, 7, 1692.<br />

161. B. S. Ebarvia, S. Cabanilla, F. Sevilla, Talanta 2005, 66, 145.<br />

162. F. Le Floch, H. A. Ho, M. Leclerc, Anal. Chem. 2006, 78, 4727.<br />

163. A. A. Entezami, B. Massoumi, Iran. Polym. J. 2006, 15, 13.<br />

164. M. R. Abidian, D. H. Kim, D. C. Martin, Adv. Mater. 2006, 18, 405.<br />

165. Y. L. Li, K. G. Neoh, E. T. Kang, J. Biomed. Mater. Res., Part A 2005, 73A, 171.<br />

166. T. F. Otero, J. M. Sansinena, Bioelectrochem. Bioener. 1997, 42, 117.<br />

167. T. F. Otero, M. T. Cortes, Sens.Actuators. B Chem. 2003, 96, 152.<br />

168. M. R. G<strong>and</strong>hi, P. Murray, G. M. Spinks, G. G. Wallace, Synth. Met. 1995, 73, 247.<br />

169. V. Mottaghitalab, B. B. Xi, G. M. Spinks, G. G. Wallace, Synth. Met. 2006, 156, 796.<br />

170. E. Smela, N. Gadegaard, Adv. Mater. 1999, 11, 953.<br />

171. L. M. Low, S. Seetharaman, K. Q. He, M. J. Madou, Sens. Actuators. B Chem. 2000, 67,<br />

149.<br />

172. E. Smela, Adv. Mater. 2003, 15, 481.<br />

173. O. Ouerghi, A. Touhami, N. Jaffrezic-Renault, C. Martelet, H. Ben Ouada, S. Cosnier,<br />

Bioelectrochemistry 2002, 56, 131.<br />

174. B. Fink, S. Dikalov, E. Bassenge, Free. Radical. Bio. Med. 2000, 28, 121.<br />

175. C. J. Hawker, A. W. Bosman, E. Harth, Chem. Rev. 2001, 101, 3661.<br />

176. Y. Katayama, N. Soh, M. Maeda, Chemphyschem 2001, 2, 655.<br />

177. R. A. Sheldon, I. W. C. E. Arends, Adv. Synth. Catal. 2004, 346, 1051.<br />

178. H. Nishide, T. Suga, Electrochim. Soc. Interface 2005, 14, 32.<br />

179. T. Osa, U. Akiba, I. Segawa, J. M. Bobbitt, Chem. Lett. 1988, 8, 1423.<br />

180. W. M. Xu, L. Z. Liu, M. Loizidou, M. Ahmed, I. G. Charles, Cell. Res. 2002, 12, 311.<br />

181. M. Valko, D. Leibfritz, J. Moncol, M. T. D. Cronin, M. Mazur, J. Telser, Int. J. Biochem.<br />

Cell. Biol. 2007, 39, 44.<br />

‐ 56 ‐


182. R. B. Weller, J. Invest. Dermatol. 2009, 129, 2335.<br />

183. H. Nishide, K. Oyaizu, Science 2008, 319, 737.<br />

- 57 -<br />

Polymerization <strong>and</strong> Applications <strong>of</strong> <strong>Biodegradable</strong> Polyesters<br />

184. K. Oyaizu, Y. Ando, H. Konishi, H. Nishide, J. Am. Chem. Soc. 2008, 130, 14459.<br />

185. Y. Takahashi, N. Hayashi, K. Oyaizu, K. Honda, H. Nishide, Polym. J. 2008, 40, 763.<br />

186. K. Koshika, N. Sano, K. Oyaizu, H. Nishide, Macromol. Chem. Phys. 2009, 210, 1989.<br />

187. K. Koshika, N. Sano, K. Oyaizu, H. Nishide, Chem. Commun. 2009, 7, 836.<br />

188. H. Nishide, K. Koshika, K. Oyaizu, Pure. Appl.Chem. 2009, 81, 1961.<br />

189. K. Oyaizu, H. Nishide, Adv. Mater. 2009, 21, 2339.<br />

190. T. Suga, H. Ohshiro, S. Sugita, K. Oyaizu, H. Nishide, Adv. Mater. 2009, 21, 1627.<br />

191. K. Koshika, N. Chikushi, N. Sano, K. Oyaizu, H. Nishide, Green. Chem. 2010, To be<br />

published.<br />

192. Y. Yonekuta, K. Susuki, K. C. Oyaizu, K. J. Honda, H. Nishide, J. Am. Chem. Soc. 2007,<br />

129, 14128.<br />

193. M. D. Lang, C. C. Chu, J. Polym. Sci., Part A: Polym. Chem. 2001, 39, 4214.<br />

194. G. Jokhadze, M. Machaidze, H. Panosyan, C. C. Chu, R. Katsarava, J. Biomater. Sci.<br />

Polym. Ed. 2007, 18, 411.<br />

195. L. Li, C. C. Chu, J. Biomater. Sci. Polym. Ed. 2009, 20, 341.<br />

196. C. C. Chu, "Polymeric Biomaterials" 2nd Edition 2002, Marcel Dekker, Inc. Chapter 5.<br />

197. C. S. Winalski, S. Shortkr<strong>of</strong>f, R. V. Mulkern, E. Schneider, G. M. Rosen, Magn. Reson.<br />

Med. 2002, 48, 965.<br />

198. H. Hayashi, S. Karasawa, A. Tanaka, K. Odoi, K. Chikama, H. Kuribayashi, N. Koga,<br />

Magn. Reson. Chem. 2009, 47, 201.<br />

199. Y. Huang, A. J. Nan, G. M. Rosen, C. S. Winalski, E. Schneider, P. Tsai, H. Gh<strong>and</strong>ehari,<br />

Macromol. Biosci. 2003, 3, 647.<br />

200. M. Gussoni, F. Greco, P. Ferruti, E. Ranucci, A. Ponti, L. Zetta, New. J. Chem. 2008, 32,<br />

323.<br />

201. K. I. Matsumoto, S. Subramanian, R. Murugesan, J. B. Mitchell, M. C. Krishna,<br />

Antioxid. Redox Signaling. 2007, 9, 1125.<br />

202. G. Yan, L. Peng, S. Q. Jian, L. Li, S. E. Bottle, Chin. Sci, Bull. 2008, 53, 3777.<br />

203. K. Saito, S. Kazama, H. Tanizawa, T. Ito, M. Tada, T. Ogata, H. Yoshioka, Biosci.<br />

Biotech. Biochem. 2001, 65, 787.<br />

204. T. Yoshitomi, D. Miyamoto, Y. Nagasaki, Biomacromolecules 2009.


Chapter 1<br />

205. T. Yoshitomi, R. Suzuki, T. Mamiya, H. Matsui, A. Hirayama, Y. Nagasaki, Bioconjug.<br />

Chem. 2009, 20, 1792.<br />

206. C. S. Xiao, H. Y. Tian, X. L. Zhuang, X. S. Chen, X. B. Jing, Sci. Chin. Ser. B Chem.<br />

2009, 52, 117.<br />

207. E. Yoshida, T. Tanaka, Colloid. Polym. Sci. 2006, 285, 135.<br />

208. E. Yoshida, T. Tanaka, Colloid. Polym. Sci. 2008, 286, 827.<br />

‐ 58 ‐


Chapter 2<br />

<strong>Synthesis</strong> <strong>and</strong> Electrochemistry <strong>of</strong> Schiff Base Cobalt(III)<br />

Complexes <strong>and</strong> Their <strong>Catalytic</strong> Activity for Copolymerization <strong>of</strong><br />

Epoxide <strong>and</strong> Carbon Dioxide<br />

2.1 Introduction<br />

2.2 <strong>Synthesis</strong> <strong>and</strong> <strong>Characterization</strong> <strong>of</strong> L-Co III -dnp Complexes<br />

2.3 Copolymerization <strong>of</strong> CO2 <strong>and</strong> rac-PO<br />

2.4 Electrochemistry <strong>of</strong> Cobalt(III) Complexes<br />

2.5 Conclusions<br />

2.6 Experimental Section<br />

References


Chapter 2<br />

2.1 Introduction<br />

Polycarbonate is one <strong>of</strong> the important thermoplastic with a good heat resistance, excellent<br />

optical <strong>and</strong> electrical properties <strong>and</strong> the unique biodegradability. 1-2 The mechanical properties<br />

are predominated by the stereostructure <strong>of</strong> the regioselectivity <strong>of</strong> the linear carbonate linkage<br />

in the polycarbonate. The enantioselectivity <strong>of</strong> the polycarbonate, head-to-tail connectivity,<br />

<strong>and</strong> carbonate linkage are important factors to determine the thermal stability <strong>and</strong> the<br />

mechanical properties. The reactivity <strong>and</strong> enantioselectivity are mainly related to the type <strong>and</strong><br />

chemical stereostructure <strong>of</strong> the catalyst system. Several efficient catalyst systems, such as zinc,<br />

aluminum, cadmium, manganese, cobalt, chromium <strong>and</strong> rare-earth-metal complexes, in<br />

combination with other reagents for the copolymerization <strong>of</strong> CO2 <strong>and</strong> alicyclic epoxides such<br />

as propylene oxide have been developed. 3-7 Recently, salen-type catalysts have become the<br />

focus <strong>of</strong> many groups. The chemical structures <strong>of</strong> the complexes as catalysts are related to the<br />

overall catalytic copolymerization performance <strong>of</strong> CO2 with an epoxide. 8 Transition metal<br />

Schiff-base complexes with N,N,O,O-tetradentate lig<strong>and</strong>s have been extensively studied, 9-16<br />

because <strong>of</strong> their various applications 17-19 <strong>and</strong> importance in the area <strong>of</strong> coordination<br />

chemistry. 20 Cobalt(III) Schiff-base complexes are characterized by the asymmetric structure<br />

containing axial lig<strong>and</strong>s, which are frequently labile, <strong>and</strong> exhibit high activities <strong>and</strong> yet a<br />

significant stability. 21 Especially, binary catalytic system <strong>of</strong> many forms <strong>of</strong> the<br />

(salen)Co(III)/Lewis base as catalysts is widely used in the fixation <strong>of</strong> CO2 with epoxide to<br />

form polycarbonates. 22-26 Recent studies have focused on how to improve the efficiency <strong>of</strong> the<br />

copolymerization <strong>and</strong> to enhance the catalytic activity <strong>of</strong> these complexes. 22, 24, 25, 27-30 In this<br />

chapter, we describe the electrochemical properties <strong>of</strong> several structurally related cobalt<br />

complexes <strong>and</strong> the X-ray crystal structure, which provided an important insight into the<br />

copolymerization mechanism <strong>of</strong> CO2 <strong>and</strong> propylene oxide catalyzed by these complexes.<br />

To elucidate factors that dominate the cobalt-catalyzed polymerization, a series <strong>of</strong> cobalt<br />

complexes as catalyst containing N,N,O,O-tetradentate Schiff bases were synthesized. These<br />

complexes were different in bridging lengths between the two nitrogen atoms in the Schiff<br />

base lig<strong>and</strong>s. The modification <strong>of</strong> the electronic <strong>and</strong> steric environments by altering the<br />

substituents on the diimine-bridge <strong>of</strong> the salen lig<strong>and</strong> around the metal center significantly<br />

affected the catalytic activity for the poly(propylene carbonate) formation, which was<br />

successfully ascribed to the stability <strong>of</strong> the six-coordinate cobalt complex <strong>and</strong> the lability <strong>of</strong><br />

the propagating polymer chain as the axial lig<strong>and</strong>.<br />

2.2 <strong>Synthesis</strong> <strong>and</strong> <strong>Characterization</strong> <strong>of</strong> L-CoIII-dnp Complexes<br />

2.2.1 <strong>Synthesis</strong> <strong>of</strong> the Lig<strong>and</strong>s<br />

‐ 60 ‐


<strong>Synthesis</strong> <strong>and</strong> Electrochemistry <strong>of</strong> Schiff Base Cobalt(III) Complexes <strong>and</strong> Their<br />

<strong>Catalytic</strong> Activity for Copolymerization <strong>of</strong> Epoxide <strong>and</strong> Carbon Dioxide<br />

The Schiff base lig<strong>and</strong>s were synthesized by the condensation <strong>of</strong> the corresponding<br />

diamine <strong>and</strong> 3,5-di-tert-butylsalicylaldehyde in a 1:2 molar ratio in methanol (Scheme 2.1).<br />

The lig<strong>and</strong>s were further purified by recrystallization from ethanol.<br />

2.2.2 <strong>Synthesis</strong> <strong>of</strong> the Complexes<br />

The L-Co III -dnp complex was prepared by the reaction <strong>of</strong> cobalt acetate tetrahydrate <strong>and</strong><br />

the lig<strong>and</strong>, followed by subsequent oxidation in the presence <strong>of</strong> 1 equiv. <strong>of</strong> 2,4-dinitrophenol<br />

<strong>and</strong> an excess amount <strong>of</strong> oxygen. The formation <strong>of</strong> the Schiff base lig<strong>and</strong> <strong>and</strong> their complexes<br />

have been confirmed by spectroscopic methods <strong>and</strong> X-ray diffraction experiments (vide infra).<br />

All the Schiff base complexes were stable at room temperature, <strong>and</strong> were soluble in<br />

methylene chloride, chlor<strong>of</strong>orm <strong>and</strong> DMSO.<br />

Scheme 2.1 Preparation <strong>of</strong> the L-Co III -dnp complexes.<br />

2.2.3 Complexes Structure <strong>Characterization</strong><br />

The molecular structure <strong>of</strong> L 1 -Co III -dnp (Figure 2.1) revealed that the complex was<br />

monomeric with a six-coordinated central cobalt atom in the solid state.<br />

‐ 61 ‐


Chapter 2<br />

Figure 2.1 Crystal structure <strong>of</strong> the L 1 -Co III -dnp complex.<br />

The geometry around the cobalt atom was octahedral with an almost linear axial<br />

O(2)–Co(1)–O(4) bond angle <strong>of</strong> 177.52(7)º, <strong>and</strong> nearly perpendicular for O(1)–Co(1)–O(3),<br />

O(3)–Co(1)–N(2), N(2)–Co(1)–N(1) <strong>and</strong> N(1)–Co(1)–O(1) <strong>of</strong> 88.03(7)º, 94.96(8)º, 83.29(8)º,<br />

<strong>and</strong> 93.71(7)º, respectively, for the equatorial donor atoms. The Co(1) atom deviated from the<br />

N(1)N(2)O(1)O(3) least-squares plane by only 0.0276 Å, representing the nearly ideal<br />

octahedral arrangement. The distances from the Co atom to the O(1), O(2), O(3), O(4), N(1)<br />

<strong>and</strong> N(2) atoms were 1.8969(15), 1.8868(15), 1.8964(16), 1.9672(16), 1.8985(18) <strong>and</strong><br />

1.902(18) Å, respectively (Table 2.1). Interestingly, the phenolate oxygen, O(3), is involved in<br />

the equatorial planed rather than the four donor atoms in the tetradentate Schiff-base lig<strong>and</strong>.<br />

Table 2.1 Selected bond distances <strong>and</strong> angles for L 1 -Co III -dnp.<br />

Bond Distance(Å) Bond Angle(º) Bond Angle(º)<br />

Co-N(1) 1.8985(18) O(1)-Co-O(2) 92.54(7) O(1)-Co-N(2) 176.99(7)<br />

Co-O(1) 1.8969(15) O(1)-Co-N(1) 93.71(7) O(2)-Co-N(1) 96.37(7)<br />

Co-O(3) 1.8964(16) N(2)-Co-N(1) 83.29(3) O(1)-Co-O(3) 88.03(7)<br />

Co-N(2) 1.9020(18) O(2)-Co-O(3) 86.58(6) N(2)-Co-O(3) 94.96(8)<br />

Co-O(2) 1.8868(15) N(1)-Co-O(3) 176.50(7) O(1)-Co-O(4) 87.41(7)<br />

Co-O(4) 1.9672(16) O(2)-Co-O(4) 177.52(7) N(2)-Co-O(4) 92.16(7)<br />

N(1)-Co-O(4) 86.11(7) O(3)-Co-O(4) 90.93(7)<br />

O(2)-Co-N(2) 88.03(7)<br />

The IR spectral data <strong>of</strong> the lig<strong>and</strong>s <strong>and</strong> the complexes in Table 2.2 showed major b<strong>and</strong>s<br />

around 1624 cm -1 assigned to the C=N stretching vibration (vC=N), <strong>and</strong> around 1610 cm -1 <strong>and</strong><br />

1525 cm -1 assigned to the ring vibrations. The peaks around 1566 cm -1 <strong>and</strong> 1327 cm -1 were<br />

assigned to the stretching <strong>of</strong> the N=O bonds, <strong>and</strong> the peaks around 1252 cm -1 <strong>and</strong> 1063 cm -1<br />

‐ 62 ‐


<strong>Synthesis</strong> <strong>and</strong> Electrochemistry <strong>of</strong> Schiff Base Cobalt(III) Complexes <strong>and</strong> Their<br />

<strong>Catalytic</strong> Activity for Copolymerization <strong>of</strong> Epoxide <strong>and</strong> Carbon Dioxide<br />

were definitely assigned to the stretching <strong>of</strong> the C–O <strong>and</strong> C–N bonds, respectively. 31, 32 The<br />

spectra <strong>of</strong> the complexes revealed that these vibrations shifted to lower wave numbers<br />

compared to the corresponding b<strong>and</strong>s observed for the free lig<strong>and</strong>s, as a result <strong>of</strong> the decrease<br />

in bond order upon complexation. Also, the characteristic b<strong>and</strong>s for the out-<strong>of</strong>-plane<br />

deforming, δOH, <strong>of</strong> the free lig<strong>and</strong>s were absent in the case <strong>of</strong> the complexes, as a consequence<br />

<strong>of</strong> the involvement <strong>of</strong> the oxygen anion in the bonds with the cobalt atom. The formation <strong>of</strong><br />

the cobalt–oxygen bonds led to major absorption b<strong>and</strong>s around 571 cm -1 assigned to a vCo–O<br />

vibration. 33 The IR spectra clearly showed the formation <strong>of</strong> the complexes by the coordination<br />

33, 34<br />

<strong>of</strong> cobalt(III) to the phenolic oxygens <strong>and</strong> the C=N nitrogens.<br />

Table 2.2. Infrared spectroscopic data for L-Co III -dnp.<br />

Complex vCo–O vC–N vC–O vC=N vAr-ring vN=O<br />

L 1 -Co III -dnp<br />

L 2 -Co III -dnp<br />

571 1063 1252 1624 1610 1566<br />

1525 1327<br />

579 1065 1257 1626 1596 1331<br />

1521 1569<br />

L 3 -Co III -dnp 583 1056 1269 1625 1597 1555<br />

1525 1310<br />

L 4 -Co III -dnp 575 1061 1261 1625 1601 1568<br />

1522 1318<br />

The –OH proton signals in the 1 H NMR spectra <strong>of</strong> the free lig<strong>and</strong>s (13.75, 13.58, 13.58,<br />

<strong>and</strong> 13.82 ppm for H2L 1 , H2L 2 , H2L 3 <strong>and</strong> H2L 4 , respectively) disappeared in the complexes<br />

indicating that the –OH group was deprotonated <strong>and</strong> bonded to the metal ion as an oxygen<br />

anion. One 2, 4-dinitrophenoxide counter ion was involved in each cobalt(III) complex to<br />

maintain electroneutrality. Comparing the 1 H NMR spectra <strong>of</strong> the free lig<strong>and</strong>s with the<br />

complexes, the imine proton resonances <strong>of</strong> the lig<strong>and</strong>s (8.29, 8.44, 8.34 <strong>and</strong> 8.40 ppm in H2L 1 ,<br />

H2L 2 , H2L 3 <strong>and</strong> H2L 4 , respectively) 15 shifted to lower fields upon the complexation (8.62,<br />

9.01, 8.41 <strong>and</strong> 8.89 ppm, respectively), as a result <strong>of</strong> the deshielding effect. The data from the<br />

elemental analysis <strong>and</strong> ESI MS also confirmed the structures <strong>of</strong> the complexes.<br />

2.3 Copolymerization <strong>of</strong> CO2 <strong>and</strong> rac-PO<br />

2.3.1 Mechanism for CO2/rac-PO Copolymerization<br />

The molecular structure <strong>of</strong> L 1 -Co III -dnp clearly demonstrated that the cobalt(III) tends to<br />

form a six coordination complex, which strongly suggested that all <strong>of</strong> the relevant<br />

L-Co III -dnp complexes characterized by the axial 2,4-dinitrophenolate lig<strong>and</strong>s were<br />

‐ 63 ‐


Chapter 2<br />

six-coordinated in the solid state. Furthermore, end group analysis <strong>of</strong> the obtained polymer<br />

suggested that the six-coordination arrangement was maintained even in solution, i.e., in the<br />

catalytic cycle. The isolated poly(carbonate) obtained by the L 1 -Co III -dnp/Bu4NBr catalyst<br />

was determined by the presence <strong>of</strong> a 2,4-dinitrophenoxyl moiety as the end group <strong>of</strong> the<br />

polymer chain (Figure 2.2).<br />

Figure 2.2 1 H NMR spectrum <strong>of</strong> the poly(carbonate) catalyzed by L 1 -Co III -dnp/ Bu4NBr (298K,<br />

(CD3)2SO, 500 MHz). The enlarged spectrum between 9.0 <strong>and</strong> 7.4 ppm indicated that the polymer<br />

ended with the 2,4-dinitrophenoxyl group.<br />

The good agreement <strong>of</strong> the number-averaged molecular weights based on the end-group<br />

analysis (Mn= 2.64 × 10 4 ) <strong>and</strong> that determined by the GPC experiment (Mn = 1.40 × 10 4 )<br />

suggested the polymerization mechanism in which the monomer was inserted to the<br />

cobalt-oxygen bond <strong>of</strong> the propagating end, as shown in Scheme 2.2.<br />

Scheme 2.2 Proposed mechanism for CO2/rac-PO copolymerization using the<br />

L-Co III -dnp/Bu4NBr catalyst systems.<br />

‐ 64 ‐


<strong>Synthesis</strong> <strong>and</strong> Electrochemistry <strong>of</strong> Schiff Base Cobalt(III) Complexes <strong>and</strong> Their<br />

<strong>Catalytic</strong> Activity for Copolymerization <strong>of</strong> Epoxide <strong>and</strong> Carbon Dioxide<br />

The Lewis basic monomer coordinated to the metal center in the axial site transferred to<br />

the propagating metal-polymer chain, thereby labilizing the metal alkoxide bond <strong>and</strong><br />

facilitating the insertion <strong>of</strong> CO2. 3, 15, 22-25 The rac-PO monomer might be first favorably<br />

inserted into the Co−OR bond <strong>of</strong> the Schiff base cobalt complexes, followed by the insertion<br />

<strong>of</strong> the CO2 monomer, as shown in Scheme 2.2. Subsequent alternating copolymerization <strong>of</strong><br />

rac-PO <strong>and</strong> CO2 afforded the alternating repeating unit structure to yield the polycarbonate.<br />

The side reaction to yield the cyclic carbonate could take place through the backbiting<br />

degradation <strong>of</strong> the growing polymer-catalyst complex, which may lead to the lower polymer<br />

yield. A similar mechanism has been proposed by Nguyen <strong>and</strong> co-workers using cobalt (salen)<br />

<strong>and</strong> N, N-dimethylaminoquinoline for the copolymerization <strong>of</strong> CO2 <strong>and</strong> PO. 15<br />

2.3.2 Complex Structure <strong>and</strong> Polymerization<br />

The copolymerizations <strong>of</strong> CO2 <strong>and</strong> rac-PO catalyzed by the series <strong>of</strong> L-Co III -dnp/Bu4NBr<br />

catalysts were studied, <strong>and</strong> the results are summarized in Table 2.3. The diimine-bridge (X)<br />

between the two nitrogen atoms in the Schiff bases significantly affected the catalytic activity.<br />

With X being (R,R)-1,2-cyclohexanediamine (L 1 -Co III -dnp in Scheme 2.1), the highest<br />

catalytic activity with a turn-over frequency (TOF) <strong>of</strong> 245 h -1 was accomplished (Table 2.3).<br />

Under the same conditions, when X was replaced by ethylenediimine or (R,<br />

R)-1,2-diphenylethylenediimine, the catalytic frequency was reduced to 210 <strong>and</strong> 190 h -1 for<br />

the L 2 -Co III -dnp/Bu4NBr <strong>and</strong> the L 3 -Co III -dnp/Bu4NBr catalysts, respectively. When X was<br />

2,2-dimethyl-1,3-propylenediimine, the L 4 -Co III -dnp/Bu4NBr catalyst showed the lowest<br />

activity. The catalytic activity <strong>of</strong> these complexes for the alternating copolymerization <strong>of</strong> CO2<br />

<strong>and</strong> rac-PO was in the order <strong>of</strong> L 1 -Co III -dnp > L 2 -Co III -dnp > L 3 -Co III -dnp >> L 4 -Co III -dnp,<br />

which revealed that the diimine-bridges with the three carbon atoms led to a lower activity.<br />

One could anticipate that the degree <strong>of</strong> polarization <strong>and</strong>/or the strength <strong>of</strong> the Co–O bond for<br />

the axial coordination should influence the rate <strong>of</strong> the monomer insertion <strong>and</strong> thus should<br />

affect the rate <strong>of</strong> the propagation during the polymerization. Although attempts to obtain all<br />

the molecular structures <strong>of</strong> the L-Co III -dnp complexes were not successful, the nature <strong>of</strong> the<br />

Co–O bond has been successfully determined by electrochemical methods (vide infra).<br />

Table 2.3 The catalytic activity for the copolymerization <strong>of</strong> CO2/rac-PO using the L-Co III -dnp/Bu4NBr<br />

catalyst systems. a)<br />

Run Complex<br />

TOF<br />

(h –1 ) b)<br />

Selectivity<br />

(%PPC) c)<br />

Head-to-tail<br />

Linkages (%) d)<br />

‐ 65 ‐<br />

Mn e)<br />

×10 4<br />

Mn f)<br />

×10 4<br />

1 L 1 -Co III -dnp 245 74 88 2.6 1.4 1.4<br />

PDI f)<br />

2 L 2 -Co III -dnp 210 60 71 1.4 1.1 1.3<br />

3 L 3 -Co III -dnp 190 75 82 4.0 2.6 1.2


Chapter 2<br />

4 L 4 -Co III -dnp 25 70 81 3.1 1.4 1.2<br />

a)<br />

The reaction was performed in neat rac-PO ([rac-PO]:[Co]:[Bu4NBr] = 1000:1:1, molar ratio) in a 100<br />

mL autoclave at 25 ºC with 2 MPa <strong>of</strong> CO2 for 2 h; b) Turnover frequency <strong>of</strong> rac-PO to products<br />

(polycarbonate <strong>and</strong> cyclic carbonate); c) Selectivity for PPC over PC as determined by 1 H NMR<br />

spectroscopy; d) Determined by 13 C NMR spectroscopy; e) Determined by 1 H NMR. f) Determined by GPC.<br />

2.4 Electrochemistry <strong>of</strong> Cobalt(III) Complexes<br />

The electrochemical data for the lig<strong>and</strong>s <strong>and</strong> the L-Co III -dnp complexes were shown in<br />

Table 2.4, <strong>and</strong> the representative voltammograms were presented in Figures 2.3 <strong>and</strong> 2.4. All<br />

CV curves for the lig<strong>and</strong>s were recorded in the potential range <strong>of</strong> -2.00 to 1.50 V. The<br />

voltammograms showed an irreversible oxidation peak at potentials between 1.08 <strong>and</strong> 1.20 V.<br />

These peaks were ascribed to the oxidation <strong>of</strong> the Schiff base lig<strong>and</strong>s (Figure 2.3). For the<br />

L-Co III -dnp complexes, the potential sweeps were made from -1.20 to 1.50 V. It is considered<br />

that the nature <strong>of</strong> the cobalt(III) atom in the Schiff base complexes should dominate the<br />

catalytic activity for the copolymerization <strong>of</strong> the epoxide <strong>and</strong> carbon dioxide, which prompted<br />

us to pay close attention to the Co III + e -<br />

Co II redox potential. The redox couple <strong>of</strong> Co III<br />

+ e -<br />

Co II was observed for L 1 -Co III -dnp, L 2 -Co III -dnp, L 3 -Co III -dnp, <strong>and</strong> L 4 -Co III -dnp<br />

at E1/2 = 107, 134, 163, <strong>and</strong> 32 mV, respectively (Figure 2.4). The absence <strong>of</strong> any redox<br />

responses near 0 to 0.5 V in the metal-free lig<strong>and</strong>s (Figure 2.3) allowed one to ascribe these<br />

quasi-reversible waves to the cobalt(II/III) redox couple.<br />

Figure 2.3 Cyclic voltammograms <strong>of</strong> 1 mM solutions <strong>of</strong> H2L in DMF containing 0.1 M<br />

NBu4ClO4 on a glassy carbon electrode (� 3.3 mm). Potential is shown in V vs. Ag/Ag + at the<br />

scan rate <strong>of</strong> 100mV s -1 . I: H2L 1 ; II: H2L 2 ; III: H2L 3 ; IV: H2L 4 .<br />

‐ 66 ‐


<strong>Synthesis</strong> <strong>and</strong> Electrochemistry <strong>of</strong> Schiff Base Cobalt(III) Complexes <strong>and</strong> Their<br />

<strong>Catalytic</strong> Activity for Copolymerization <strong>of</strong> Epoxide <strong>and</strong> Carbon Dioxide<br />

Figure 2.4. Cyclic voltammograms <strong>of</strong> 1 mM solutions <strong>of</strong> L-Co III -dnp in DMF<br />

containing 0.1 M NBu4ClO4 on a carbon electrode. Potential is shown in V vs.<br />

Ag/Ag + at the scan rate <strong>of</strong> 100 mV s -1 . I: L 1 -Co III -dnp; II: L 2 -Co III -dnp; III:<br />

L 3 -Co III -dnp; IV: L 4 -Co III -dnp.<br />

Co II for L-Co III -dnp (Table 2.4) revealed that the length<br />

<strong>and</strong> steric structures <strong>of</strong> the diimine-bridges between the two nitrogen atoms in the Schiff bases<br />

significantly affected the redox potential. These complexes had the same auxiliary lig<strong>and</strong>,<br />

2,4-dinitrophenolate, but different diimine-bridges. The gradual shift <strong>of</strong> E1/2 to negative<br />

potentials for the L 3 -Co III -dnp, L 2 -Co III -dnp, <strong>and</strong> L 1 -Co III -dnp complexes indicated that the<br />

election-donating character <strong>of</strong> the lig<strong>and</strong> increased in this order, stabilizing the axial Co–O<br />

(phenolate) bond in Figure 2.1, which is an advantage when forming the transition state<br />

(Scheme 2.2) <strong>and</strong> thus leads to a faster copolymerization <strong>of</strong> CO2 <strong>and</strong> PO. However,<br />

The E1/2 data <strong>of</strong> Co III + e -<br />

L 4 -Co III -dnp has the most negative E1/2 <strong>of</strong> 32 mV <strong>and</strong> yet the lowest activity for the<br />

copolymerization. It is believed that the diimine-bridges containing two carbon atoms <strong>and</strong><br />

three carbon atoms leads to a square pyramidal (sqp) <strong>and</strong> a trigonal bipyramidal (tbp)<br />

geometry for the central five-coordinated metal atom, respectively. Thus, the low catalytic<br />

activity <strong>of</strong> L 4 -Co III -dnp during the copolymerization <strong>of</strong> CO2 <strong>and</strong> PO may be ascribed to the<br />

unfavorable tbp geometry <strong>of</strong> the central cobalt atom, which prevents Bu4NBr from<br />

coordinating with cobalt in an axial position opposite to the Co–O (phenolate) bond, leading<br />

to the decrease in the activity. The utilization <strong>of</strong> carbon dioxide (CO2) has received much<br />

attention because <strong>of</strong> its potential use as an abundant, economical <strong>and</strong> biorenewable resource<br />

<strong>and</strong> its weakening <strong>of</strong> global warming or the greenhouse effect. The copolymerization <strong>of</strong> CO2<br />

as the monomer with epoxide was firstly reported by Inoue et al. in the 1960s. 35 The<br />

combination <strong>of</strong> the X-ray crystallographic <strong>and</strong> electrochemical results in the present<br />

investigation provided important insights into the factors dominating the catalytic activity<br />

‐ 67 ‐


Chapter 2<br />

useful for designing a highly efficient catalytic system.<br />

Table 2.4 Cyclic voltammogram data for the transition between Co III + e -<br />

Complex Epc (mV) E1/2 (mV) ΔE ipa/ipc<br />

L 1 -Co III -dnp 32 107 150 0.64<br />

L 2 -Co III -dnp 84 134 99 0.83<br />

L 3 -Co III -dnp 52 163 221 0.82<br />

L 4 -Co III -dnp -71 32 206 0.78<br />

‐ 68 ‐<br />

Co II <strong>of</strong> L-Co III -dnp. a)<br />

a) Determined using 1 mM solution <strong>of</strong> L-Co III -dnp in DMF in the presence <strong>of</strong> 0.1 M NBu4ClO4 at carbon<br />

electrode vs. Ag/Ag + at the scan rate <strong>of</strong> 100 mV s -1 .<br />

2.5 Conclusions<br />

A series <strong>of</strong> Cobalt Schiff-base complexes were investigated as the catalyst for the<br />

alternating copolymerization <strong>of</strong> CO2 <strong>and</strong> rac-PO in the presence <strong>of</strong> Bu4NBr. The<br />

poly(propylene carbonate) (PPC) <strong>and</strong> cyclic propylene carbonate (PC) selectivity <strong>of</strong> the<br />

resultant copolymers were determined by modification <strong>of</strong> the length <strong>of</strong> the diimine bridges<br />

between the two nitrogen atoms in the lig<strong>and</strong>s. The L 1 -Co III -dnp/Bu4NBr catalyst exhibited<br />

the highest activity, PPC/PC selectivity, <strong>and</strong> degree <strong>of</strong> head-to-tail linkages. The<br />

L 2 -Co III -dnp/Bu4NBr catalyst showed slightly lower head-to-tail linkages. For the dimine<br />

bridges containing three-carbon chains between the two nitrogen atoms in the Schiff bases,<br />

the corresponding L 4 -Co III -dnp complex displayed the lowest catalytic activity. Based on the<br />

electrochemical measurements, the half-wave potentials <strong>of</strong> the complexes were obtained <strong>and</strong><br />

the catalytic activity <strong>of</strong> these complexes was compared. The higher stability for the axial<br />

group’s metal–O (phenolate) bond indicated the more negative E1/2 <strong>of</strong> the Co III + e -<br />

redox couple in the L-Co III -dnp complexes, <strong>and</strong> the higher catalytic activity for the<br />

copolymerization <strong>of</strong> PO <strong>and</strong> CO2. In the case <strong>of</strong> the cobalt complexes with more positive<br />

Co(II/III) potentials <strong>and</strong> lower electron density on the cobalt center, the end <strong>of</strong> propagating<br />

chain to the cobalt center was considered to determine the overall polymerization rate.<br />

However, the L 4 -Co III -dnp with the most negative E1/2 possessed the lowest catalytic activity,<br />

which may be due to the unfavorable trigonal bipyramidal geometry during the transition state,<br />

preventing the Lewis bases from bonding to the metal center.<br />

2.6 Experimental Part<br />

Materials<br />

All experiments involving air- <strong>and</strong>/or water-sensitive compounds were carried out using<br />

st<strong>and</strong>ard Schlenk techniques under a dry argon atmosphere. All solvents were purified by<br />

Co II


<strong>Synthesis</strong> <strong>and</strong> Electrochemistry <strong>of</strong> Schiff Base Cobalt(III) Complexes <strong>and</strong> Their<br />

<strong>Catalytic</strong> Activity for Copolymerization <strong>of</strong> Epoxide <strong>and</strong> Carbon Dioxide<br />

st<strong>and</strong>ard procedures before use.<br />

(R,R)-N,N�-(1,2-cyclohexene)bis(3,5-di-tert-butylsalicylideneimine) (H2L 1 ),<br />

N,N�-ethylenebis(3,5-di-tert-butylsalicylideneimine) (H2L 2 ),<br />

(R,R)-N,N�-(1,2-diphenylethylethylene)bis(3,5-di-tert-butylsalicylideneimine) (H2L 3 ), <strong>and</strong><br />

N,N�-(2,2-dimethyl-1,3-propylene)bis(3,5-di-tert-butylsalicylideneimine) (H2L 4 ) were<br />

synthesized according to previously published procedures. 22, 30 All other reagents were<br />

commercially available, <strong>and</strong> were used as received.<br />

Instrumentation <strong>and</strong> Methods<br />

The NMR spectra were recorded by a Bruker AV 400M or a Bruker AV 500M instrument<br />

in CDCl3 or (CD3)2SO at room temperature. Tetramethylsilane was used as the internal<br />

st<strong>and</strong>ard. The FT-IR spectra were obtained using a Perkin-Elmer 2000 FT-IR spectrometer.<br />

The EIS-MS data were obtained using a Finnigan LCQ TM ion trap mass spectrometer. It was<br />

operated with the following parameters: positive ion <strong>and</strong> negative ion mode, the electrospray<br />

voltage at 4.5 kV, the capillary voltage at 20 V, the tube lens <strong>of</strong>fset voltage at 15 V <strong>and</strong> the<br />

capillary temperature between 80─100 . High-purity nitrogen (N2) was used as the sheath<br />

gas, <strong>and</strong> its flow rate was 60 arbitrary units. The sample solutions were injected at 5 µL min -1<br />

via a syringe pump. The GPC measurements were performed by a Waters 410 GPC with THF<br />

as the eluent (flow rate: 1 ml min -1 at 25 ). Polystyrene was used as the internal st<strong>and</strong>ard.<br />

The cyclic voltammetry experiments were performed using the BAS 660C electrochemical<br />

system. The sample solutions typically contained 1 mM <strong>of</strong> the cobalt complex in DMF <strong>and</strong><br />

tetrabutylammoium perchlorate (TBAP) (0.1 M) as the supporting electrolyte at room<br />

temperature. The cyclic voltammograms were obtained at the scan rate <strong>of</strong> 100 m V s −1 under<br />

a nitrogen atmosphere. The potential sweep ranges were between −2.00 <strong>and</strong> 1.50 V for the<br />

lig<strong>and</strong>s <strong>and</strong> between −1.20 <strong>and</strong> 1.50 V for the cobalt complexes.<br />

A suitable crystal for the X-ray diffraction experiment was obtained by slow evaporation<br />

<strong>of</strong> a saturated toluene solution <strong>of</strong> L 1 -Co III -dnp at room temperature, which was carefully<br />

collected <strong>and</strong> mounted on a Bruker SMART APEX CCD diffractometer with<br />

graphite-monochromated Mo-Kα (λ = 0.71073Å) radiation at room temperature. The crystal<br />

structure was solved by the direct method. Refinement was carried out by the full matrix<br />

least-squares methods based on F 2 using the SHELXL-97 s<strong>of</strong>tware package.<br />

Preparation <strong>of</strong> the Complexes<br />

(2,4-Dinitrophenolato)(R,R)-N,N�-(1,2-cyclohexene)bis(3,5-di-tert-butylsalicylideneimina<br />

to)]cobalt(III) (L1-CoIII-dnp):<br />

(R,R)-N,N�-1,2-Cyclohexenebis(3,5-di-tert-butylsalicylideneiminato)cobalt(II) (L 1 -Co II )<br />

‐ 69 ‐


Chapter 2<br />

was first synthesized according to the method described in a previous report 29 with some<br />

modifications as follows. A solution <strong>of</strong> cobalt acetate tetrahydrate (0.75 g, 3.0 mmol) in<br />

methanol (200 mL) was added to a solution <strong>of</strong> the lig<strong>and</strong>, H2L 1 (1.81 g, 3.0 mmol), in CH2Cl2<br />

(20 mL) via a cannula under an atmosphere <strong>of</strong> argon. A brick-red precipitate was observed<br />

before all <strong>of</strong> the cobalt acetate solution was added. The residue on the wall <strong>of</strong> the reaction<br />

flask was rinsed with methanol (20 mL), <strong>and</strong> the collected mixture was allowed to stir for a<br />

further 15 min at room temperature, <strong>and</strong> then for 30 min at 0 . The solids were collected by<br />

filtration <strong>and</strong> rinsed with cold (0 ) methanol (3 × 50 mL) before drying at 60 in a vacuum<br />

for 24 h. The product L 1 -Co II was obtained in 95% yield. FT-IR (KBr): 1610 (s, C=N), 1595<br />

(s, Ar–ring), 1527 (s, Ar–ring), 1254 (s, C–O), 572 cm -1 (w, Co–O). ESI-MS: m/z = 603.6<br />

(C36H52N2O2Co) + .<br />

(2,4-Dinitrophenolato)(R,R)-N,N�-(1,2-cyclohexene)bis(3,5-di-tert-butylsalicylideneiminat<br />

o)]cobalt(III) (L 1 -Co III -dnp) was synthesized according to a method described in the<br />

literature with modifications as follows. 22, 29 To a stirred solution <strong>of</strong> L 1 -Co II (1.48 g, 2.0 mmol)<br />

in CH2Cl2 (150 mL), 2,4-dinitrophenol (0.368 g, 2.0 mmol, 1 equiv.) in CH2Cl2 (20 mL) was<br />

added. The solution was stirred under dry oxygen at room temperature for 60 min. The<br />

solvents were removed under vacuum to leave a crude dark solid in an approximately 100%<br />

yield. The residue was further rinsed with a mixture <strong>of</strong> diethyl ether <strong>and</strong> hexane, <strong>and</strong> then<br />

dried at 60 under vacuum for 24 h, yielding a dark needle-like crystal (1.47 g, 94%). FT-IR<br />

(KBr): 1624 (s, C=N), 1610 (s, Ar–ring), 1566 (w, N=O), 1525 (s, Ar–ring), 1327 (s, N=O),<br />

1252 (s, C–O), 1063 (w, C–N), 571 cm -1 (w, Co–O). C42H55N4O7Co (786.8): Calcd. C 64.11,<br />

H 7.05, N 7.12; Found C 64.01, H 7.01, N 6.94. ESI-MS: m/z = 786.3, (C42H55N4O7Co) + ,<br />

603.5 (C36H52N2O2Co) + , 183.2 (C6H3N2O5) - . Crystal data for L 1 -Co III -dnp: C42H55CoN4O20,<br />

M = 786.83, Triclinic space group P 1.<br />

a = 11.2788(13) Å, b = 14.4552(17) Å, c = 14.5806(17)<br />

Å, � = 90.382(2), = 108.936(2), � = 111.141(2), V = 2076.5(4) Å 3 , Z = 2, Dcalcd = 1.258 g<br />

cm -3 , (Mo K ) = 0.466 mm -1 . The data were collected at 293 K. Of the 11165 reflections,<br />

7503 were unique (Rint = 0.017). The non-hydrogen atoms were anisotropically refined. The<br />

hydrogen atoms were geometrically positioned <strong>and</strong> refined as riding atoms. The final cycle <strong>of</strong><br />

the full-matrix least-squares refinement converged with R = 0.0447 <strong>and</strong> Rw = 0.1040 based on<br />

6046 reflections with I > 2(I).<br />

(2,4-Dinitrophenolato)[N,N�-ethylenebis(3,5-di-tert-butylsalicylideneiminato)]cobalt(I<br />

II) (L 2 -Co III -dnp) was synthesized by a procedure similar to that <strong>of</strong> L 1 -Co III -dnp. Yield: 89%.<br />

IR (KBr): 1626 (s, C=N), 1596 (s, Ar–ring), 1569 (w, N=O), 1521 (s, Ar–ring), 1331 (s, N=O),<br />

1257 (s, C–O), 1065 (w, C–N), 583 cm -1 (w, Co–O). C38H49N4O7Co (732.8): Calcd. C 62.29,<br />

H 6.74, N 7.65; Found C 62.12, H 6.67, N 7.52. ESI-MS: m/z = 603.5 (C32H46N2O2Co) + ,<br />

183.2 (C6H3N2O5) - .<br />

‐ 70 ‐


<strong>Synthesis</strong> <strong>and</strong> Electrochemistry <strong>of</strong> Schiff Base Cobalt(III) Complexes <strong>and</strong> Their<br />

<strong>Catalytic</strong> Activity for Copolymerization <strong>of</strong> Epoxide <strong>and</strong> Carbon Dioxide<br />

(2,4-Dinitrophenolato)[(R,R)-N,N�-1,2-diphenylethylenebis(3,5-di-tert-butylsalicylide<br />

neiminato)]cobalt(III) (L 3 -Co III -dnp) was synthesized by a method similar to that <strong>of</strong><br />

L 1 -Co III -dnp. Yield: 90%. IR (KBr): 1625 (s, C=N), 1597 (s, Ar–ring), 1555 (w, N=O), 1525<br />

(s, Ar–ring), 1310 (s, N=O), 1269 (s, C–O), 1056 (w, C–N), 571 cm -1 (w, Co–O).<br />

C50H57N4O7Co (884.9): Calcd. C 67.86, H 6.49, N 6.33; Found C 66.05, H 6.34, N 5.97.<br />

ESI-MS: m/z = 883.8, (C50H57N4O7Co) + , 701.5 (C44H54N2O2Co) + , 183.2 (C6H3N2O5) - .<br />

(2,4-Dinitrophenolato)[N,N�-2,2-dimethyl-1,3-propylenebis(3,5-di-tert-butylsalicylide<br />

neiminato)]cobalt(III) (L 4 -Co III -dnp) was synthesized by a procedure similar to that <strong>of</strong><br />

L 1 -Co III -dnp. Yield: 95%. IR (KBr): 1625 (s, C=N), 1601 (s, Ar–CH), 1568 (s, N=O), 1522 (s,<br />

Ar–ring), 1318 (s, N=O), 1261 (s, C–O), 1061 (w, C–N), 575 cm -1 (w, Co–O). C41H55N4CoO7<br />

(774.8): Calcd. C 63.55, H 7.15, N 7.23; Found C 63.39, H 7.06, N 7.14. ESI-MS: m/z =<br />

774.1 (C41H55N4O7Co) + , 591.4 (C35H52N2O2Co) + , 183.1 (C6H3N2O5) - .<br />

Typical Copolymerization Procedure<br />

A mixture <strong>of</strong> the cobalt complex(0.1 mmol) <strong>and</strong> tetra(n-butyl)ammonium bromide<br />

(Bu4NBr)(0.1 mmol) was dissolved in 3.5 mL <strong>of</strong> racemic propylene oxide (rac-PO) under a<br />

nitrogen atmosphere. The mixture was injected into a high pressure reactor equipped with a<br />

magnetic stirrer under a CO2 atmosphere. The reactor with a 2.5 MPa CO2 pressure was<br />

placed in an oil bath. The mixture was stirred at 40 for the allotted reaction time <strong>and</strong> then<br />

vented in a fume hood. A small aliquot <strong>of</strong> the polymerization mixture was sampled from the<br />

reactor for the 1 H NMR spectroscopic analysis. The remaining polymerization mixture was<br />

then dissolved in CH2Cl2, quenched with 5% HCl in methanol, <strong>and</strong> transferred to a<br />

pre-weighed vial. The product mixture was dried under vacuum to a constant weight. The<br />

crude yield was carefully determined after subtracting the catalyst weight. The product was<br />

then dissolved in CH2Cl2 <strong>and</strong> precipitated from methanol again. The polymer was collected<br />

by filtration <strong>and</strong> dried under vacuum to a constant weight. Analytical data for the typical<br />

product (with the L 1 -Co III -dnp catalyst) was as follows. IR (KBr): 1749 (s, C=O), 1609 (w,<br />

Ar-ring), 1580 (w, N=O), 1531 (w, Ar-ring), 1314 (s, N=O), 1235 cm -1 (s, C-O). 1 H NMR<br />

(500 MHz, CDCl3, ppm): = 1.24 (–CHCH3), 4.19 (–CH2CH–), 4.90 (–CH2CH–). 13 C NMR<br />

(400 MHz, CDCl3, ppm): = 16.1 (–CH3), 72.3 (–CH2–), 69.1 (–CH–), 154.1 (C=O) .<br />

References<br />

1. J. K. Lee, S. W. Park, J. W. Lee, M. R. Kim, Polym. Int. 2006, 55, 849.<br />

2. H. Sugimoto, S. Inoue, J. Polym. Sci., Part A: Polym. Chem. 2004, 42, 5561.<br />

3. D. J. Darensbourg, R. M. Mackiewicz, A. L. Phelps, D. R. Billodeaux, Acc. Chem. Res.<br />

2004, 37, 836.<br />

4. M. Cheng, E. B. Lobkovsky, G. W. Coates, J. Am. Chem. Soc. 1998, 120, 11018.<br />

‐ 71 ‐


Chapter 2<br />

5. Z. L. Quan, X. H. Wang, X. J. Zhao, F. S. Wang, Polymer 2003, 44, 5605.<br />

6. D. M. Cui , M. Nishiura , O. Tardif , Z. M. Hou, Organometallics 2008, 27, 2428.<br />

7. L. P. Guo, C. M. Wang, W. J. Zhao, H. R. Li, W. L. Sun, Z. Q. Shen, Dalton. Trans. 2009,<br />

5406.<br />

8. X. B. Lu, L. Shi, Y. M. Wang, R. Zhang, Y. J. Zhang X. J. Peng, Z. C. Zhang, B. Li, J. Am.<br />

Chem. Soc. 2006 128, 1664.<br />

9. J. Vargas, J. Costamagna, R. Latorre, A. Alvardo, G. Mena, Coord. Chem. Rev. 1992, 119,<br />

67.<br />

10. M. D. Hobday, T. D. Smith, Coord. Chem. Rev. 1972, 9, 311.<br />

11. 11a. S. Yamada, Coord. Chem. Rev. 1999, 192, 537. 11b. R. D. Jones, D. A. Summerville,<br />

F. Basolo, Chem. Rev. 1979, 79, 139.<br />

12. A. Christensen, H. S. Jensen, V. McKee, C. J. McKenzie, M. Munch, Inorg. Chem. 1997,<br />

36, 6080.<br />

13. S. H. Rehaman, H. Chowdhury, D. Bose, R. Ghosh, C. H. Hung, B. K. Ghosh, Polyhedron<br />

2005, 24, 1755.<br />

14. H. J. Kim, W. Kim, A. J. Lough, B. M. Kim, J. Chin, J. Am. Chem. Soc. 2005, 127,<br />

16776.<br />

15. R. L. Paddock, S. T. Nguyen, Macromolecules 2005, 38, 6251.<br />

16. M. Fondo, N. Ocampo, A. M. Garcia-Deibe, M. Corbella, M. S. El Fallah, J. Cano,<br />

Sanmartin, M. R. Bermejo, Dalton. Trans. 2006, 7, 4905.<br />

17. A. Chaudhary, N. Bansal, A. Gajraj, R. V. Singh, J. Inorg. Biochem. 2003, 93, 393.<br />

18. M. R. Malachowski, B. T. Dorsey, M. J. Parker, M. E. Adams, R. S. Kelly, Polyhedron<br />

1998, 17, 1289.<br />

19. H. Z. Du, A. H. Velders, P. J. Dijkstra, Z. Y. Zhong, X. S. Chen, F. J. Jan,<br />

Macromolecules 2009, 42, 1058.<br />

20. S. Ch<strong>and</strong>ra, K. Gupta, Trans. Metal. Chem. 2002, 27, 196.<br />

21. A. Bottcher, T. Takeuchi, K. I. Hardcastle, T. J. Meade, H. B. Gray, Inorg. Chem. 1997,<br />

36, 2498.<br />

22. Y. Niu, W. Zhang, X. Pang, X. Chen, X. Zhuang, X. Jing. J. Polym. Sci., Part A: Polym.<br />

Chem. 2007, 45, 5050.<br />

23. G. W. Coates, D. R. Moore, Angew. Chem., Int. Ed. 2004, 43, 6618.<br />

24. D. J. Darensbourg, J. C. Yarbrough, J. Am. Chem. Soc. 2002, 124, 6335.<br />

25. D. J. Darensbourg, J. C. Yarbrough, C. Ortiz, C. Fang, J. Am. Chem. Soc. 2003, 125, 7586.<br />

26. R. Eberhardt, M. Allmendinger, B. Rieger, Macromol. Rapid. Commun. 2003, 24, 194.<br />

27. S. Chen, Z. Hua, Z. Fang, G. Qi, Polymer 2004, 45, 6519.<br />

28. 28a. Y. Liu, K. Huang, D. Peng, H. Wu, Polymer 2006, 47, 8453; 28b. L. Lu, K. Huang, J.<br />

Polym. Sci., Part. A: Polym. Chem. 2005, 43, 2468.<br />

‐ 72 ‐


<strong>Synthesis</strong> <strong>and</strong> Electrochemistry <strong>of</strong> Schiff Base Cobalt(III) Complexes <strong>and</strong> Their<br />

<strong>Catalytic</strong> Activity for Copolymerization <strong>of</strong> Epoxide <strong>and</strong> Carbon Dioxide<br />

29. S. E. Schaus, B. D. Br<strong>and</strong>es, J. F. Larrow, M. Tokunaga, K. B. Hansen, A. E. Gould, M. E.<br />

Furrow, E. N. Jacobsen, J. Am. Chem. Soc. 2002, 124, 1307.<br />

30. D. J. Darensbourg, R. M. Mackiewicz, J. L. Rodgers, C. Fang, D. R. Billodeaux, J. H.<br />

Reibenspies, Inorg. Chem. 2004, 43, 6024.<br />

31. T. Joseph, S. B. Halligudi, C. Satyanarayan, D. P. Sawant, S. Gopinathan, J. Mol. Catal. A:<br />

Chem. 2001, 168, 87.<br />

32. K. Nakamoto, Infrared <strong>and</strong> Raman Spectra <strong>of</strong> Inorganic <strong>and</strong> Coordination Compounds,<br />

Part B, 5th ed., Wiley 1997.<br />

33. A. Ramach<strong>and</strong>raiah, P. Rao, M. Ramaiah, Indian J. Chem. A 1989, 28, 309.<br />

34. X. Tai, X. Yin, Q. Chen, M. Tan, Molecules 2003, 8, 439.<br />

35. S. Inoue, H. Koinuma, T. Tsuruta, J. Polym. Sci., Part B: Polym. Lett. 1969, 130, 210.<br />

‐ 73 ‐


Chapter 3<br />

Polymerization <strong>of</strong> Lactic O-Carboxylic Anhydride using<br />

Organometallic Catalysts<br />

3.1 Introduction<br />

3.2 Polymerization <strong>of</strong> LacOCA using Co(III) Complexes with Schiff Base Lig<strong>and</strong>s<br />

3.3 Polymerization <strong>of</strong> LacOCA using Tin(II) or Al(III) Based Complexes<br />

3.4 <strong>Characterization</strong> <strong>of</strong> the Obtained PLA<br />

3.5 Conclusions<br />

3.6 Experimental Section<br />

References


Chapter 3<br />

3.1 Introduction<br />

Poly(lactic acid) (PLA) has been receiving much attention as biodegradable materials in<br />

recent years. 1, 2 It has found many applications in the field <strong>of</strong> biomedical purposes, such as<br />

suture, drug delivery, <strong>and</strong> tissue engineering. 3-5 Because the starting material <strong>of</strong> PLA is<br />

agricultural products such as corn, PLA is an environmentally sustainable alternative to<br />

petrochemically-derived products. 6 It can be employed in the preparation <strong>of</strong> biodegradable<br />

plastics. PLA can be prepared by directly polycondensation <strong>of</strong> lactic acid. 7 However, it is<br />

usually difficult to obtain high molecular weight PLA by polycondensation. Currently,<br />

ring-opening polymerization (ROP) <strong>of</strong> lactide is a preferred route to prepare PLA because <strong>of</strong><br />

the higher controllability <strong>of</strong> the polymerization. 1 Indeed, high molecular-weight PLA can be<br />

obtained by ROP. The driving force for the ROP <strong>of</strong> lactide is the ring strain <strong>of</strong> the monomer.<br />

As a polymerizable six-membered ring, the ring strain <strong>of</strong> lactides is modest. Therefore, high<br />

temperatures are usually required for the ROP <strong>of</strong> lactides. 8 Undesirable transesterification <strong>and</strong><br />

racemization <strong>of</strong>ten take place in these cases, 1 which significantly reduce the mechanical<br />

properties <strong>of</strong> the corresponding PLA. Consequently, a polymerizable activated monomer <strong>of</strong><br />

the lactide equivalent is highly desirable. 5-Methyl-1,3-dioxolane-2,4-dione is a<br />

five-membered O-carboxyanhydride (LacOCA) derived from the lactic acid (Figure 3.1).<br />

Figure 3.1 Polymerization <strong>of</strong> LacOCA.<br />

It can be readily obtained by the reaction <strong>of</strong> lactate salt with diphosgene. 9 Bourissou et al.<br />

reported that the ROP <strong>of</strong> LacOCA proceeded in the presence <strong>of</strong> an organocatalyst,<br />

dimethylaminopyridine (DMAP) 10, 11 or lipase. 12 The ROP <strong>of</strong> LacOCA catalyzed by DMAP is<br />

well controllable. PLA with high molecular weight <strong>and</strong> narrow polydispersity index has been<br />

obtained under mild polymerization conditions. This indicates that the ROP <strong>of</strong> LacOCA is an<br />

alternative route to prepare PLA. 10 Organometallic complexes have been widely used as<br />

catalysts in a variety <strong>of</strong> polymerization. For instance, stannous octoate, 8 alkylaluminum,<br />

aluminum alkoxides, 13-18 zinc alkyl, 19 calcium alkoxides, 20, 21 <strong>and</strong> strontium alkoxide 22 have<br />

been used as the catalyst/initiator for the ROP <strong>of</strong> lactides. However, to the best <strong>of</strong> our<br />

‐ 76 ‐


Polymerization <strong>of</strong> Lactic O‐Carboxylic Anhydride using Organometallic Catalysts<br />

knowledge, there has been no report on the ROP <strong>of</strong> LacOCA using organometallic complexes<br />

as catalysts. Considering the high toxicity <strong>of</strong> DMAP, it is important to explore the<br />

organometallic complexes as the catalyst <strong>of</strong> the ROP <strong>of</strong> LacOCA. We have focused on<br />

Co(III) complexes with Schiff base lig<strong>and</strong>s as the versatile <strong>and</strong> highly active catalysts, based<br />

on their potential activities to catalyze many kinds <strong>of</strong> reactions including Baeyer-Villiger<br />

oxidation, 23 hydrolysis <strong>of</strong> epoxides, 24 cyclizations <strong>of</strong> dianhydro sugar alcohols, 25 <strong>and</strong><br />

copolymerization <strong>of</strong> epoxides <strong>and</strong> carbon dioxide. 26 Tin(II) aliphatates 8 <strong>and</strong> Al(III)<br />

complexes with Schiff base lig<strong>and</strong>s 27, 28 have also been reported as the conventional catalysts<br />

for the ROP <strong>of</strong> lactides (Figure 3.2). In this chapter, the ROP <strong>of</strong> LacOCA using Co(III)<br />

complexes with Schiff base lig<strong>and</strong>s, Tin(II) aliphatates <strong>and</strong> Al(III) complexes with Schiff<br />

base lig<strong>and</strong>s as catalysts were described in detail.<br />

Figure 3.2 Structures <strong>of</strong> Co(III) complexes with Schiff base lig<strong>and</strong>s employed in this study.<br />

3.2 Polymerization <strong>of</strong> LacOCA using Co(III) Complexes with Schiff Base Lig<strong>and</strong>s<br />

Polymerization <strong>of</strong> LacOCA using Co(III) complexes with Schiff base lig<strong>and</strong>s were<br />

performed at 70°C in toluene. The results <strong>of</strong> the polymerization were listed in Table 3.1.<br />

Poly(lactic acid) (PLA) was obtained in all experiments using the Co(III) complexes with<br />

Schiff base lig<strong>and</strong>s as the catalyst. When the monomer/initiator ratio was 100/1 (Table 3.1),<br />

the number average molecular weight (Mn) <strong>of</strong> PLA was in the range <strong>of</strong> 3.3 – 9.6×10 3 , <strong>and</strong> the<br />

‐ 77 ‐


Chapter 3<br />

polydispersity index (PDI) <strong>of</strong> PLA was in the range <strong>of</strong> 1.1–1.5. The Mn <strong>of</strong> the obtained PLA<br />

was low in all experiments except the Co(III)-c-(NO2)2 initiating system, which gave PLA<br />

with a high Mn (9.6×10 3 ). The molecular weight distribution <strong>of</strong> the product obtained with the<br />

Co(III)-c-(NO2)2 catalyst was relatively broad (PDI = 1.4). When the Co(III)-b-NO2/ i PrOH<br />

initiating system (where i PrOH was isopropanol) was employed, the Mn <strong>of</strong> PLA increased<br />

from 3.3×103 to 6.0×103 <strong>and</strong> the PDI broadened from 1.1 to 1.5 as the monomer/initiator<br />

ratio increased from 100/1 to 200/1 (Table 3.1, Entries 5 <strong>and</strong> 6). The Co(III)-b-NO2 initiating<br />

system without i PrOH also catalyzed the polymerization <strong>of</strong> LacOCA. Compared with the<br />

Co(III)-b-NO2/ i PrOH initiating system, the Mn <strong>of</strong> the resulting PLA using the Co(III)-b-NO2<br />

initiating system was slightly higher (Mn = 7.6×10 3 ). The molecular weight distribution <strong>of</strong><br />

PLA was also broad (PDI = 1.4), which was similar to that <strong>of</strong> the Co(III)-b-NO2/ i PrOH<br />

initiating system. These demonstrated that the Co(III) complexes with Schiff base lig<strong>and</strong>s<br />

acted as the catalysts for the ROP <strong>of</strong> LacOCA. The present ROP method somehow allowed<br />

the control <strong>of</strong> the molecular weight <strong>of</strong> PLA, but the molecular weight distribution was broad<br />

in most cases <strong>of</strong> the polymerization examined in this study, especially in longer<br />

polymerization time. The large polydispersity implies that the ROP may have been<br />

accompanied by undesired intra- <strong>and</strong>/or inter-transesterification reactions.<br />

Table 3.1 Polymerizations <strong>of</strong> LacOCA<br />

Entry Initiating system [M]/[I] Temp.<br />

‐ 78 ‐<br />

/°C<br />

Time<br />

1 Co(III)-a-(NO2)2/ i PrOH 100/1 70 35 4.0 1.3<br />

2 Co(III)-b-(NO2)2/ i PrOH 100/1 70 35 4.3 1.2<br />

3 Co(III)-c-(NO2)2/ i PrOH 100/1 70 35 9.6 1.4<br />

4 Co(III)-a-NO2/ i PrOH 100/1 70 35 6.4 1.4<br />

5 Co(III)-b-NO2/ i PrOH 100/1 70 35 3.3 1.1<br />

6 Co(III)-b-NO2/ i PrOH 200/1 70 60 6.0 1.5<br />

/hr<br />

Mn<br />

/10 3<br />

PDI


Polymerization <strong>of</strong> Lactic O‐Carboxylic Anhydride using Organometallic Catalysts<br />

7 Co(III)-b-NO2 200/1 70 60 7.6 1.4<br />

8 Co(III)-c-NO2/ i PrOH 100/1 70 35 3.6 1.2<br />

9 Stannous benzoate/ i PrOH 120/1 70 15 6.4 1.2<br />

10 Stannous benzoate 120/1 70 15 7.0 1.3<br />

11 Stannous octoate/ i PrOH 150/1 60 15 7.9 1.4<br />

12 (Cyclohexylsalen)AlEt/BnOH 150/1 70 11 N/A N/A<br />

13 (Dimethylsalen)AlEt/BnOH 150/1 70 11 N/A N/A<br />

i PrOH: isopropyl alcohol. BnOH: benzyl alcohol.<br />

3.3 Polymerization <strong>of</strong> LacOCA using Tin(II) or Al(III) Based Complexes<br />

Tin(II) <strong>and</strong> Al(III) complexes are also widely used as catalysts for the ROP <strong>of</strong> lactides.<br />

Based on the activity <strong>of</strong> the Co(III) complexes (vide supra), we decided to examine the ROP<br />

<strong>of</strong> LacOCA using these metal complexes with the similar (or identical) lig<strong>and</strong>s as the<br />

catalysts. Stannous benzoate, stannous octate, (cyclohexylsalen)AlEt, <strong>and</strong><br />

(dimethylpropylsalen)AlEt (Figure 3.3) were employed in this chapter. Both stannous<br />

benzoate <strong>and</strong> stannous octate were efficient catalysts for the ROP <strong>of</strong> LacOCA. When the<br />

stannous benzoate/ i PrOH initiating system was used as the catalyst <strong>and</strong> the monomer/initiator<br />

ratio was 120/1, the Mn <strong>and</strong> PDI <strong>of</strong> the obtained PLA were 6.4×10 3 <strong>and</strong> 1.2, respectively<br />

(Table 3.1, Entry 9). When the stannous benzoate initiating system was applied, the resulting<br />

Mn <strong>of</strong> 7.0×10 3 was slightly higher than that obtained with the benzoate/ i PrOH initiating<br />

system (Table 3.1, Entry 9). This result suggested that the addition <strong>of</strong> i PrOH was not effective<br />

to control the Mn <strong>of</strong> the product <strong>of</strong> the LacOCA polymerization when stannous benzoate was<br />

used as the catalyst, which was in contrast to the DMAP/alcohol mediated ROP <strong>of</strong> LacOCA<br />

10, 11<br />

where the alcohol was able to control the Mn <strong>of</strong> PLA.<br />

(Cyclohexylsalen)AlEt 29, 30 <strong>and</strong> (dimethylpropylsalen)AlEt 28 (Figure 3.3) are efficient Al<br />

based catalysts for the ROP <strong>of</strong> lactides. However, no polymer was isolated from the ROP <strong>of</strong><br />

LacOCA using those catalysts (Table 3.1, Entries 12 <strong>and</strong> 13), which indicated that the<br />

‐ 79 ‐


Chapter 3<br />

aluminum based complexes were totally inactive for the ROP <strong>of</strong> LacOCA.<br />

3.4 <strong>Characterization</strong> <strong>of</strong> the Obtained PLA<br />

Figure 3.3 Structures <strong>of</strong> aluminum based complexes.<br />

The 1 H NMR spectrum <strong>of</strong> the obtained PLA was shown in Figure 3.4. The doublet peaks<br />

at 1.73 ppm was assigned to CH3 resonance <strong>of</strong> the polymer chain. The quartet peaks at<br />

5.15 ppm was attributed to that <strong>of</strong> the CHCH3 <strong>of</strong> PLA. 13 C NMR spectrum <strong>of</strong> the PLA was<br />

displayed in Figure 3.5. The signals at 169.5, 69.0 <strong>and</strong> 16.6 ppm corresponded to the carbon<br />

resonances <strong>of</strong> carbonyl, methine, <strong>and</strong> methyl groups <strong>of</strong> the PLA chain, respectively. These<br />

results were consistent with the reported NMR spectra <strong>of</strong> PLA, 28 which demonstrated that<br />

PLA was formed by the polymerization <strong>of</strong> LacOCA. The carbonyldioxy group was<br />

eliminated from the polymerization system by the liberation <strong>of</strong> carbon dioxide.<br />

8 6 4 2 0<br />

ppm<br />

Figure 3.4 1 H NMR spectrum <strong>of</strong> poly(lactic acid) from LacOCA.<br />

The DSC thermogram <strong>of</strong> the obtained PLA was shown in Figure 3.6. The glass transition<br />

‐ 80 ‐


Polymerization <strong>of</strong> Lactic O‐Carboxylic Anhydride using Organometallic Catalysts<br />

temperature (Tg) <strong>of</strong> the corresponding PLA was 55 °C. This was in consistent with the<br />

reported data in a literature (Tg ~ 55 °C). 31 The melting point Tm <strong>of</strong> the PLA was 154 °C,<br />

which was lower than that for the poly(L-lactide) reported in the literature (Tm ~ 175 °C). 31<br />

TGA revealed that T5%, the temperature corresponding to 5% weight loss, was 231°C for the<br />

PLA produced by the LacOCA polymerization. This was lower than that <strong>of</strong> the previously<br />

reported poly(L-lactide) (T5% ~ 320 °C). 32 These lower characteristic thermal transition<br />

temperatures were ascribed to the relatively low molecular weights <strong>of</strong> the PLA produced by<br />

the LacOCA polymerization. Increasing the molecular weights by suppressing the proposed<br />

side reactions such as the transesterification by tuning the catalytic activity with a suitably<br />

designed lig<strong>and</strong> around the cobalt center will be the topics <strong>of</strong> our continuous research.<br />

200 150 100 50 0<br />

ppm<br />

Figure 3.5 13 C NMR spectrum <strong>of</strong> poly(lactic acid) from LacOCA.<br />

Heat flow (W/g)<br />

0.5<br />

0.0<br />

-0.5<br />

-1.0<br />

-1.5<br />

-2.0<br />

50 100 150<br />

Temperature ( o C)<br />

Figure 3.6. DSC thermogram <strong>of</strong> poly(lactic acid) (Table 3.1, Entry 3).<br />

‐ 81 ‐


Chapter 3<br />

3.5 Conclusions<br />

Co(III) complexes with Schiff base lig<strong>and</strong>s <strong>and</strong> Tin(II) alphatates were found to be the<br />

catalysts for the ROP <strong>of</strong> LacOCA to produce poly(lactic acid). Intra- <strong>and</strong>/or<br />

inter-transesterification were suggested to coincide with the polymerization, which resulted in<br />

a relatively large polydispersity. The carbonyldioxy group was eliminated from the<br />

polymerization system. The corresponding PLA showed similar Tg but lower T5% compared<br />

with PLA obtained from lactides, as a result <strong>of</strong> the lower molecular weights. Al(III)<br />

complexes with Schiff base lig<strong>and</strong>s were not catalytically active for the polymerization <strong>of</strong><br />

LacOCA.<br />

3.6 Experimental Section<br />

General Procedures<br />

LacOCA <strong>and</strong> Co(III) complexes with Schiff base lig<strong>and</strong>s (Figure 3.2) were prepared<br />

according to the already published procedures. 33, 34 Toluene was dried by distillation over<br />

sodium under the protection <strong>of</strong> nitrogen. 1 H NMR spectra were recorded on Bruker AV 300<br />

MHz, Bruker AV 400 MHz or Bruker AV 600 MHz spectrometer in CDCl3 at 25ºC. Chemical<br />

shifts were referenced to tetramethylsilane (TMS) for the 1 H NMR measurement. Gel<br />

permeation chromatography (GPC) was performed in THF (Flow rate 1.00 ml/min at 35ºC)<br />

on a Waters 410 GPC. The columns were calibrated against polystyrene (PS) as the external<br />

st<strong>and</strong>ard. Differential scanning calorimetry (DSC) analyses were carried out at a heating rate<br />

<strong>of</strong> 10 °C/min on a Perkin Elmer Instruments DSC-7.<br />

Polymerization <strong>of</strong> LacOCA<br />

All <strong>of</strong> the procedures were carried out under the protection <strong>of</strong> dry argon. In a typical<br />

procedure, LacOCA (2.92 mmol, 0.34 g), 2-propanol (0.029 mmol, in 0.34 mL <strong>of</strong> toluene),<br />

Co(III)-c-(NO2)2 (0.029 mmol, 0.0258g), <strong>and</strong> toluene (5.5 mL) were added to a dried<br />

ampoule equipped with a magnetic stirrer bar. The initial concentration <strong>of</strong> LacOCA monomer<br />

in toluene was 0.5 mol/mL. The ampoule was placed in an oil bath at 70 °C. After<br />

polymerization, the polymer was dissolved in chlor<strong>of</strong>orm <strong>and</strong> isolated by precipitation into<br />

cold ethanol, which was collected by filtration <strong>and</strong> dried under vacuum at room temperature<br />

for 24h.<br />

‐ 82 ‐


References<br />

Polymerization <strong>of</strong> Lactic O‐Carboxylic Anhydride using Organometallic Catalysts<br />

1. H. R. Kricheldorf, M. Berl, N. Scharnagl, Macromolecules 1988, 21, 286.<br />

2. C. G. Pitt, M. M. Gratzl, G. L. Kimmel, J. Surles, A. Schindler, Biomaterials 1981, 2,<br />

215.<br />

3. K. J. L. Burg, S. Porter, J. F. Kellam, Biomaterials 2000, 21, 2347.<br />

4. J. P. Penning, H. Dijkstra, A. J. Pennings, Polymer 1993, 34, 942.<br />

5. R. Bodmeier, K. H. Oh, Chen, H. Int. J. Pharm. 1989, 51, 1.<br />

6. H. Ohara, U. Doi, M. Otsuka, H. Okuyama, S. Okada, Seibutsu-Kogaku Kaishi-J. Soc.<br />

Ferment. Bioeng. 2001, 79, 142.<br />

7. J. Zhang, P. Krishnamachari, J. Z. Lou, A. Shahbazi, In Proceedings <strong>of</strong> the 2007 National<br />

Conference on Environmental Science <strong>and</strong> Technology, Uzochukwu, G. A., Ed. Springer:<br />

New York, 2009; pp 3-8.<br />

8. G. Schwach, J. Coudane, R. Engel, M. Vert, Biomaterials 2002, 23, 993.<br />

9. L. Tang, L. Deng, J. Am. Chem. Soc. 2002, 124, 2870.<br />

10. O. T. Du Boullay, E. Marchal, B. Martin-Vaca, F. P. Cossio, D. Bourissou, J. Am. Chem.<br />

Soc. 2006, 128, 16442.<br />

11. C. Bonduelle, B. Martin-Vaca, F. P. Cossio, D. Bourissou, Chem. Eur. J. 2008, 14, 5304.<br />

12. C. Bonduelle, B. Martin-Vaca, D. Bourissou, Biomacromolecules 2009, 10, 3069.<br />

13. X. Pang, H. Z. Du, X. S. Chen, X. H. Wang, X. B. Jing, Chem. Eur. J. 2008, 14, 3126.<br />

14. H. Z. Du, X. Pang, H. Y. Yu, X. L. Zhuang, X. S. Chen, D. M. Cui, X. H. Wang, X. B.<br />

Jing, Macromolecules 2007, 40, 1904.<br />

15. Z. H. Tang, V. C. Gibson, Eur. Polym. J. 2007, 43, 150.<br />

16. X. Pang, H. Z. Du, X. S. Chen, X. L. Zhuang, D. M. Cui, X. B. Jing, J. Polym. Sci., Part<br />

A: Polym. Chem. 2005, 43, 6605.<br />

17. Z. H. Tang, Y. K. Yang, X. Pang, J. L. Hu, X. S. Chen, N. H. Hu, X. B. Jing, J. Appl.<br />

Polym. Sci. 2005, 98, 102.<br />

18. A. Kowalski, A. Duda, S. Penczek, Macromolecules 1998, 31, 2114.<br />

19. B. M. Chamberlain, M. Cheng, D. R. Moore, T. M. Ovitt, E. B. Lobkovsky, G. W. Coates,<br />

J. Am. Chem. Soc. 2001, 123, 3229.<br />

20. L. H. Piao, Z. L. Dai, M. X. Deng, X. S. Chen, X. B. Jing, Polymer 2003, 44, 2025.<br />

21. Z. Y. Zhong, P. J. Dijkstra, C. Birg, M. Westerhausen, J. Feijen, Macromolecules 2001,<br />

34, 3863.<br />

‐ 83 ‐


Chapter 3<br />

22. Z. H. Tang, X. S. Chen, Q. Z. Hang, X. C. Bian, L. X. Yang, L. H. Piao, X. B. Jing, J.<br />

Polym. Sci., Part A: Polym. Chem. 2003, 41, 1934.<br />

23. T. Uchida, T. Katsuki, K. Ito, S. Akashi, A. Ishii, T. Kuroda, Helv. Chim. Acta 2002, 85,<br />

3078.<br />

24. S. D. Choi, G. J. Kim, Catal. Lett. 2004, 92, 35.<br />

25. T. Satoh, T. Imai, S. Umeda, K. Tsuda, H. Hashimoto, T. Kakuchi, Carbohydr. Res. 2005,<br />

340, 2677.<br />

26. W. M. Ren, Z. W. Liu, Y. Q. Wen, R. Zhang, X. B. Lu, J. Am. Chem. Soc. 2009, 131,<br />

11509.<br />

27. Z. H. Tang, X. S. Chen, Y. K. Yang, X. Pang, J. R. Sun, X. F. Zhang, X. B. Jing, J. Polym.<br />

Sci., Part A: Polym. Chem. 2004, 42, 5974.<br />

28. Z. H. Tang, X. S. Chen, X. Pang, Y. K. Yang, X. F. Zhang, X. B. Jing,<br />

Biomacromolecules 2004, 5, 965.<br />

29. Z. Y. Zhong, P. J. Dijkstra, J. Feijen, J. Am. Chem. Soc. 2003, 125, 11291.<br />

30. Z. Y. Zhong, P. J. Dijkstra, J. Feijen, Angew. Chem.Int. Edit. 2002, 41, 4510.<br />

31. D. Garlotta, J. Polym. Environ. 2001, 9, 63.<br />

32. G. X. Chen, J. S. Yoon, Polym. Degrad. Stabil. 2005, 88, 206.<br />

33. Y. S. Niu, W. X. Zhang, H. C. Li, X. S. Chen, J. R. Sun, X. L. Zhuang, X. B. Jing,<br />

Polymer 2009, 50, 441.<br />

34. Y. S. Niu, W. X. Zhang, X. Pang, X. S. Chen, X. L. Zhuang, X. B. Jing, J. Polym. Sci.,<br />

Part A: Polym. Chem. 2007, 45, 5050.<br />

‐ 84 ‐


Chapter 4<br />

<strong>Synthesis</strong> <strong>of</strong> Lactide-Grafted Poly(TEMPO-acrylamide) <strong>and</strong> its<br />

Electrochemical Property <strong>and</strong> Biocompatibility<br />

4.1 Introduction<br />

4.2 <strong>Synthesis</strong> <strong>of</strong> PTAm-g-PLA<br />

4.3 Electrochemical Properties<br />

4.4 Thermal Properties<br />

4.5 Phase Separation Behavior <strong>of</strong> PTAm-g-PLA<br />

4.6 Cytotoxicity <strong>of</strong> PTAm-g-PLA<br />

4.7 Cell Adhesion <strong>and</strong> Spreading<br />

4.8 Conclusions<br />

4.9 Experimental Section<br />

References


Chapter 4<br />

4.1 Introduction<br />

Stimulus-responsive polymers show great potential application as new “smart”<br />

biomaterials. 1 The electroactive polymer, being electrical stimulus responsive, has attracted<br />

increasing attentions since it has been established that variety <strong>of</strong> cells, including many <strong>of</strong><br />

fibroblasts, nerve cells <strong>and</strong> osteoblasts are sensitive to the electrical effects generated from<br />

the electrodes in vitro or in vivo. 2, 3 Besides, they had been widely used in biological systems<br />

due to the key advantage <strong>of</strong> facile control <strong>of</strong> the intensity <strong>and</strong> duration <strong>of</strong> stimulation. So far,<br />

polypyrrole (PPy) <strong>and</strong> polyaniline (PANi) are still the most widely used two kinds <strong>of</strong><br />

electroactive materials in tissue engineering although the researches on them have lasted a<br />

long time. 4-7 More recently, there are some new electroactive polymers being developed, such<br />

as polythiophene (PT) <strong>and</strong> other types <strong>of</strong> conductive polymers. 8 However, there are still few<br />

examples attempting to explore the application <strong>of</strong> new electroactive materials in tissue<br />

engineering.<br />

In recent years, polymers with stable free radicals, known as radical polymers, have<br />

attracted more <strong>and</strong> more attention due to their potential applications as electron transport<br />

materials <strong>and</strong> organic ferromagnets. 9, 10 Radical polymers are readily oxidized <strong>and</strong> reduced<br />

electrochemically, which allowed the use in charge-storage applications for new-types <strong>of</strong><br />

batteries. 11, 12 However, little attention has been paid to investigate the application <strong>of</strong><br />

nitroxide radical-containing polymers in the field <strong>of</strong> biomedical materials. Chu et al.<br />

synthesized biodegradable nitroxide radical-containing polymers through the condensation <strong>of</strong><br />

4-amino-2,2,6,6-tetramethylpiperidine-1-oxy <strong>and</strong> polyglycolide. This material showed<br />

pr<strong>of</strong>ound retardation <strong>of</strong> the proliferation <strong>of</strong> smooth muscle cell in vitro. 13, 14 However, the<br />

hydrolytic degradation <strong>of</strong> PGA will lead to the release <strong>of</strong> nitroxide radicals which may have<br />

adverse side effects to the body. Using radical polymers as the source <strong>of</strong> stable radical is an<br />

alternative way to solve this problem. These polymers usually have non-biodegradable<br />

backbones which will avoid the release <strong>of</strong> radical agents after the degradation <strong>of</strong> the<br />

main-chain. PTAm <strong>and</strong> its analogs are typical radical polymers <strong>and</strong> mostly examined as the<br />

cathode materials. 15, 16 To the best <strong>of</strong> our knowledge, there has few work to investigate the<br />

biomedical application <strong>of</strong> the radical polymers such as PTAm. In this chapter, we<br />

copolymerized it with a biodegradable polylactide (PLA) macromonomer, to obtain a novel<br />

biologically active <strong>and</strong> biocompatible copolymer. PLA are well known as very important<br />

synthetic biodegradable materials. Due to its low immunogenicity, good biocompatibility <strong>and</strong><br />

‐ 86 ‐


‐ 87 ‐<br />

<strong>Synthesis</strong> <strong>of</strong> Lactide‐Grafted Poly(TEMPO‐acrylamide) <strong>and</strong><br />

its Electrochemical Property <strong>and</strong> Biocompatibility<br />

excellent mechanical properties, they are widely used in pharmaceutical <strong>and</strong> other medical<br />

applications, such as sutures, implants for bone fixation, carriers in drug delivery, <strong>and</strong><br />

temporary matrices or scaffolds in tissue engineering. 17-19 The incorporation <strong>of</strong> PLA fragment<br />

into PTAm will improve the biocompatibility <strong>of</strong> PTAm, <strong>and</strong> the resulting biomaterials will<br />

have promising application in tissue engineering.<br />

4.2 <strong>Synthesis</strong> <strong>of</strong> PTAm-g-PLA<br />

The synthetic route <strong>of</strong> PTAP-g-PLA was shown in Scheme 4.1. Firstly, double bond ended<br />

PLA was synthesized by the ring-opening polymerization <strong>of</strong> lactide initiated by HEMA. Then<br />

the macromonomer was copolymerized with 2,2,6,6-tetrametylpiperidine-4-yl acrylamide<br />

prepared by the reaction <strong>of</strong> 4-amino-2,2,6,6-tetrametylpiperidine with acryl chloride<br />

according to our previous method. 20 The resulted copolymer was oxidized by the peroxide to<br />

give PTAm-g-PLA.<br />

Scheme 4.1 Synthetic routes <strong>of</strong> PTAm-g-PLA.<br />

As shown in Figure 4.1A, the peaks around 6.13 <strong>and</strong> 5.58 ppm were assigned to the<br />

characteristic peaks <strong>of</strong> the double bond <strong>of</strong> HEMA, <strong>and</strong> the peak at 5.19 ppm was attributed to


Chapter 4<br />

the proton <strong>of</strong> methenyl groups in the lactide backbone. The molecular weight <strong>of</strong> PLA-HEMA<br />

was 3000, calculated from the integration area ratio <strong>of</strong> peaks 1 <strong>and</strong> 6 (Figure 4.1A). To<br />

prevent the possible polymerization <strong>of</strong> HEMA under the high temperature, hydroquinone was<br />

used as an inhibitor which was removed during the purification step <strong>of</strong> the polymeric product.<br />

After the copolymerization with the TAP monomer, the peak associated with the double bond<br />

disappeared, as shown in Figure 4.1B, which suggested the polymerization <strong>of</strong> PLA-HEMA.<br />

The structure <strong>of</strong> copolymer was also confirmed by 13 C NMR (Figure 4.2).<br />

Figure 4.1 1 H NMR <strong>of</strong> PLA-HEMA (A) <strong>and</strong> PTAP-g-PLA2 (B).<br />

Figure 4.2 13 C NMR spectra <strong>of</strong> PTAP (A), PTAP-g-PLA (B) <strong>and</strong> PLA-HEMA (C).<br />

The contents <strong>of</strong> PTAP in the copolymers were determined by the integration area ratio <strong>of</strong><br />

the peaks at 5.19 <strong>and</strong> 4.19 ppm. It was noted that the peak at 4.19 ppm contained the<br />

resonance from the –CH2CH2- groups capped at the PLA chain end, so this part <strong>of</strong> integration<br />

‐ 88 ‐


‐ 89 ‐<br />

<strong>Synthesis</strong> <strong>of</strong> Lactide‐Grafted Poly(TEMPO‐acrylamide) <strong>and</strong><br />

its Electrochemical Property <strong>and</strong> Biocompatibility<br />

was subtracted to determine the monomer composition (Figure 4.1B). The compositions <strong>of</strong><br />

the copolymers evaluated by the 1 H NMR spectra were summarized in Table 4.1. All the<br />

copolymer had a relatively lower PLA content compared with the feeding ratio. Presumably<br />

the PLA-HEMA macromonomers had a relatively low reactive efficiency <strong>and</strong> only part <strong>of</strong> the<br />

macromonomers copolymerized with the 2,2,6,6-tetrametylpiperidine-4-yl acrylamide<br />

monomer. The low yield <strong>of</strong> the copolymerization can also be attributed to the low reactivity<br />

<strong>of</strong> the macromonomers. The GPC results showed that the molecular weight <strong>of</strong> the copolymers<br />

decreased as the PLA content increased, indicating that the introduction <strong>of</strong> the PLA lowered<br />

the reactivity <strong>of</strong> the propagating copolymer radical, making them more readily terminated<br />

during the polymerization.<br />

Sample<br />

Table 4.1 Feed <strong>and</strong> result compositions <strong>of</strong> copolymers. a<br />

PLA-HEMA/Monomer<br />

(w/w)<br />

Feed Result b<br />

Yield<br />

(%)<br />

Mn c PDI c<br />

Unpaired<br />

electron<br />

density<br />

(spin/g) d<br />

1 3 : 7 1 : 9.0 42.0 7.0 × 10 4 1.98 2.14 × 10 21<br />

2 5 : 5 1 : 4.3 32.5 3.9 × 10 4 2.03 1.64 × 10 21<br />

3 7 : 3 1 : 2.4 40.8 1.2 × 10 4 1.97 1.39 × 10 21<br />

a Reaction temperature was 85 o C <strong>and</strong> duration was 24 h;. b Calculated from 1 H NMR; c Determined from GPC in DMF. d<br />

Determined from SQUID after oxidation.<br />

TEMPO are generally prepared through the oxidation <strong>of</strong> 2,2,6,6-tetramethylpiperidine.<br />

Several oxidants, including hydrogen peroxide, peracids <strong>and</strong> lead(IV) oxide, can be used to<br />

obtain the nitrooxide radical. In this chapter, the oxidation <strong>of</strong> PTAP-g-PLA was accomplished<br />

in the presence <strong>of</strong> 3-chloroperoxybenzoic acid. The obtained PTAm-g-PLA was orange<br />

powder.<br />

To further investigate the oxidation reaction <strong>of</strong> copolymer, the FTIR spectra before <strong>and</strong><br />

after the oxidation were obtained (Figure 4.3). The peak at 1359 cm -1 was obviously<br />

enhanced after the oxidation, which was assigned to the characteristic peak <strong>of</strong> the nitroxide<br />

radical. 21, 22 The peak at 1758 cm -1 (νCO) was characteristic peak <strong>of</strong> the PLA segment, which<br />

did not change after the oxidation. The peaks at 1653 cm -1 (νCO) <strong>and</strong> 1548 cm -1 (νCO-NH) were


Chapter 4<br />

attributed to the amide group. The GPC elution curve for PTAm-g-PLA gave a unimodal peak,<br />

which indicated that there was no undesired homopolymer residue in the final product. Based<br />

on these results, it was concluded that the PTAm-g-PLA copolymers were successfully<br />

prepared by this method.<br />

Figure 4.3 FTIR spectra <strong>of</strong> PTAP-g-PLA2 (A) <strong>and</strong> PTAm-g-PLA2 (B).<br />

4.3 Electrochemical Properties<br />

Based on the presence <strong>of</strong> the nitroxide radicals, the copolymers exhibited<br />

electrochemically reversible properties similar to that <strong>of</strong> TEMPO. The cyclic voltammogram<br />

(CV) <strong>of</strong> the copolymer at different scan rates were shown in Figure 4.4. The CV curves for<br />

the copolymer were recorded in the potential range <strong>of</strong> 0.00 to 1.30 V. The response revealed a<br />

highly reversible redox property at around 0.63 V vs. Ag/AgCl, which was assigned to the<br />

oxidation <strong>of</strong> the nitroxide to the corresponding oxoammonium cation. The redox potential<br />

(0.63 V vs. Ag/AgCl) was in good agreement with that <strong>of</strong> the PTAm homopolymer (0.68 V vs.<br />

Ag/AgCl). Moreover, the peak current is proportional to the sweep rate, suggesting a fast<br />

electron transfer process in the copolymer film.<br />

‐ 90 ‐


‐ 91 ‐<br />

<strong>Synthesis</strong> <strong>of</strong> Lactide‐Grafted Poly(TEMPO‐acrylamide) <strong>and</strong><br />

its Electrochemical Property <strong>and</strong> Biocompatibility<br />

Figure 4.4 CV <strong>of</strong> the copolymer (PTAm-g-PLA2) at different scan rates in an aqueous<br />

electrolyte solution <strong>of</strong> 0.1 M NaBF4.<br />

The spin concentration <strong>of</strong> the copolymer was determined by magnetic measurement using<br />

a SQUID technique. As determined from the Curie plots <strong>and</strong> the values for saturated<br />

magnetization (Figure 4.5), the radical concentration <strong>of</strong> PTAm-g-PLA2 was 1.64×10 21 spin/g.<br />

As calculated from the ratio <strong>of</strong> the radical amount determined by SQUID to the precursors<br />

derived from the 1 H NMR spectra <strong>of</strong> PTAP-g-PLA, it can be found that more than 80% <strong>of</strong> the<br />

TAP units was oxidized. Considering that almost all <strong>of</strong> the piperidine units were oxidized in<br />

the PTAm homopolymer under the same experimental conditions, the PLA segment probably<br />

inhibited the oxidation <strong>of</strong> the piperidine unit to a small extent.<br />

Figure 4.5 Curie <strong>and</strong> χT (inset) vs. T plots for the radical copolymers obtained by SQUID measurements.


Chapter 4<br />

4.4 Thermal Properties<br />

The DSC thermograms <strong>of</strong> the PTAm homopolymer, PLA-HEMA <strong>and</strong> PTAm-g-PLA2 were<br />

recorded <strong>and</strong> shown in Figure 4.6A. PTAm showed no thermal transition at temperatures<br />

above 20 o C. The melting temperature Tm <strong>of</strong> PLA-HEMA was low, around 133 o C, due to its<br />

low molecular weight. For PTAm-g-PLA copolymer, no Tg or Tm was observed in the tested<br />

temperature ranges, which may be ascribed to the low PLA content in the copolymer. The<br />

TGA curves <strong>of</strong> the PTAm homopolymer, PLA-HEMA <strong>and</strong> PTAm-g-PLA also demonstrated<br />

that the copolymers had similar thermal properties with that <strong>of</strong> PTAm (Figure 4.6B).<br />

Figure 4.6 (A) DSC traces <strong>of</strong> the PLA-HEMA, PTAm, <strong>and</strong> PTAm-g-PLA2 in second heating. (B)<br />

TGA curves <strong>of</strong> PLA-HEMA, PTAm, <strong>and</strong> PTAm-g-PLA.<br />

4.5 Phase Separation Behavior <strong>of</strong> PTAm-g-PLA<br />

The phase behavior <strong>of</strong> the PTAm-g-PLA copolymer is very important to its application.<br />

Therefore, the phase images <strong>of</strong> blend <strong>of</strong> PTAm <strong>and</strong> PLA, blend <strong>of</strong> PLA <strong>and</strong> PTAm-g-PLA2,<br />

PTAm-g-PLA2 <strong>and</strong> PTAm-g-PLA1 films cast from their chlor<strong>of</strong>orm solutions were examined<br />

by AFM (Figure 4.7). It was observed in Figure 4.7B <strong>and</strong> 4.7C that the film <strong>of</strong> copolymer<br />

exhibited two different phases. One was continuous <strong>and</strong> another phase was uniformly<br />

dispersed in the film. Considering the low PLA content in PTAm-g-PLA2 <strong>and</strong> PTAm-g-PLA1,<br />

it can be safely concluded that the continuous phase was PTAm. This unique phase behavior<br />

can probably be attributed to the immiscible nature <strong>of</strong> PTAm <strong>and</strong> PLA as shown in Figure<br />

4.7A. It had been demonstrated that the incorporation <strong>of</strong> two immiscible chains into one<br />

copolymer prevented the macro-phase separation <strong>of</strong> these incompatible materials, which can<br />

restrict the partition <strong>of</strong> the components to nanoscopic (1-100 nm) domains. 23 As shown in<br />

‐ 92 ‐


‐ 93 ‐<br />

<strong>Synthesis</strong> <strong>of</strong> Lactide‐Grafted Poly(TEMPO‐acrylamide) <strong>and</strong><br />

its Electrochemical Property <strong>and</strong> Biocompatibility<br />

Figure 4.7D, the blend film <strong>of</strong> PLA <strong>and</strong> PTAm-g-PLA2 also exhibited phase separation but<br />

the dispersion <strong>of</strong> PLA was more uniform, indicating that PTAm-g-PLA acted as a<br />

compatibilizer which modified the PLA <strong>and</strong> PTAm interface.<br />

4.6 Cytotoxicity <strong>of</strong> PTAm-g-PLA<br />

To determine the cytotoxicity <strong>of</strong> the PTAm-g-PLA copolymer in comparison with PLA <strong>and</strong><br />

PTAm, the MTT assay by the RSC96 cell line was conducted, <strong>and</strong> the results were shown in<br />

Figure 4.8. The copolymers exhibited similar cytocompatibility as compared to the PTAm<br />

homopolymer. This is mainly attributed to the reason that all the polymers tested are<br />

hydrophobic <strong>and</strong> no cytotoxic small organic molecules or polymers were dissolved during the<br />

extraction process.<br />

Figure 4.7 AFM Phase image: blend <strong>of</strong> PLA <strong>and</strong> PTAm with a weight ratio <strong>of</strong> 1:4 (A);<br />

PTAm-g-PLA2 (B), PTAm-g-PLA1(C), <strong>and</strong> blend <strong>of</strong> PLA <strong>and</strong> PTAm-g-PLA2 with a weight ratio<br />

<strong>of</strong> 1:4 (D).


Chapter 4<br />

Figure 4.8 MTT assay for the cytotoxicity <strong>of</strong> PLA-HEMA, PTAm <strong>and</strong> PTAm-g-PLA. Statistical<br />

significance: *p < 0.05<br />

4.7 Cell Adhesion <strong>and</strong> Spreading<br />

Biocompatibility is also a great concern for medical applications. The biocompatibility <strong>of</strong><br />

PTAm-g-PLA copolymers was evaluated by observing adhesion <strong>and</strong> spreading <strong>of</strong> RSC96<br />

cells on copolymer films. Figure 4.9 showed the micrographs <strong>of</strong> RSC96 cells incubated for<br />

48 h on PLA, PTAm-g-PLA2 <strong>and</strong> PTAm films, respectively. After 48 h, almost all cells<br />

adhered to the PLA <strong>and</strong> PTAm-g-PLA2 films <strong>and</strong> got spread. The cells extended from<br />

original elliptical shape to the irregular polygonal shape. Moreover, the filopodias appeared<br />

on the surface <strong>of</strong> cells <strong>and</strong> networks were formed among cells. All the above observation<br />

demonstrated that the adhesion <strong>and</strong> spreading <strong>of</strong> RSC96 cells on the PTAm-g-PLA2 were<br />

good. On the pure PTAm film, the cells also adhered well but their shapes were still original<br />

elliptical shape, which showed the cells did not spread on the film. The cells are separated<br />

<strong>and</strong> few filopodias or networks were observed. After copolymerized with PLA, the<br />

copolymers had improved biocompatibility compared with PTAm, which were comparable<br />

with PLA, indicating great potential use in tissue engineering.<br />

‐ 94 ‐


‐ 95 ‐<br />

<strong>Synthesis</strong> <strong>of</strong> Lactide‐Grafted Poly(TEMPO‐acrylamide) <strong>and</strong><br />

its Electrochemical Property <strong>and</strong> Biocompatibility<br />

Figure 4.9 Cell morphology <strong>of</strong> F-actin distribution in RSC96 cells incubated on cover-slip coated<br />

with PLA (A), PTAm-g-PLA (B) <strong>and</strong> PTAm (C). RSC96 cells were incubated for 48 h <strong>and</strong> stained<br />

for F-actin. Representative fluorescence microscopy photographs were shown here.<br />

4.8 Conclusions<br />

Novel PTAm-g-PLA copolymers were successfully prepared by a PLA macromonomer<br />

approach. The obtained copolymer contained stable radical units which was biologically<br />

active. The copolymers were amorphous which were determined from DSC measurement due<br />

to the relatively low PLA content. The copolymer exhibited an improved biocompatibility<br />

compared to PTAm homopolymer, which indicated that it was a promising biomaterial for the<br />

tissue engineering.<br />

4.9 Experimental Section<br />

Materials <strong>and</strong> Methods<br />

4-Amino-2,2,6,6-tetramethylpiperidine (>97%, TCI), hydroquinone (sinopharm chemical<br />

reagent, >98%), 3-chloroperoxybenzonic acid (sinopharm chemical reagent, >85%) were<br />

used as received. Acryloyl chloride was purchased from Best Chemical (China) <strong>and</strong> purified<br />

by vacuum distillation. 2-Hydroxyethyl methacrylate (HEMA, 96%, Acros) was distilled<br />

under reduced pressure before use. L-Lactide (LA) was purchased from Purac <strong>and</strong><br />

recrystallized in ethyl acetate three times. 2,2'-Azoisobutyronitrile (AIBN, Beijing Chemical<br />

Co., China) was recrystallized twice from methanol. All the organic solvents used were dried<br />

over CaH2 <strong>and</strong> distilled prior to use.


Chapter 4<br />

The 1 H NMR <strong>and</strong> 13 C NMR measurements were performed on Bruker DMX300<br />

spectrometer with tetramethylsilane (TMS) as an internal reference. FTIR spectra were<br />

measured on a Bruker Vertex 70 Fourier Transform Infrared spectrometer using the KBr disk<br />

method. Atomic force microscopy (AFM) images were acquired in a commercial<br />

SPA300HV/SPI3800N Probe Station, Seiko Instruments, Japan, in tapping mode. A silicon<br />

microcantilever (spring constant 2 N/m <strong>and</strong> resonance frequency ~ 70 kHz, Olympus, Japan)<br />

with an etched conical tip was used for scan. Silicon wafers were cleaned by a 30 min dip in<br />

fresh 1:1 H2SO4 (concentrated)/H2O2 (30%) solution. Subsequently, the acids were removed<br />

by a thorough rinse in Millipore water. Polymer thin films were obtained by spin-coating a<br />

copolymer chlor<strong>of</strong>orm solution onto silicon wafers. Magnetization <strong>and</strong> magnetic<br />

susceptibility <strong>of</strong> the powder polymer sample were measured using a Quantum Design<br />

MPMS-7 SQUID (superconducting quantum interference device) magnetometer. The<br />

magnetic susceptibility was measured from 10 to 300 K under a 1.0 T field. Cyclic<br />

voltammetry was performed using a normal potentiostat system (BAS Inc. ALS660B) with a<br />

conventional three-electrode cell under a dry argon atmosphere. A platinum disk, coiled<br />

platinum wire, <strong>and</strong> Ag/AgCl were used as the working, auxiliary <strong>and</strong> reference electrode,<br />

respectively. The cyclic voltammogram was measured in a dichloromethane solution in the<br />

presence <strong>of</strong> 0.1 M tetrabutylammonium tetrafluoroborate as the supporting electrolyte. The<br />

film <strong>of</strong> copolymer was 100 nm in thickness. The formal potential <strong>of</strong> the<br />

ferrocene/ferrocenium redox couple was 0.45 V vs the Ag/AgCl reference electrode. The<br />

GPC analysis was performed using TOSOH HLC-8220 with 0.1 M LiCl DMF as solvent.<br />

<strong>Synthesis</strong> <strong>of</strong> PLA-HEMA<br />

PLA-HEMA was easily prepared by the ring-opening polymerization <strong>of</strong> L-lactide in the<br />

presence <strong>of</strong> HEMA <strong>and</strong> stannous octoate (Sn(Oct)2). First, 0.48 g HEMA, 5.0 g L-lactide, 8<br />

mg hydroquinone, <strong>and</strong> 5 mg Sn(Oct)2 were added into a dried glass reactor already<br />

flame-dried <strong>and</strong> nitrogen-purged three times. After injection <strong>of</strong> 35 mL toluene, the reactor<br />

was sealed <strong>and</strong> maintained at 120 o C for 24 h. The product was precipitated with an excess <strong>of</strong><br />

ethanol, purified by precipitating twice into ethanol from chlor<strong>of</strong>orm solution, to give a white<br />

product. Yield: 34%.<br />

<strong>Synthesis</strong> <strong>of</strong> PTAP-g-PLA<br />

Typically, 0.15 g PLA-HEMA <strong>and</strong> 0.35 g 2,2,6,6-tetrametylpiperidine-4-yl acrylamide<br />

‐ 96 ‐


‐ 97 ‐<br />

<strong>Synthesis</strong> <strong>of</strong> Lactide‐Grafted Poly(TEMPO‐acrylamide) <strong>and</strong><br />

its Electrochemical Property <strong>and</strong> Biocompatibility<br />

(TAP) were added into a dried glass reactor already flame-dried <strong>and</strong> nitrogen-purged three<br />

times. After injection <strong>of</strong> 6 mL dioxane, 6 mg AIBN was added to the glass in glove box. Then<br />

the system was degassed by three freeze-vacuum-thaw cycles. Finally, the glass reactor was<br />

sealed under Ar atmosphere <strong>and</strong> then immersed in an oil bath at 85 o C under stirring. After 24<br />

h, the reaction solution was poured into excessive methanol. The homopolymer <strong>of</strong><br />

PLA-HEMA was precipitated <strong>and</strong> filtered. The poly(2,2,6,6-tetrametylpiperidine-4-yl<br />

acrylamide)-g-polylactide (PTAP-g-PLA) was dispersed in the methanol solution <strong>and</strong><br />

collected by evaporation <strong>of</strong> methanol under reduced pressure. Then the product was purified<br />

by precipitating twice into diethyl ether from chlor<strong>of</strong>orm solution, <strong>and</strong> eventually dried under<br />

vacuum at room temperature for 24 h.<br />

<strong>Synthesis</strong> <strong>of</strong> PTAm-g-PLA<br />

PTAP-g-PLA was added to an ice-cold solution <strong>of</strong> 3-chloroperoxybenzoic acid in THF <strong>and</strong><br />

stirred for 3 h at room temperature. The weight ratio <strong>of</strong> PTAP to 3-chloroperoxybenzoic acid<br />

was 1:4. The polymer was solved in THF with oxidation progress. The solution was added<br />

dropwise to diethyl ether/hexane (1/1 v/v 200 ml). The precipitate was collected by filtration,<br />

<strong>and</strong> dried under reduced pressure for 12 h. Yield: 77%.<br />

Cytotoxicity assay<br />

The cytotoxicity <strong>of</strong> the various polymer complexes were assessed using the MTT assay.<br />

PTAm-g-PLA, PLA <strong>and</strong> PTAm were dispersed in DMEM medium (Dulbecco's Modified<br />

Eagle Medium) supplemented with 10% fetal calf serum (Gibco) in a humidified incubator at<br />

37 <strong>and</strong> 5% CO2, respectively. After 24 h, extracted solutions <strong>of</strong> polymers were diluted into<br />

different concentrations. RSC96 cells were seeded at 1.2 ×10 4 cells/well in 96-well plates <strong>and</strong><br />

cultured for 24 h. Then the culture media were replaced by the extracted solutions <strong>and</strong> the<br />

plates were returned to the incubator for an additional 24 h. At the end <strong>of</strong> the experiments, 20<br />

μL <strong>of</strong> 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) solution (5<br />

mg/mL) was added to each well. The plates were returned to the incubator. After 4 h, the<br />

MTT solutions were carefully removed from each well, <strong>and</strong> 200 mL DMSO was added to<br />

dissolve the MTT formazan crystals. The plate was incubated for an additional 10 min before<br />

the absorbance at 492nm was recorded by an ELISA microplate reader (Bio-Rad). The cell<br />

viability (%) was calculated according to the following equation:


Chapter 4<br />

Cell viability (%) = (Asample / Acontrol) × 100 (1)<br />

where Asample was the absorbance <strong>of</strong> the polymers extracted solutions treated cells <strong>and</strong> Acontrol<br />

was the absorbance <strong>of</strong> the untreated cells. Each experiment was done in triplicate <strong>and</strong><br />

repeated a minimum <strong>of</strong> three times.<br />

Cell Adhesion <strong>and</strong> Spreading<br />

RSC96 cells (rat Schwann cell) were used to investigate the cell adhesion <strong>and</strong> viability <strong>of</strong><br />

the several materials. The cells were cultured in cell culture flasks with DMEM medium<br />

supplemented with 10% fetal calf serum in a humidified incubator at 37 <strong>and</strong> 5% CO2.<br />

The PLA, PTAm-g-PLA2 <strong>and</strong> PTAm were dissolved in CHCl3 (10 mg/ml), <strong>and</strong> the<br />

solution was cast onto glass slides. The coated slides were treated under vacuum for 48 h to<br />

remove the trace <strong>of</strong> CHCl3, followed by sterilizing by ultraviolet radiation for 2 h.<br />

Approximately, 30,000 RSC96 cells were seeded on the cover-slip coated with PLA,<br />

PTAm-g-PLA, PTAm <strong>and</strong> allowed to grow for 48 hours at 37 o C/5% CO2. RSC96 cells were<br />

routinely grown in DMEM medium containing 10% FBS <strong>and</strong> antibiotics (100 U/mL<br />

penicillin, 100 μg/mL streptomycin). Subsequently, the cells were fixed in 3%<br />

paraformaldehyde in PBS pH 7.4 for 15 min at room temperature, <strong>and</strong> washed twice with ice<br />

cold PBS. In order to improve the penetration <strong>of</strong> the antibody, cells were incubated for 10<br />

min with PBS containing 0.25% Triton X-100. Cells were incubated with 1% BSA in PBST<br />

for 30 min to block unspecific binding <strong>of</strong> the antibodies. Cells were incubated in the diluted<br />

F-actin antibody (Abcam) in 1% BSA in PBST in a humidified chamber for 1 hour at room<br />

temperature. The solution was decanted <strong>and</strong> the cells were washed three times in PBS, 5 min<br />

each time. Cells were incubated with the goat polyclonal to Rabbit IgG conjugated with<br />

Rhodamine secondary antibody (Abcam) in 1% BSA for 1 hour at room temperature in dark.<br />

The secondary antibody solution was decanted <strong>and</strong> washed three times with PBS for 5 min<br />

each in dark. Cells were incubated on 1 μg/ml DAPI (Sigma) for 2 min <strong>and</strong> rinsed with PBS.<br />

The cells were observed with a Nikon Eclipse TE2000-U fluorescence microscope equipped<br />

with a DXM1200F digital camera.<br />

References<br />

1. E. S. Gil, S. A. Hudson, Prog. Polym. Sci. 2004, 29, 1173.<br />

2. E. G. Fine, R. F. Valentini, R. Bellamkonda, P. Aebischer, Biomaterials 1991, 12, 775.<br />

‐ 98 ‐


3. I. Giaever, C. R. Keese, Pro. Natl. Acad. Sci-Biol. 1984, 81, 3761.<br />

‐ 99 ‐<br />

<strong>Synthesis</strong> <strong>of</strong> Lactide‐Grafted Poly(TEMPO‐acrylamide) <strong>and</strong><br />

its Electrochemical Property <strong>and</strong> Biocompatibility<br />

4. J. Y. Wong, R. Langer, D. E. Ingber, Proc. Natl. Acad. Sci. U.S.A 1994, 91, 3201.<br />

5. H. K. Song, B. Toste, K. Ahmann, D. H<strong>of</strong>fman-Kim, G. T. R. Palmore, Biomaterials 2006,<br />

27, 473.<br />

6. M. Y. Li, Y. Guo, Y. Wei, A. G. MacDiarmid, P. I. Lelkes, Biomaterials 2006, 27, 2705.<br />

7. L. H. Huang, X. L. Zhuang, J. Hu, L. Lang, P. B. Zhang, Y. S. Wang, X. S. Chen, Y. Wei, X.<br />

B. Jing, Biomacromolecules 2008, 9, 850.<br />

8. M. Waugaman, B. Sannigrahi, P. McGeady, I. M. Khan, Eur. Polym. J. 2003, 39, 1405.<br />

9. M. Tanaka, S. Imai, T. Tanii, Y. Numao, N. Shimamoto, I. Ohdomari, H. Nishide, J. Polym.<br />

Sci., Part A: Polym. Chem. 2007, 45, 521.<br />

10. H. Oka, H. Kouno, H. Tanaka, J. Mater. Chem. 2007, 17, 1209.<br />

11. K. Nakahara, S. Iwasa, M. Satoh, Y. Morioka, J. Iriyama, M. Suguro, E. Hasegawa, Chem.<br />

Phys. Lett. 2002, 359, 351.<br />

12. H. Nishide, K. Oyaizu, Science 2008, 319, 737.<br />

13. M. D. Lang, C. C. Chu, J. Polym. Sci., Part A: Polym. Chem. 2001, 39, 4214.<br />

14. L. Li, C. C. Chu, J. Biomater. Sci. Polym. Ed. 2009, 20, 341.<br />

15. H. Nishide, S. Iwasa, Y. J. Pu, T. Suga, K. Nakahara, M. Satoh, Electrochim. Acta 2004,<br />

50, 827.<br />

16. J. K. Kim, G. Cheruvally, J. W. Choi, J. H. Ahn, S. H. Lee, D. S. Choi, C. E. Song, Solid<br />

State Ionics 2007, 178, 1546.<br />

17. R. Bhardwaj, A. K. Mohanty, J. Biobased Mater. Bio. 2007, 1, 191.<br />

18. A. Duda, S. Penczek, Polimery 2003, 48, 16.<br />

19. R. Jain, N. H. Shah, A. W. Malick, C. T. Rhodes, Drug. Dev. Ind. Pharm. 1998, 24, 703.<br />

20. K. Koshika, Waseda University, 2009, PhD. Thesis.<br />

21. T. Endo, K. Takuma, T. Takata, C. Hirose, Macromolecules 1993, 26, 3227.<br />

22. T. Kurosaki, K. Wanlee, M. Okawara, J. Polym. Sci., Part A: Polym. Chem. 1972, 10,<br />

3295.<br />

23. M. W. Matsen, F. S. Bates, Macromolecules 1996, 29, 1091.


Chapter 5<br />

<strong>Synthesis</strong> <strong>of</strong> Triblock Copolymers <strong>of</strong> TEMPO-Acrylamide <strong>and</strong><br />

Lactate by RAFT Polymerization<br />

5.1 Introduction<br />

5.2 <strong>Synthesis</strong> <strong>of</strong> Triblock Copolymer<br />

5.3 Thermal Properties<br />

5.4 Cytotoxicity <strong>and</strong> Biocompatibility <strong>of</strong> Copolymers<br />

5.5 Electrospun <strong>of</strong> Copolymers<br />

5.6 Conclusions<br />

5.7 Experimental Section<br />

References


Chapter 5<br />

5.1 Introduction<br />

Radical polymers bearing pendant 2,2,6,6-tetramethylpiperidin-1-oxy (TEMPO) groups<br />

have attracted increasing interest due to their unique characteristics associated with the<br />

unpaired electrons present in free radicals. Up to now, radical polymers have been found<br />

numerous potential applications such as catalyst for redox reaction, 1 one-dimensional<br />

throw-bond organic ferromagnets, 2 probe for in vivo electron paramagnetic resonance (EPR) 3<br />

<strong>and</strong> cathode material for bendable batteries. 4 In Chapter 4 we synthesized graft copolymers<br />

composed <strong>of</strong> poly(2,2,6,6-tetramethylpiperidine-1-oxyl-4-yl acrylamide) (PTAm) <strong>and</strong><br />

polylactide (PLA) <strong>and</strong> preliminarily investigated their application in biomedical field.<br />

However, the graft copolymers did not possessed definite structure due to the r<strong>and</strong>om<br />

location <strong>of</strong> PLA on the PTAm backbone. Therefore, it is necessary to design block<br />

copolymers <strong>of</strong> PLA <strong>and</strong> PTAm with well-defined structure <strong>and</strong> controlled molecular weight.<br />

The block copolymer <strong>of</strong> PTAm <strong>and</strong> PLA should combine the electroactive property <strong>of</strong> radical<br />

polymers <strong>and</strong> biodegradable property <strong>of</strong> PLA together.<br />

In this chapter, well-defined triblock copolymers composed <strong>of</strong> a radical polymer, PTAm,<br />

<strong>and</strong> PLA were synthesized by combination <strong>of</strong> ring-opening polymerization (ROP) <strong>and</strong><br />

reversible addition-fragmentation chain transfer (RAFT) polymerization methods. Compared<br />

to another controlled radical polymerization method, atom transfer radical polymerization<br />

(ATRP), RAFT appears more practical in terms <strong>of</strong> monomer selection <strong>and</strong> reaction<br />

conditions. 5-7 Moreover, RAFT technique requires fewer purification steps after the<br />

polymerization.<br />

The resulted copolymers were then electrospun into mats with varied morphology.<br />

Electro-spinning technique was originally developed to produce ultra-fine polymer fibers, <strong>and</strong><br />

has recently re-emerged as a novel tool for generating micro- or nano-scale biopolymer<br />

scaffolds for tissue engineering. 8-10 The scaffolds fabricated by electro-spinning have highly<br />

porous microstructure with interconnected pores <strong>and</strong> an extremely large specific surface-area,<br />

which allow accommodation <strong>of</strong> a large number <strong>of</strong> cells <strong>and</strong> facilitate uniform distribution <strong>of</strong><br />

cells <strong>and</strong> diffusion <strong>of</strong> oxygen <strong>and</strong> nutrients. We envisioned that this material could exhibit<br />

improved biocompatibility, biodegradability, solubility <strong>and</strong> processability. More importantly,<br />

it may have promising application as electroactive biomedical materials.<br />

‐ 102 ‐


5.2 <strong>Synthesis</strong> <strong>of</strong> Triblock Copolymer<br />

‐ 103 ‐<br />

<strong>Synthesis</strong> <strong>of</strong> Triblock Copolymers <strong>of</strong> TEMPO‐Acrylamide<br />

<strong>and</strong> Lactate by RAFT Polymerization<br />

The schematic procedures <strong>of</strong> the synthesis <strong>of</strong> the PTAm-b-PLA-b-PTAm copolymer were<br />

shown in Scheme 1. The first step was to prepare PLA bearing two hydroxyl groups at each<br />

chain end. The molecular weight <strong>of</strong> the HO-PLA-OH can be determined from 1 H NMR as<br />

1.0×10 4 (Figure 1A). Then the CPAD, used as a chain transfer agent, was coupled onto the<br />

PLA chain ends via condensation reaction in the presence <strong>of</strong> DCC <strong>and</strong> DMAP. The 1 H NMR<br />

spectra <strong>of</strong> CPAD-PLA-CPAD showed signals at 7.4 – 7.9 ppm, corresponded to the resonance<br />

<strong>of</strong> the protons from the phenyl in the CPAD (Figure 1B). The PTAP-b-PLA-b-PTAP was<br />

synthesized by RAFT polymerization using CPAD-PLA-CPAD as macro-RAFT agent. The<br />

monomer, TAP, is not suitable for ATRP polymerization due to its pendant imine group,<br />

which may competitively complex with the copper <strong>and</strong> form species with lower catalytic<br />

activity. Therefore, we synthesized the block copolymer via RAFT polymerization. Finally,<br />

the PTAm-b-PLA-b-PTAm was obtained after the oxidation <strong>of</strong> PTAP-b-PLA-b-PTAP.<br />

Scheme 1. Synthetic route <strong>of</strong> PTAm-b-PLA-b-PTAm.


Chapter 5<br />

Figure 1. 1 H NMR spectra <strong>of</strong> HO-PLA-OH (A) <strong>and</strong> CPAD-PLA-CPAD (B).<br />

Figure 2 showed the 1 H NMR spectra <strong>of</strong> PTAm-b-PLA-b-PTAm triblock copolymers. All<br />

the resonances attributed to PLA <strong>and</strong> PTAm repeat units were detected, which clearly<br />

indicated the successful preparation <strong>of</strong> the triblock copolymers. The degree <strong>of</strong> polymerization<br />

<strong>of</strong> PTAm block can be calculated based on the integration <strong>of</strong> proton peak c at 4.19 ppm from<br />

piperidine <strong>and</strong> the proton peak a at 5.18 ppm <strong>of</strong> the methine in PLA. Triblock copolymers<br />

with different PTAm block length can be obtained by varying the monomer feed. The<br />

molecular weights <strong>of</strong> the copolymers are summarized in Table 1. The yields were higher than<br />

80%, indicating the reaction is more effective than that described in Chapter 4. The<br />

copolymers had low polydispersities, <strong>and</strong> the GPC pr<strong>of</strong>iles <strong>of</strong> the copolymers were all single<br />

<strong>and</strong> symmetrical peak. These implied that all the macromonomers have initiated the TAP<br />

monomers successfully <strong>and</strong> no homopolymerization occurred.<br />

Figure 2. 1 H NMR spectrum <strong>of</strong> PTAP-b-PLA-b-PTAP.<br />

‐ 104 ‐


Sample<br />

Table 1. Feed <strong>and</strong> result compositions <strong>of</strong> copolymers.<br />

CPAD-PLA-CPAD/Monomer<br />

(mol)<br />

Feed Result a<br />

‐ 105 ‐<br />

<strong>Synthesis</strong> <strong>of</strong> Triblock Copolymers <strong>of</strong> TEMPO‐Acrylamide<br />

<strong>and</strong> Lactate by RAFT Polymerization<br />

Yield<br />

(%)<br />

PDI b<br />

TLT1 1 : 100 1 : 70 81.0 1.40<br />

TLT2 1 : 200 1 : 180 90.0 1.42<br />

a Calculated from 1 H NMR; b Determined from GPC in DMF.<br />

The cyclic voltammogram (CV) <strong>of</strong> the copolymer at different scan rates were shown in<br />

Figure 3. The CV curves also confirmed the success <strong>of</strong> the oxidation.<br />

Figure 3. CV <strong>of</strong> the copolymer (TLT1) at different scan rates in acetonitrile solution <strong>of</strong> 0.1 M<br />

tetrabutylammonium perchlorate.<br />

5.3 Thermal Properties<br />

The typical DSC traces for PLA, TLT1, TLT2 <strong>and</strong> PTAm were collected in Figure 4A. The<br />

PLLA homopolymer has a Tm at 155.6 o C <strong>and</strong> a Tc at 97.4 o C, while TLT1 exhibits a Tm at<br />

160.0 o C <strong>and</strong> TLT2 exhibits no Tm or Tg in the test temperature range. These differences<br />

indicated that the physical properties <strong>of</strong> the PLA in the block copolymers were really<br />

different from those <strong>of</strong> homo-PLA. In the triblock copolymers, the PLA block was still<br />

crystalline, but its crystalline parameters changed much due to the restriction <strong>of</strong> the PTAm


Chapter 5<br />

blocks, which exhibiting no Tm or Tg in the test temperature range. Figure 4B showed the<br />

decomposition behaviors <strong>of</strong> PLA, TLT1, TLT2 <strong>and</strong> PTAm. PLA began to decompose at 200<br />

o C, which originating from the chain end, <strong>and</strong> lost all the weight at 300 o C. The<br />

decomposition temperature <strong>of</strong> PTAm occured at 50 o C <strong>and</strong> lasted to 500 o C. For TLT1 <strong>and</strong><br />

TLT2, their thermo-decomposition behaviors were more similar to PTAm because the end<br />

hydroxyl groups <strong>of</strong> PLA were protected by the PTAm chains.<br />

Figure 4. (A) DSC traces <strong>of</strong> the PLA, PTAm, <strong>and</strong> TLT in second heating. (B) TGA curves <strong>of</strong><br />

PLA, PTAm, <strong>and</strong> TLT.<br />

5.4 Cytotoxicity <strong>and</strong> Biocompatibility <strong>of</strong> Copolymers<br />

In the biomedical application, TLT copolymer should be non-toxic <strong>and</strong> able to support cell<br />

adhesion <strong>and</strong> growth. Like described in Chapter 4, the cytotoxicity <strong>and</strong> cell compatibility <strong>of</strong><br />

the block copolymers were assessed in vitro. To better observe the shape <strong>of</strong> the adherent cells<br />

(particularly whether irregular polygonal or original elliptical), the samples were prepared by<br />

regular cell-staining method, i.e. staining separately with nucleus <strong>and</strong> F-actin.<br />

Figure 5. A: Phase contrast microscopy images <strong>of</strong> RSC96 cells after incubation on different<br />

polymer films for 48h. (1) PLA film, (2) PTAm film, (3) TLT1 film <strong>and</strong> (4) TLT2 film. Scale bar<br />

‐ 106 ‐


= 25 μm. B: MTT assay for the cytotoxicity <strong>of</strong> PLA, PTAm, TLT1 <strong>and</strong> TLT2.<br />

‐ 107 ‐<br />

<strong>Synthesis</strong> <strong>of</strong> Triblock Copolymers <strong>of</strong> TEMPO‐Acrylamide<br />

<strong>and</strong> Lactate by RAFT Polymerization<br />

It can be seen from Figure 5A that cells exhibited original elliptical on PTAm film, while<br />

cells seeded on TLT films extended well as on PLA film. Considering the results shown in<br />

Figure 5B, it can be concluded that block copolymers had improved cell viability <strong>and</strong> were<br />

more suitable to support the proliferation <strong>of</strong> RSC96 cells compared to PTAm homopolymer<br />

in higher concentration.<br />

5.5 Electrospun <strong>of</strong> Copolymers<br />

It is known that the size <strong>and</strong> morphology <strong>of</strong> electrospun materials can be influenced either<br />

by the intrinsic properties <strong>of</strong> solutions such as viscosity, surface tension, <strong>and</strong> conductivity, or<br />

by the spinning parameters such as voltage, flow rate, <strong>and</strong> working distance. 11-14 In our study,<br />

all electro-spinnings were conducted with identical spinning parameters, so the size <strong>and</strong><br />

morphology <strong>of</strong> the nan<strong>of</strong>ibers were closely related to the intrinsic properties <strong>of</strong> the<br />

electro-spinning solutions. As shown in Figure 6, the morphologies <strong>of</strong> the TLT1 electrospun<br />

at various concentrations are quite different. It can be seen that at low concentration (10 wt%)<br />

TLT1 copolymers formed tyre-like shape. As the concentration increased to 20wt%,the<br />

copolymer was electrospun to connected microspheres. When the concentration reached to 30<br />

wt%, the resulted felt exhibited fibroid shape. The shape evolution <strong>of</strong> the polymers is mainly<br />

attributed to the viscosity difference. In the case <strong>of</strong> low viscosity solutions (Figure 6A <strong>and</strong> B),<br />

the jet breaks up into droplets, which also known as electrospraying. For high viscosity<br />

liquids (Figure 6C), the jet does not break up <strong>and</strong> formed continous fibers.<br />

Figure 6. SEM photograph <strong>of</strong> the electrospun TLT1 at concentration <strong>of</strong> 10 wt%(A), 20 wt%<br />

(B) <strong>and</strong> 30 wt% (C). (Scale bar: A <strong>and</strong> B = 50 μm, C = 5 μm)


Chapter 5<br />

5.6 Conclusions<br />

Triblock copolymers composed <strong>of</strong> PLA <strong>and</strong> PTAm with well-defined structure was<br />

prepared by combination <strong>of</strong> ROP <strong>and</strong> RAFT methods. Since the living characteristics <strong>of</strong> the<br />

RAFT polymerization, the triblock copolymers could be obtained with well-defined structure<br />

<strong>and</strong> compositions, which was confirmed by the comparison <strong>of</strong> their yields <strong>and</strong> PDIs with that<br />

<strong>of</strong> graft copolymers synthesized in Chapter 4. Biological evaluation <strong>of</strong> the material showed<br />

that the material has good biocompatibility <strong>and</strong> low cytotoxicity. Electrospinning technology<br />

was then adopted to prepare the electro-active porous polymer matrices with relatively large<br />

surface area for biomedical applications. The morphology <strong>of</strong> electrospun copolymers is<br />

mainly attributed to the viscosity difference from tyre-like shape at lower concentration to<br />

fibroid shape at higher concentration.<br />

5.7 Experimental Section<br />

Materials<br />

Lactide was purchased from Purac, Holl<strong>and</strong>. Stannous octoate (Sn(Oct)2, 95%) <strong>and</strong><br />

1,4-butanediol (BDO) were purchased from Aldrich. 2, 2'-Azoisobutyronitrile (AIBN, Beijing<br />

Chemical Co., China) was recrystallized twice from methanol. 4-Cyanopentnoic acid<br />

dithiobenzoate (CPAD) was prepared as described in literature.<br />

4-Amino-2,2,6,6-tetramethylpiperidine (TCI, >97%) <strong>and</strong> Acrylyl chloride (Alfa Aesar, 96%)<br />

were used to prepare the 2,2,6,6-tetra-methylpiperidinyl acrylamide (TAP) monomer.<br />

N,N'-Dicyclohexylcarbodiimide (DCC, 99%) <strong>and</strong> 4-dimethylaminopyridine (DMAP, 98%)<br />

were purchased from Sigma Chemical Co. <strong>and</strong> used as received. All other reagents were used<br />

without further purification.<br />

Instrument<br />

1 H NMR spectra were recorded on Bruker AV 400 NMR spectrometer in CDCl3. Gel<br />

permeation chromatography was performed with a DMF eluent using a Tosoh HLC-8220<br />

instrument, <strong>and</strong> the molecular weight <strong>and</strong> polydispersity were calibrated with polystyrene<br />

st<strong>and</strong>ards. Scanning electron microscope (SEM) images were obtained by using an ESEM<br />

(Model XL 30 ESEM FEG from Micro FEI Philips). The specimens were mounted on metal<br />

stubs using a double-sided adhesive tape <strong>and</strong> vacuum coated with a platinum layer prior to<br />

examination. Samples for cyclic voltammetry were prepared by depositing thin LM PAP film<br />

‐ 108 ‐


‐ 109 ‐<br />

<strong>Synthesis</strong> <strong>of</strong> Triblock Copolymers <strong>of</strong> TEMPO‐Acrylamide<br />

<strong>and</strong> Lactate by RAFT Polymerization<br />

on an indium tin oxide (ITO) electrode as working electrode <strong>and</strong> Ag/Ag + as reference<br />

electrode. Cyclic voltammograms were recorded on a CHI 630 potentiostat with different<br />

scanning rate.<br />

<strong>Synthesis</strong> <strong>of</strong> Hydroxyl-capped PLA<br />

After recrystallization in ethyl acetate for three times, lactide (1 mol) was added to a<br />

flame-dried <strong>and</strong> nitrogen-purged glass ampoule, into which a 150 mL toluene solution <strong>of</strong><br />

BDO (7.2 mmol) <strong>and</strong> Sn(Oct)2 (0.3% <strong>of</strong> the BDO, mol/mol) was then transferred. The<br />

reaction vessel was immersed into a thermostatic oil bath maintained at 120 o C, under<br />

magnetic stirring for 24 h. The reaction product was precipitated into ethanol, filtered <strong>and</strong><br />

dried at 40 o C in vacuum for 48 h.<br />

<strong>Synthesis</strong> <strong>of</strong> Macromolecular CTA: CPAD-PLA-CPAD<br />

Typically, PLA 5 g (0.5 mmol), CPAD 0.28 g (1mmol), DCC 0.42 g (2 mmol) <strong>and</strong> DMAP<br />

0.025 g (0.2 mmol) were dissolved in 20 mL anhydrous methylene chloride. The solution was<br />

stirred at room temperature for 48 hours. After filtration <strong>of</strong> the precipitate, the solution was<br />

poured into 200 mL cold ethyl ether, yielding pink precipitate. The pink precipitate was<br />

redissolved in methylene chloride <strong>and</strong> precipitated again in cold ethyl ether. After washing<br />

with additional two times with ethyl ether, the pink product CPAD-PLA-CPAD was finally<br />

obtained, Yield 92%.<br />

<strong>Synthesis</strong> <strong>of</strong> PTAm-b-PLA-b-PTAm<br />

Typically, 0.20 g CPAD-PLA-CPAD (0.02 mmol), 0.29 g (1.4 mmol) TAP monomer <strong>and</strong><br />

1.6 mg AIBN (0.01 mmol) ware dissolved in 2 mL THF. The mixture was deaerated by<br />

applying three times freeze-pump-thaw cycle. Afterwards, the tube was immerged in 90 o C<br />

oil bath <strong>and</strong> stirred for 12 hours. After cooling to room temperature, the mixture was<br />

precipitate in cold ethyl ether. The slight pink precipitate was collected <strong>and</strong> washed twice<br />

with ethyl ether. The product was obtained after vacuum drying for 24 h with a yield <strong>of</strong> 81%.<br />

In order to obtain PTAm-b-PLA-b-PTAm, an oxidation step was needed. Typically, 110 mg <strong>of</strong><br />

the aforementioned PTAP-b-PLA-b-PTAP dissolved in 5 mL THF <strong>and</strong> stirred in ice-cold bath,<br />

to which 500 mg <strong>of</strong> m-chloroperbenzoic acid (mCPBA) was added. The mixture was allowed<br />

to move to the room temperature <strong>and</strong> stirred for further 4 hours. The final product was<br />

obtained by precipitate the mixture into 120 mL ethyl ether/n-hexane (v/v, 1:1), Yield: 92%.


Chapter 5<br />

Cytotoxicity assay, Cell Adhesion <strong>and</strong> Spreading<br />

See Chapter 4, experimental section.<br />

Electrospinning Process<br />

Polymers (PLA, PTAm or PTAm-b-PLA-b-PTAm) were dissolved in dichloromethane to<br />

achieve a certain concentration. The solution was stirred for 48 h at ice water bath <strong>and</strong> then<br />

loaded into the spinner with a capillary outlet <strong>of</strong> 1 mm in diameter, which as an anode was 20<br />

cm apart from the grounded collecting drum. The voltage applied between the spinner outlet<br />

<strong>and</strong> the drum was 15 kV. The resulted mats were collected on an aluminum sheet fixed on the<br />

rotating drum for 3 h. Then the obtained product was dried in a vacuum oven for 24 h.<br />

Reference<br />

1. M. F. Semmelhack, C. R. Schmid, D. A. Cortes, C. S. Chou, J. Am. Chem. Soc. 1984, 106,<br />

3374.<br />

2. H. Nishide, T. Kaneko, Magnetic properties <strong>of</strong> organic material. Dekker: New York,<br />

1999.<br />

3. E. E. Voest, E. Vanfaassen, J. J. M. Marx, Free Radical Biol. Med. 1993, 15, 589.<br />

4. K. Koshika, N. Sano, K. Oyaizu, H. Nishide, Chem. Commun. 2009, 7, 836.<br />

5. R. Barbey, L. Lavanant, D. Paripovic, N. Schuwer, C. Sugnaux, S. Tugulu, H. A. Klok,<br />

Chem. Rev. 2009, 109, 5437.<br />

6. A. Vora, K. Singh, D. C. Webster, Polymer 2009, 50, 2768.<br />

7. J. L. Zhu, X. Z. Zhang, H. Cheng, Y. Y. Li, Cheng, S. X.; Zhuo, R. X., J. Polym. Sci.,<br />

Part A: Polym. Chem. 2007, 45, 5354.<br />

8. A. K. Ekaputra, Y. F. Zhou, S. M. Cool, D. W. Hutmacher, Tissue Engineering Part A<br />

2009, 15, 3779.<br />

9. X. L. Xu, X. S. Chen, A. X. Liu, Z. K. Hong, X. B. Jing, Eur. Polym. J. 2007, 43, 3187.<br />

10. Y. L. Hong, Y. A. Li, X. L. Zhuang, X. S. Chen, X. B. Jing, J. Biomed. Mater. Res. Part A<br />

2009, 89A, 345.<br />

11. M. Li, M. J. Mondrinos, M. R. G<strong>and</strong>hi, F. K. Ko, A. S. Weiss, P. I. Lelkes, Biomaterials<br />

2005, 26, 5999.<br />

12. J. A. Matthews, G. E. Wnek, D. G. Simpson, G. L. Bowlin, Biomacromolecules 2002, 3,<br />

232.<br />

‐ 110 ‐


‐ 111 ‐<br />

<strong>Synthesis</strong> <strong>of</strong> Triblock Copolymers <strong>of</strong> TEMPO‐Acrylamide<br />

<strong>and</strong> Lactate by RAFT Polymerization<br />

13. J. M. Deitzel, J. Kleinmeyer, D. Harris, N. C. Beck Tan, Polymer 2001, 42, 261.<br />

14. Y. J. Ryu, H. Y. Kim, K. H. Lee, H. C. Park, D. R. Lee, Eur. Polym. J. 2003, 39, 1883.


Chapter 6<br />

<strong>Synthesis</strong> <strong>of</strong> Amphiphilic Triblock Copolymers <strong>of</strong> Acrylamide,<br />

6.1 Inroduction<br />

Lactate, <strong>and</strong> Ethyleneoxide<br />

6.2 <strong>Synthesis</strong> <strong>of</strong> MPEG-CPAD <strong>and</strong> MPEG-b-PLA-CPAD<br />

6.3 <strong>Synthesis</strong> <strong>of</strong> MPEG-b-PTAm <strong>and</strong> MPEG-b-PLA-b-PTAm<br />

6.4 Micelles preparation <strong>and</strong> characterization<br />

6.5 EPR study <strong>of</strong> the micelles in PBS containing ascorbic acid<br />

6.6 Cytotoxicity assay<br />

6.7 Conclusions<br />

6.8 Experimental section<br />

References


Chapter 6<br />

6.1 Introduction<br />

Since the discovery <strong>of</strong> nitroxyl radicals in 1956, there have been <strong>of</strong> continually pr<strong>of</strong>ound<br />

interest in the synthesis <strong>and</strong> investigation <strong>of</strong> compounds containing nitroxyl free radicals.<br />

Applications such as spin-labeling/trapping, MR/ESR in vivo imaging, oxidation<br />

organocatalysts <strong>and</strong> mediating radical polymerization have emerged based on the unique<br />

physical, chemical <strong>and</strong> biological properties <strong>of</strong> the nitroxyl radicals. Polymers bearing stable<br />

nitroxyl radicals in the pendant, known as radical polymers, also have been investigated<br />

during past several decades initially for their potential reactivities, however, it is not until<br />

recently that the radical polymers was intensively investigated for organic electronic devices<br />

applications. A series <strong>of</strong> radical polymers have been synthesized <strong>and</strong> evaluated by our group<br />

as the electro-active materials for secondary batteries <strong>and</strong> organic memory devices. 1-4<br />

Aside from the electronic applications, radical polymers are also receiving increasing<br />

interest for the application as biomaterials due to nitroxyl radical which plays critical roles in<br />

various metabolic processes. C. C. Chu <strong>and</strong> co-workers are pioneers in this field by<br />

conjugating 2,2,6,6-tetramethylpiperidinyloxy (TEMPO) radicals with biodegradable<br />

polymers. The resultant TEMPO-functionalized polymers could be degraded in the<br />

physiological condition to release the TEMPO nitroxyl radicals, which may found potential<br />

use as the scaffold or surface coating for the cardiovascular tissue engineering, the<br />

macromolecular drugs for anticancer therapy <strong>and</strong> the wound dressing with improved healing<br />

<strong>and</strong> antimicrobial capability. 5-7 On the other h<strong>and</strong>, the magnetic resonance relaxivity <strong>and</strong><br />

electron paramagnetic resonance (EPR) property have endowed the nitroxyl radical-contained<br />

materials with the potential application as bioimaging probes. Much attentions have been paid<br />

on the development <strong>of</strong> the nitroxyl radical-based magnetic resonance imaging (MRI) probes<br />

since they have lower toxicity in comparison with the commercially gadolinium derivatives<br />

agents. 8-10 However, the MR relaxivities <strong>of</strong> the nitroxyl radical is relatively lower. In order to<br />

solve this issue, several strategies such as copolymerization or attachment <strong>of</strong> nitroxyl radical<br />

compound to the water soluble hyperbranched polymer have been investigated. 11-14 Also, the<br />

nitroxyl radical compounds, such as TEMPO, are now being investigated as EPR probes<br />

applied in vivo because <strong>of</strong> the sensitivity <strong>of</strong> the EPR is much higher than that <strong>of</strong> MRI. 15-19<br />

However, the reducing environment in vivo, which could reduce or eliminate the EPR signal<br />

by reducing the nitroxyl radical, has brought a great obstacle to the application <strong>of</strong> these<br />

‐ 114 ‐


<strong>Synthesis</strong> <strong>of</strong> Amphiphilic Triblock Copolymers <strong>of</strong> Acrylamide, Lactate, <strong>and</strong> Ethyleneoxide<br />

materials. It has been reported that the half-time <strong>of</strong> free TEMPO radical in the blood was<br />

about 15 s which was far from satisfactory for practical use. 20<br />

Therefore, our purpose is to develop a kind <strong>of</strong> polymeric deliver system that can carry<br />

stable nitroxyl radicals <strong>and</strong> resist to the reduction environment, which may find potential use<br />

as the EPR probe for bioimaging in vivo. For this study, amphiphilic triblock copolymer,<br />

consisting <strong>of</strong> a hydrophilic PEG segment, a hydrophobic PLA segment <strong>and</strong> a TEMPO<br />

radical-contained hydrophobic PTAm segment, was synthesized. An amphiphilic diblock<br />

copolymer only contains PEG <strong>and</strong> PTAm segments were also synthesized. Both <strong>of</strong> them can<br />

self-assemble into micelles in aqueous solution with radicals in the core <strong>and</strong> generate stable<br />

EPR signal. The stability <strong>of</strong> the micelles against reducing environment was also evaluated.<br />

6.2 <strong>Synthesis</strong> <strong>of</strong> MPEG-CPAD <strong>and</strong> MPEG-b-PLA-CPAD<br />

4-Cyanopentnoic acid dithiobenzoate (CPAD) was prepared as described in the literature<br />

26. 1 H NMR spectra with proton resonance absorption assignments was shown in Figure 6.1<br />

A. Then, the macromolecular chain transfer agents, MEPG-CPAD <strong>and</strong> MPEG-b-PLA-CPAD,<br />

were synthesized by chemical coupling the CPAD to the chain-end <strong>of</strong> MPEG <strong>and</strong><br />

MPEG-b-PLA in the presence <strong>of</strong> DCC <strong>and</strong> DMAP, respectively. The structures <strong>of</strong> the<br />

obtained polymers were verified by 1 H NMR characterization as shown in Figure 6.1 with the<br />

presence <strong>of</strong> the proton resonance absorption peaks from CPAD.<br />

Figure 6.1 1 H NMR characterizations <strong>of</strong> (A) CPAD; (B) MPEG-CPAD; (C) MPEG-b-PLA-CPAD<br />

‐ 115 ‐


Chapter 6<br />

6.3 <strong>Synthesis</strong> <strong>of</strong> MPEG-b-PTAm <strong>and</strong> MPEG-b-PLA-b-PTAm<br />

Controlled living radical polymerization such as nitroxide-mediated polymerization<br />

(NMP), atom transfer radical polymerization (ATRP) <strong>and</strong> reversible addition-fragmentation<br />

chain transfer (RAFT) polymerization have recently received great attention for their ability<br />

to synthesize polymer materials with controlled molecular weight, various functional groups<br />

<strong>and</strong> diverse architectures. We used RAFT technique to obtain the novel block copolymers in<br />

virtue <strong>of</strong> the convenient <strong>and</strong> effective RAFT polymerization towards TAP monomer. The<br />

preparation <strong>of</strong> MPEG-b-PTAm <strong>and</strong> MPEG-b-PLA-b-PTAm precursor polymers were<br />

conducted in 1,4-dioxane/methanol mixture using MPEG-CPAD <strong>and</strong> MPEG-b-PLA-CPAD as<br />

the macromolecular chain transfer agent (CTA), respectively (Scheme 1 <strong>and</strong> Scheme 2).<br />

Scheme 6.1 <strong>Synthesis</strong> <strong>of</strong> macromolecular chain transfer agents (CTA).<br />

To better underst<strong>and</strong> the polymerization process, the kinetic <strong>of</strong> the RAFT polymerization<br />

<strong>of</strong> TAP was studied <strong>and</strong> the results were shown in Figure 6.2. The polymerization could reach<br />

a high monomer conversion up to ~87% within 6 hours. The linear dependence <strong>of</strong> the<br />

molecular weight on the monomer conversion strongly indicated that the RAFT<br />

polymerization towards TAP monomer had living characters. However, the molecular weight<br />

distribution <strong>of</strong> the polymer was increased from 1.20 at 0 hour to 1.57 at 6 hour, <strong>and</strong> further to<br />

1.73 if the polymerization proceeded more than10 hours which may be assigned either to<br />

intermolecular chain transfer or possibly to cross-termination at the high monomer<br />

conversions. This feature has already been observed in most <strong>of</strong> the dithiobenzoate-mediated<br />

RAFT system.<br />

‐ 116 ‐


<strong>Synthesis</strong> <strong>of</strong> Amphiphilic Triblock Copolymers <strong>of</strong> Acrylamide, Lactate, <strong>and</strong> Ethyleneoxide<br />

Scheme 6.2 <strong>Synthesis</strong> <strong>and</strong> self-assembly <strong>of</strong> MPEG-b-PTAm <strong>and</strong> MPEG-b-PLA-b-PTAm<br />

Figure 6.2 RAFT polymerization <strong>of</strong> TAP monomer using MEPG-CPAD as the CTA. (A)<br />

Monomer conversion versus time; (B) Dependence <strong>of</strong> molecular weights <strong>and</strong> molecular weight<br />

distributions <strong>of</strong> the block copolymers on the monomer conversion.<br />

Based on the aforementioned study, we then conducted the preparation <strong>of</strong> the block<br />

copolymer for 8 hours in order to achieve high monomer conversion <strong>and</strong> meanwhile<br />

minimum side reactions which might induce broaden <strong>of</strong> the molecular weight distribution. 1 H<br />

NMR was used to study the structure <strong>of</strong> the precursor block copolymers. The spectra shown<br />

in Figure 6.3 clearly confirmed the structure <strong>of</strong> the block copolymers. The degree <strong>of</strong><br />

polymerization was calculated to be 46 <strong>and</strong> 42 for MPEG-b-PTAm <strong>and</strong><br />

MPEG-b-PLA-b-PTAm, respectively, based on the integration <strong>of</strong> proton peak c at 4.19 ppm<br />

from piperidine <strong>and</strong> the proton peak b at 3.64 ppm for the methylene in MPEG. The obtained<br />

precursor block copolymers were then oxidized by m-chloroperbenzoic acid to convert the<br />

N-H groups into nitroxyl radicals, leading to MPEG-b-PTAm <strong>and</strong> MPEG-b-PLA-b-PTAm<br />

block copolymers with nitroxyl radical-contained segment. 1 H NMR (data not shown) was<br />

‐ 117 ‐


Chapter 6<br />

used to characterize the resultant block copolymer, the disappearance <strong>of</strong> the resonance<br />

absorption from the protons in the PTAm segment indicated that the N-H groups had been<br />

oxidized into paramagnetic nitroxyl radicals that would induce no magnetic resonance<br />

absorption. The GPC results (Figure 6.4) also revealed the two obtained block copolymers<br />

with unimodel GPC traces, demonstrating the successful preparation <strong>of</strong> the block<br />

copolymers.<br />

Figure 6.3<br />

1 H NMR spectra <strong>of</strong> (A) MPEG-b-PTAm precursor polymer; (B)<br />

MPEG-b-PLA-b-PTAm precursor polymer.<br />

Figure 6.4 GPC traces <strong>of</strong> MPEG-b-PTAm <strong>and</strong> MPEG-b-PLA-b-PTAm<br />

‐ 118 ‐


6.4 Micelles preparation <strong>and</strong> characterization<br />

<strong>Synthesis</strong> <strong>of</strong> Amphiphilic Triblock Copolymers <strong>of</strong> Acrylamide, Lactate, <strong>and</strong> Ethyleneoxide<br />

The micelles were prepared by dialysis method. The weighted block copolymer was<br />

dissolved in THF <strong>and</strong> subsequently added dropwise into deionic water to form micelles<br />

aggregation. The mixtures were then dialyzed against deionic water to remove the residual<br />

THF. The obtained micelles solutions were diluted to a concentration <strong>of</strong> 1.0 mg.mL -1 . The<br />

average hydrodynamic diameters for MPEG-b-PTAm <strong>and</strong> MPEG-b-PLA-b-PTAm micelles<br />

were characterized by DLS to be 48 nm <strong>and</strong> 58 nm (see Figure 6.5 C), respectively. The<br />

slightly increased size <strong>of</strong> the micelles prepared from MPEG-b-PLA-b-PTAm may be ascribed<br />

to the more hydrophobic molecular weight ratio <strong>of</strong> MPEG-b-PLA-b-PTAm compared to that<br />

<strong>of</strong> MPEG-b-PTAm. 21 TEM was also used to confirm the micelles structure <strong>of</strong> the two block<br />

copolymers as shown in Figure 6.5 A <strong>and</strong> B. Cyclic voltammograms <strong>of</strong> MPEG-b-PTAm <strong>and</strong><br />

MPEG-b-PLA-b-PTAm micelles were also conducted in aqueous HCl solution at a scanning<br />

rate <strong>of</strong> 10 mV/min. The results showed in Figure 6.6 demonstrated that both <strong>of</strong> the micelles<br />

produced a reversible redox wave at around 0.7 V vs Ag/AgCl, indicating that the nitroxyl<br />

radicals had been successfully incorporated into the micelles.<br />

Figure 6.5 TEM <strong>and</strong> DLS characterizations <strong>of</strong> MPEG-b-PTAm micelles (A) <strong>and</strong><br />

MPEG-b-PLA-b-PTAm micelles (B)<br />

‐ 119 ‐


Chapter 6<br />

Figure 6.6 Cyclic voltammograms <strong>of</strong> (A) MPEG-b-PTAm <strong>and</strong> (B) MPEG-b-PLA-b-PTAm<br />

micelles in 0.1 M HCl aqueous solution.<br />

6.5 EPR study <strong>of</strong> the micelles in PBS containing ascorbic acid<br />

It is known that a reducing environment is necessary for biological systems. The<br />

antioxidants such as ascorbic acid (Vc) <strong>and</strong> glutathione (GSH) are important to keep the<br />

redox status in living cells <strong>and</strong> exist in most <strong>of</strong> parts in vivo. Thus, we evaluated the reduction<br />

resistance <strong>of</strong> the micelles in the presence <strong>of</strong> antioxidants. Since the reduction <strong>of</strong> nitroxyl<br />

radical into hydroxylamine group by GSH is much slower than Vc <strong>and</strong> there is higher<br />

concentration <strong>of</strong> Vc in the blood stream, 22, 23 we selected antioxidant Vc for our study. The<br />

EPR measurements were conducted in the presence <strong>of</strong> Vc at concentration <strong>of</strong> 5.0 mM which<br />

is 50 times more than the concentration <strong>of</strong> Vc in the blood.<br />

As shown in Figure 6.7 A, the EPR signal decreased over time in PBS solution<br />

containing Vc, ascribing to the reduction <strong>of</strong> nitroxyl radicals. And the EPR signal pattern <strong>of</strong><br />

both MPEG-b-PTAm <strong>and</strong> MPEG-b-PLA-b-PTAm micelles were broad single, which is<br />

distinct from typical triplet signal <strong>of</strong> free TEMPOL radical in diluted solution, due to the<br />

spin-spin interaction <strong>of</strong> TEMPO in the side chains. 19, 24, 25 Furthermore, we observed that the<br />

EPR signal in aqueous media is broader than in chlor<strong>of</strong>orm, which is consisted with previous<br />

report. 18 These results indicated that the polymer segment containing nitroxyl radicals had<br />

been incorporated into micelles in aggregated state in the hydrophobic core. By virtue <strong>of</strong> the<br />

micelle structure, the nitroxyl radicals in the core could resist reduction for a longer time than<br />

free TEMPOL radical even in the presence <strong>of</strong> higher concentration <strong>of</strong> Vc (Figure 6.7 B). The<br />

half-life time for both MPEG-b-PTAm <strong>and</strong> MPEG-b-PLA-b-PTAm micelles were about 8<br />

‐ 120 ‐


<strong>Synthesis</strong> <strong>of</strong> Amphiphilic Triblock Copolymers <strong>of</strong> Acrylamide, Lactate, <strong>and</strong> Ethyleneoxide<br />

min, however, the time for free TEMPOL radical was less than 1 min. The prolonging time<br />

for both MPEG-b-PTAm <strong>and</strong> MPEG-b-PLA-b-PTAm micelles in the reducing environment<br />

would endow the systems have the promising application as the stable EPR probes.<br />

Figure 6.7 (A) EPR spectra <strong>of</strong> MPEG-b-PTAm <strong>and</strong> MPEG-b-PLA-b-PTAm micelles in PBS<br />

containing 5.0 mM ascorbic acid over time; (B) Dependence <strong>of</strong> EPR signal intensity on the time.<br />

6.6 Cytotoxicity assay<br />

MTT cell toxicity assay was used to determine the cytoxicity <strong>of</strong> the synthesized block<br />

copolymers MEPG-b-PTAm <strong>and</strong> MPEG-b-PLA-b-PTAm. The results shown in Figure 6.8<br />

revealed that both MEPG-b-PTAm <strong>and</strong> MPEG-b-PLA-b-PTAm had as low cytotoxicity as the<br />

MPEG5K <strong>and</strong> MPEG-b-PLA. The good biocompatibility <strong>of</strong> the MEPG-b-PTAm <strong>and</strong><br />

MPEG-b-PLA-b-PTAm, presumably ascribing to the hydrophiphic shell <strong>of</strong> the MPEG<br />

segment in aqueous solution, would make these materials suitable for applications in vivo. In<br />

‐ 121 ‐


Chapter 6<br />

vivo cytotoxicity was also determined by injecting the micelles solution into the kunming<br />

mice (ca. 17~23 g). 0.5 mL <strong>of</strong> micelles solution at concentrations <strong>of</strong> 2 mg/mL, 5mg/mL <strong>and</strong><br />

15mg/ml were injected into the five kunming mice in each group. All the mice were observed<br />

to show no significant indisposed nor death during the 2 day’s observation, indicating the<br />

good in vivo biocompatibility <strong>of</strong> the micelles.<br />

Figure 6.8 MTT cell toxicity assay. B1: MPEG-b-PLA; B2: MPEG-b-PTAm; B3: MPEG-b-PLA-b-PTAm.<br />

6.7 Conclusions<br />

Amphiphilic triblock copolymer containing stable nitroxyl radicals was successfully<br />

prepared by combination <strong>of</strong> ring-opening polymerization <strong>and</strong> RAFT technique. The RAFT<br />

polymerization towards TAP monomer was monitored to show living features by using<br />

MPEG-CPAD as the CTA. Amphiphilic diblock copolymer containing stable nitroxyl radicals<br />

but without PLA segment was also prepared for the control study. The block copolymer<br />

structures were confirmed by 1 H NMR <strong>and</strong> GPC characterization. Both <strong>of</strong> the copolymers can<br />

self-assemble into micelles in aqueous solution with nitroxyl radicals in the hydrophobic core.<br />

All the micelles prepared from either triblock or diblock copolymer can produce EPR signal<br />

<strong>and</strong> protect the free radicals from being reduced for a longer time as compared to the free<br />

TEMPOL radicals in the presence <strong>of</strong> high concentration Vc. Additionally, the micellar<br />

structure also endow the two materials with low cytotoxicity. Therefore, such micelles with<br />

stable free radicals in the core should be promising to be used as the EPR probe for in vivo<br />

imaging application.<br />

‐ 122 ‐


6.8 Experimental section<br />

Materials <strong>and</strong> methods<br />

<strong>Synthesis</strong> <strong>of</strong> Amphiphilic Triblock Copolymers <strong>of</strong> Acrylamide, Lactate, <strong>and</strong> Ethyleneoxide<br />

2,2’-Azoisobutyronitrile (AIBN, Beijing Chemical Co., China) was recrystallized twice<br />

from methanol. Monomethoxy poly(ethylene glycol) (MPEG Mw=5000 Da),<br />

4-hydroxyl-TEMPO (TEMPOL, 97%) N,N'-Dicyclohexylcarbodiimide (DCC, 99%) <strong>and</strong><br />

4-dimethylaminopyridine (DMAP, 98%) were purchased from Aldrich Chemical Co. <strong>and</strong><br />

used as received. 4-Cyanopentnoic acid dithiobenzoate (CPAD) was prepared as described in<br />

the literature 21. 26 4-Amino-2,2,6,6-tetramethylpiperidine (TCI, >97%) <strong>and</strong> Acryloyl chloride<br />

(Alfa Aesar, 96%) were used without purification to prepare the<br />

2,2,6,6-tetramethylpiperidinyl acrylamide (TAP) monomer. 27 All other reagents were used<br />

without further purification.<br />

1<br />

H NMR spectra were recorded on Bruker AV 300 or Bruker AV 600 NMR spectrometer<br />

in CDCl3. Gel permeation chromatography was performed with 0.1 M LiCl DMF eluent using<br />

a Tosoh HLC-8220 instrument, <strong>and</strong> the molecular weight <strong>and</strong> polydispersity were calibrated<br />

with polystyrene st<strong>and</strong>ards. For kinetic study, gel permeation chromatography was performed<br />

with a Schamback SFD GmnH mGPC2000 instrument using 0.1 M LiCl DMF as the eluent at<br />

40 o C, <strong>and</strong> the molecular weight <strong>and</strong> polydispersity were calibrated with polystyrene<br />

st<strong>and</strong>ards. Dynamic light scattering (DLS) were determined by DAWN EOS 18 Angles Laser<br />

Light Scattering Instructment (Wyatt Technology). Transmission electron microscopy (TEM<br />

JEM-2010 electron microscope, JEOL, Japan) was also used to characterize the morphology<br />

<strong>of</strong> the micelles. Electron paramagnetic resonance (EPR) spectra were recorded on a<br />

JES-FA200 EPR spectrometer.<br />

<strong>Synthesis</strong> <strong>of</strong> Macromolecular CTA: MPEG-CPAD <strong>and</strong> MPEG-b-PLA-CPAD.<br />

Typically, MPEG5k 2.5 g (0.5 mmol), CPAD 0.28 g (1mmol) DCC 0.42 g (2 mmol) <strong>and</strong><br />

DMAP 0.025 g (0.2 mmol) were dissolved in 20 mL anhydrous methylene chloride. The<br />

solution was stirred at room temperature for 48 hours. After filtration <strong>of</strong> the precipitate, the<br />

solution was poured into 200 mL cold ethyl ether, yielding pink precipitate. The pink<br />

precipitate was redissolved in methylene chloride <strong>and</strong> precipitated again in cold ethyl ether.<br />

After washing with additional two times with ethyl ether, the pink product MPEG-CPAD was<br />

finally obtained, yield 92%. The preparation <strong>of</strong> MPEG-b-PLA-CPAD was similar as<br />

MPEG-CPAD, yield 89%. The structure <strong>of</strong> the two macromolecular CTA was verified by 1 H<br />

‐ 123 ‐


Chapter 6<br />

NMR. <strong>Characterization</strong> <strong>of</strong> MPEG-CPAD: IR: � ~ (cm-1) 1737 (C=O) 1044 (C=S). 1 H NMR<br />

(300M, CDCl3, δ ppm): 7.89 7.57 7.40 (5H, CH×3 in benzene), 4.27 (2H, CH2OC(O)-), 3.64<br />

(CH2 in PEG), 3.38 (3H, CH3O-), 2.46~2.75 (4H, CH2×2 in CPAD), 1.94 (3H, CH3 in CPAD).<br />

<strong>Characterization</strong> <strong>of</strong> MPEG-b-PLA-CPAD: IR: � ~ (cm-1) 1755 (C=O) 1044 (C=S). 1 H NMR<br />

(300M, CDCl3, δ ppm): 7.88 7.57 7.40 (5H, CH×3 in benzene), 5.15 (CH in PLA), 4.26 (2H,<br />

CH2OC(O)-), 3.62 (CH2 in PEG), 3.36 (3H, CH3O-), 2.46~2.75 (4H, CH2×2 in CPAD), 1.95<br />

(3H, CH3 in CPAD).<br />

<strong>Synthesis</strong> <strong>of</strong> MPEG-b-poly(2,2,6,6-tetramethylpiperidinyloxy acrylamide)<br />

(PEG-b-PTAm).<br />

For kinetic study, 0.22 g MPEG-CPAD (0.042 mmol), 0.53 g TAP (2.5 mmol) <strong>and</strong> 3.4 mg<br />

AIBN (0.021 mmol) were dissolved in 8 mL 1,4-dioxane/methanol (v/v, 3:1). The mixture<br />

was deaerated by applying three times freeze-pump-thaw cycle. Afterwards, the tube was<br />

immerged in 90 o C oil bath. Samples were taken out at various time intervals <strong>and</strong> quenched<br />

immediately into liquid nitrogen. The freezing samples were lyophilized to gain pink powders<br />

under reduced pressure <strong>and</strong> then used for 1 H NMR analysis to determine the monomer<br />

conversions. Molecular weights <strong>and</strong> molecular weight distributions were determined by GPC<br />

after the oxidation <strong>of</strong> the samples by m-chloroperbenzoic acid (mCPBA) in THF. The<br />

detailed oxidation processes are as follow.<br />

For preparation <strong>of</strong> MPEG-b-PTAm, 0.10 g MPEG-CPAD (0.019 mmol), 0.25 g TAP<br />

monomer (1.2 mmol) <strong>and</strong> 1.6 mg AIBN (0.01 mmol) were dissolved in 4mL<br />

1,4-dioxane/methanol (v/v, 3:1). The mixture was deaerated by applying three times<br />

freeze-pump-thaw cycle. Afterwards, the tube was immerged in 90 o C oil bath <strong>and</strong> stirred for<br />

8 hours. After cooling to room temperature, the mixture was precipitated from cold ethyl<br />

ether. The slight pink precipitate was collected <strong>and</strong> washed twice with ethyl ether. The<br />

product was obtained after vacuum drying for 24 h with a yield <strong>of</strong> 81% (0.29 g). In order to<br />

gain MPEG-b-PTAm, an oxidation step was needed. Typically, 110 mg <strong>of</strong> the aforementioned<br />

product dissolved in 5 mL THF <strong>and</strong> stirred in ice-cold bath, to which 500 mg <strong>of</strong> mCPBA was<br />

added. The mixture was allowed to move to the room temperature <strong>and</strong> stirred for further 4<br />

hours. The final product was obtained by precipitating the mixture into 120 mL ethyl<br />

ether/n-hexane (v/v, 1:1), yield 92%. (Mn=25200, PDI=1.38 by GPC).<br />

‐ 124 ‐


<strong>Synthesis</strong> <strong>of</strong> MPEG-b-PLA-b-PTAm.<br />

<strong>Synthesis</strong> <strong>of</strong> Amphiphilic Triblock Copolymers <strong>of</strong> Acrylamide, Lactate, <strong>and</strong> Ethyleneoxide<br />

0.56 g MPEG-b-PLA-CPAD (0.04 mmol), 0.50 g (2.4 mmol) TAP monomer <strong>and</strong> 3.3 mg<br />

AIBN (0.02 mmol) ware dissolved in 6mL 1,4-dioxane/methanol (9:1, v/v). The mixture was<br />

deaerated by applying three times freeze-pump-thaw cycle. Afterwards, the tube was<br />

immerged in 90 o C oil bath <strong>and</strong> stirred for 8 hours. After cooling to room temperature, the<br />

mixture was precipitate in cold ethyl ether. The pale pink precipitate was collected <strong>and</strong><br />

washed twice with ethyl ether. The product was obtained after vacuum drying for 24 h with a<br />

yield <strong>of</strong> 79% (0.84 g). An additional oxidation procedure was also needed to gain the final<br />

product MPEG-b-PLA-b-PTAm in good yield. (Mn=32300, PDI=1.42 by GPC)<br />

Micelles preparation <strong>and</strong> characterization.<br />

The preparation procedures for the MPEG-b-PTAm <strong>and</strong> MPEG-b-PLA-b-PTAm micelles<br />

are the same. Typically, 50 mg <strong>of</strong> the block copolymer was firstly dissolved in 5 mL <strong>of</strong> THF<br />

<strong>and</strong> then added dropwise into 10 mL deionized water. The mixture was then transferred into a<br />

dialysis bag (molecular weight cut-<strong>of</strong>f = 3500 Da) <strong>and</strong> dialyzed against deionized water for<br />

24h. Afterwards, the micelles solution was diluted to 50 mL with deioinized water. The final<br />

concentrations <strong>of</strong> the micelles were 1.0 mg mL -1 . The structure <strong>of</strong> the micelles was<br />

characterized by DLS <strong>and</strong> TEM.<br />

EPR study.<br />

To evaluate the stability <strong>of</strong> the nitroxyl radicals containing micelles in the reducing<br />

environment, EPR study <strong>of</strong> the micelles in PBS buffering solution containing ascorbic acid at<br />

high concentration was conducted. 0.5 mL <strong>of</strong> the prepared micelle solution (1.0 mg.mL -1 )<br />

was mixed vigorously with 0.5 mL <strong>of</strong> PBS solution (pH 7.4, 0.2 M containing 10.0 mM<br />

ascorbic acid). The mixture was then transferred into capillary tube immediately for EPR<br />

measurement over time. The final concentrations <strong>of</strong> the micelle <strong>and</strong> the ascorbic acid were<br />

0.5 mg.mL -1 <strong>and</strong> 5.0 mM, respectively. The same assay was performed toward TEMPOL at<br />

concentration <strong>of</strong> 0.5 mg mL -1 in the same buffer as the control.<br />

Cytotoxicity assay.<br />

The cytotoxicity <strong>of</strong> the block copolymers were assessed using MTT method. Polymer<br />

samples were dispersed in DMEM medium (Dulbecco's Modified Eagle Medium)<br />

supplemented with 10% fetal calf serum (Gibco) in a humidified incubator at 37 <strong>and</strong> 5%<br />

‐ 125 ‐


Chapter 6<br />

CO2 for 24 h. RSC96 cells were seeded at 1.2 ×104 cells/well in 96-well plates <strong>and</strong> cultured<br />

for 24 h. Then the culture media were replaced by the polymer solutions at various diluted<br />

concentrations. After incubation for additional 24 h, 20 μL <strong>of</strong><br />

3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) solution (5 mg/mL)<br />

was added to each well. The plates were return to the incubator for a period <strong>of</strong> 4 h culture.<br />

The supernatant were then carefully removed from each well, <strong>and</strong> 200 μL DMSO was added<br />

to dissolve the MTT formazan crystals. The plate was incubated for an additional 10 min<br />

before the absorbance at 492 nm was recorded by an ELISA microplate reader (Bio-Rad).<br />

The cell viability (%) was calculated according to the following equation:<br />

Cell viability (%) = (Asample / Acontrol) × 100 (1)<br />

Where Asample was the absorbance <strong>of</strong> the polymers extracted solutions treated cells <strong>and</strong><br />

Acontrol was the absorbance <strong>of</strong> the untreated cells. Each experiment was done in triplicate <strong>and</strong><br />

repeated a minimum <strong>of</strong> three times.<br />

References<br />

1. H. Nishide, K. Koshika, K. Oyaizu, Pure Appl. Chem. 2009, 81, 1961.<br />

2. H. Nishide, K. Oyaizu, Science 2008, 319, 737.<br />

3. K. Oyaizu, H. Nishide, Adv. Mater. 2009, 21, 2339.<br />

4. T. Suga, Y. J. Pu, K. Oyaizu, H. Nishide, Bull. Chem. Soc. <strong>of</strong> Jap. 2004, 77, (12),<br />

2203-2204.<br />

5. M. D. Lang, C. C. Chu, J. Polym. Sci. Part A: Polym. Chem. 2001, 39, 4214.<br />

6. G. Jokhadze, M. Machaidze, H. Panosyan, C. C. Chu, R. Katsarava, J. Biomater. Sci.<br />

Polym. Ed. 2007, 18, 411.<br />

7. L. Li, C. C. Chu, J. Biomater. Sci. Polym. Ed. 2009, 20, 341.<br />

8. Z. Zhelev, R. Bakalova, I. Aoki, K. Matsumoto, V. Gadjeva, K. Anzai, I. Kanno, Mol<br />

Pharm 2009, 6, 504.<br />

9. Z. Zhelev, R. Bakalova, I. Aoki, K. Matsumoto, V. Gadjeva, K. Anzai, I. Kanno, Chem.<br />

Commun. 2009, 1, 53.<br />

10. Z. Zhelev, R. Bakalova, I. Aoki, K. I. Matsumoto, V. Gadjeva, K. Anzai, I. Kanno, Mol.<br />

Pharmacol. 2009.<br />

11. C. S. Winalski, S. Shortkr<strong>of</strong>f, R. V. Mulkern, E. Schneider, G. M. Rosen, Magn. Reson.<br />

‐ 126 ‐


Med. 2002, 48, 965.<br />

<strong>Synthesis</strong> <strong>of</strong> Amphiphilic Triblock Copolymers <strong>of</strong> Acrylamide, Lactate, <strong>and</strong> Ethyleneoxide<br />

12. Y. Huang, A. J. Nan, G. M. Rosen, C. S. Winalski, E. Schneider, P. Tsai, H. Gh<strong>and</strong>ehari,<br />

Macromol. Biosci. 2003, 3, 647.<br />

13. Y.Sato, H. Hayashi, M. Okazaki, M. Aso, S. Karasawa, S. Ueki, H. Suemune, N. Koga,<br />

Magn. Reson. Chem. 2008, 46, 1055.<br />

14. H. Hayashi, S. Karasawa, A. Tanaka, K. Odoi, K. Chikama, H. Kuribayashi, N. Koga,<br />

Magn. Reson. Chem. 2009, 47, 201.<br />

15. L. J. Berliner, V. Khramtsov, H. Fujii, T. L. Clanton, Free Radic. Biol. Med. 2001, 30,<br />

489.<br />

16.G. L. He, Y. M. Deng, H. H. Li, P. Kuppusamy, J. L. Zweier, Magn. Reson. Med. 2002, 47,<br />

571.<br />

17. B. P. Soule, F. Hyodo, K. Matsumoto, N. L. Simone, J. A. Cook, M. C. Krishna, J. B.<br />

Mitchell, Free Radic. Biol. Med. 2007, 42, 1632.<br />

18. T. Yoshitomi, D. Miyamoto, Y. Nagasaki, Biomacromolecules 2009, 10, 569.<br />

19. T. Yoshitomi, R. Suzuki, T. Mamiya, H. Matsui, A. Hirayama, Y. Nagasaki, Bioconjug.<br />

Chem. 2009, 20, 1792.<br />

20. K. Takechi, H. Tamura, K. Yamaoka, H. Sakurai, Free Radic. Res. 1997, 26, 483.<br />

21. H. T. Duong, T. U. Ngayen, M. H. Stenzel, Polym. Chem. 2010, 1, 171.<br />

22. J. Fuchs, N. Groth, T. Herrling, G. Zimmer, Free Radic. Biol. Med. 1997, 22, 967.<br />

23. J. Trnka, F. H. Blaikie, R. A. J. Smith, Murphy, M. P., Free Radic. Biol. Med. 2008, 44,<br />

1406.<br />

24. E. Yoshida, T. Tanaka, Colloid Polym. Sci. 2006, 285, 135.<br />

25. E. Yoshida, T. Tanaka, Colloid Polym. Sci. 2008, 286, 827.<br />

26. Y. Mitsukami, M. S. Donovan, A. B. Lowe, C. L. McCormick, Macromolecules 2001, 34,<br />

2248.<br />

27. K. Koshika, Waseda University PhD. Thesis. 2009.<br />

‐ 127 ‐


Chapter 7<br />

Conclusion <strong>and</strong> Future Prospects<br />

7.1. Conclusion<br />

7.2. Future Prospects


Chapter 7<br />

7.1. Conclusion<br />

In this thesis, the author described the catalytic synthesis <strong>and</strong> characterization <strong>of</strong><br />

biodegradable polyesters <strong>and</strong> their radical copolymers. In this chapter, the author has<br />

described the synthesis <strong>of</strong> biodegradable polyesters by using a series <strong>of</strong> newly synthesized<br />

cobalt-Schiff-base complexes catalysts. The catalyst activity, polymerization mechanism,<br />

functionality <strong>of</strong> PLA by incorporating with stable TEMPO radical-substituted polymers <strong>and</strong><br />

important conclusions derived from this study are summarized.<br />

In chapter 1, the author summarized “copolymer <strong>of</strong> epoxide <strong>and</strong> carbon dioxide”,<br />

“polylactide”, “others biodegradable polyesters”, “electroactive polymers” <strong>and</strong> “radical<br />

polymers”, respectively.<br />

In chapter 2, a series <strong>of</strong> Cobalt Schiff-base complexes were investigated as the catalyst for<br />

the alternating copolymerization <strong>of</strong> CO2 <strong>and</strong> rac-PO in the presence <strong>of</strong> Bu4NBr. The<br />

poly(propylene carbonate) (PPC) <strong>and</strong> cyclic propylene carbonate (PC) selectivity <strong>of</strong> the<br />

resultant copolymers were determined by modification <strong>of</strong> the length <strong>of</strong> the diimine bridges<br />

between the two nitrogen atoms in the lig<strong>and</strong>s. The L 1 -Co III -dnp/Bu4NBr catalyst exhibited<br />

the highest activity, PPC/PC selectivity, <strong>and</strong> degree <strong>of</strong> head-to-tail linkages. The<br />

L 2 -Co III -dnp/Bu4NBr catalyst showed slightly lower head-to-tail linkages. For the dimine<br />

bridges containing three-carbon chains between the two nitrogen atoms in the Schiff bases,<br />

the corresponding L 4 -Co III -dnp complex displayed the lowest catalytic activity. Based on the<br />

electrochemical measurements to determine the half-wave potentials <strong>of</strong> the complexes, the<br />

catalytic activity <strong>of</strong> these complexes was compared. The higher stability for the axial group’s<br />

metal–O (phenolate) bond reflected the more negative E1/2 <strong>of</strong> the Co III + e -<br />

‐ 130 ‐<br />

Co II redox<br />

couple for the L-Co III -dnp complexes with the diimine-bridge composed <strong>of</strong> two carbon atoms,<br />

which led to the higher catalytic activity for the copolymerization <strong>of</strong> PO <strong>and</strong> CO2. In the case<br />

<strong>of</strong> the cobalt complexes with more positive Co(II/III) potentials <strong>and</strong> thus with a lower electron<br />

density on the cobalt center, the coordination <strong>of</strong> a labile propagating chain end to the cobalt<br />

center is considered to determine the overall polymerization rate.<br />

In chapter 3, Co(III) complexes with Schiff base lig<strong>and</strong>s <strong>and</strong> Tin(II) alphatates, were used<br />

as catalysts for the ROP <strong>of</strong> LacOCA to produce poly(lactic acid). Intra- <strong>and</strong>/or<br />

inter-transesterification were suggested to coincide with the polymerization, which resulted in<br />

a relatively large polydispersity in molecular weights. The carbonyldioxy group was


‐ 131 ‐<br />

Conclusion <strong>and</strong> Future Prospects<br />

eliminated from the polymerization system. The corresponding PLA showed similar Tg but<br />

lower T5% compared with PLA obtained from lactides, as a result <strong>of</strong> the lower molecular<br />

weights. Al(III) complexes with Schiff base lig<strong>and</strong>s were not catalytically active for the<br />

polymerization <strong>of</strong> LacOCA.<br />

In Chapter 4, novel PTAm-g-PLA copolymers were successfully prepared by a PLA<br />

macromonomer approach. The obtained copolymer contained stable radical units which was<br />

biologically active. The copolymers were amorphous determined from DSC measurement<br />

due to the relatively low PLA content. The copolymer exhibited an improved<br />

biocompatibility compared to PTAm homopolymer.<br />

In chapter 5, Triblock copolymers composed <strong>of</strong> PLA <strong>and</strong> PTAm with well-defined<br />

structure was prepared by combination <strong>of</strong> ROP <strong>and</strong> RAFT methods. Biological evaluation <strong>of</strong><br />

the material showed that the material has good biocompatibility <strong>and</strong> low cytotoxicity.<br />

Electrospinning technology was then adopted to prepare the electro-active porous polymer<br />

matrices with relatively large surface area for biomedical applications.<br />

In chapter 6, two kinds <strong>of</strong> amphiphilic block copolymers containing stable nitroxyl<br />

radicals were synthesized. The RAFT polymerization towards TAP monomer was monitored<br />

to show living features. The block copolymer structures were confirmed by 1 H-NMR <strong>and</strong><br />

GPC characterization. Both <strong>of</strong> the copolymers can self-assemble into micelles in aqueous<br />

solution with nitroxyl radicals in the hydrophobic core. The micelles can produce EPR signal<br />

<strong>and</strong> protect the free radicals from being reduced for a longer time as compared to the free<br />

TEMPOL radicals in the presence <strong>of</strong> high concentration Vc. Additionally, the micelle<br />

structure also endow the materials with low cytotoxicity. Therefore, such micelles with stable<br />

free radicals in the core should be promising to be used as the EPR probe for in vivo imaging<br />

application.


Chapter 7<br />

7.2. Future Prospects<br />

Recent advances <strong>of</strong> the biomedical use <strong>of</strong> polyesters have been paid increasing attention<br />

on dealing with the interaction <strong>and</strong> communication between the materials <strong>and</strong> life systems,<br />

for which the materials are termed as bioactive materials. In order to obtain bioactive<br />

materials, some modifications towards polyesters are usually taken to achieve bioactive<br />

functionalities, including hydrophobicity-hydrophilicity, s<strong>of</strong>tness-hardness, charge-charge<br />

density, stimuli responsiveness, <strong>and</strong> biological functionality. In the past few years, some<br />

efforts have been made by our lab on the functionalization <strong>of</strong> PLA <strong>and</strong> their use in the<br />

biomedical fields. For example, copolymerization <strong>of</strong> lactide monomer with functionalized<br />

cyclic monomers leads to biodegradable polymers with reactive amino or carboxyl groups<br />

which are subsequently used for conjugation with bioactive molecules such as folate, RGD,<br />

sugars, antibody <strong>and</strong> drugs. Then, the use <strong>of</strong> these bio-functionalized materials in, such as<br />

tissue engineering scaffolds or smart drug delivery systems have also been investigated.<br />

Though great efforts have been made, there are still challenges in synthesis <strong>of</strong> biodegradable<br />

polyesters with precise <strong>and</strong> smart properties, such as improved cell adhesions <strong>and</strong><br />

proliferation, precise targeting drug delivery <strong>and</strong> triggered release, promoted cell<br />

internalization <strong>and</strong> so on. Fortunately, by virtue <strong>of</strong> the recent development on the click<br />

chemistry <strong>and</strong> living radical polymerization technique (ATRP, RAFT et al.), it is possible for<br />

us to prepare polyesters with diverse structures <strong>and</strong> architectures, ie. different kinds <strong>of</strong><br />

polymers (natural polymers, synthetic polyesters <strong>and</strong> free radical polymers) can be<br />

conjugated together <strong>and</strong> great amount <strong>of</strong> vinyl monomers can be used to bring additional<br />

properties to the polyesters. Thus, it is possible for us to prepare biomedical polymers with<br />

more pr<strong>of</strong>ound <strong>and</strong> precise functions for specific biomedical use.<br />

A tentative effort has been made in this thesis on the synthesis <strong>of</strong> the biodegradable<br />

copolymers <strong>of</strong> PLA <strong>and</strong> TEMPO-contained PTAm through combination <strong>of</strong> ring-opening<br />

polymerization <strong>and</strong> RAFT living radical polymerization. The biocompatibility <strong>and</strong> potential<br />

biomedical use were also evaluated. However, it is reasonably attractive to make step further<br />

in the application <strong>of</strong> the TEMPO-based radical polymers in biomedical fields based on the<br />

unique electronic, magnetic <strong>and</strong> biological properties <strong>of</strong> the nitroxyl radicals. Tasks are still<br />

remained in the further investigation <strong>of</strong> these materials interaction <strong>and</strong> communication with<br />

cells through electronic transport or nitroxyl radical triggered signal pathway. Materials in the<br />

‐ 132 ‐


‐ 133 ‐<br />

Conclusion <strong>and</strong> Future Prospects<br />

forms <strong>of</strong> solid film, nano/micro fibrous film, mesoporous scaffold or drug-eluting stent<br />

coating should be constructed <strong>and</strong> evaluated in the field including induced cell differentiation<br />

<strong>and</strong> proliferation, cell adhension or retardation, antibacterials <strong>and</strong> so on. Additionally, the MR<br />

<strong>and</strong> ESR properties <strong>of</strong> the TMEPO radical also endow the copolymers with promising use as<br />

the “s<strong>of</strong>t” bioimaging probes in vivo. As shown in our chapter 6, the core-shell micellar<br />

structure was able to protect the TEMPO from being reduced by reductant in a relatively<br />

longer time as compared to the free TEMPO molecules <strong>and</strong> reduced cytotoxicity was also<br />

observed. However, the stability <strong>of</strong> the micelles in the blood stream is still the problem<br />

encountered. The further improved stability <strong>and</strong> MR or EPR signal intensity are now under<br />

investigation. One <strong>of</strong> the strategies is to prepare core cross-linked nanogels, which may<br />

significantly improve the stability <strong>of</strong> the nanoparticles. And the nanogels can also be produced<br />

with stimuli responsive monomers, so that the obtained nano-probe could generate different<br />

signals in response to diverse environment for applications in probing various inner parts in<br />

vivo.


List <strong>of</strong> Publications<br />

1. Xiuli Zhuang, Kenichi Oyaizu, Yongsheng Niu, Kenichiroh Koshika, Xuesi Chen, Hiroyuki<br />

Nishide, “<strong>Synthesis</strong> <strong>and</strong> Electrochemistry <strong>of</strong> Schiff Base Cobalt(III) Complexes <strong>and</strong> Their<br />

<strong>Catalytic</strong> Activity for Copolymerization <strong>of</strong> Epoxide <strong>and</strong> Carbon Dioxide” Macromol. Chem.<br />

Phys., 210, 669-676 (2009).<br />

2. Xiuli Zhuang, Han Zhang, Natsuru Chikushi, Changwen Zhao, Kenichi Oyaizu, Xuesi Chen,<br />

Hiroyuki Nishide, “<strong>Biodegradable</strong> <strong>and</strong> Electroactive TEMPO-Substituted Acrylamide/Lactide<br />

Copolymer” Macromol. Biosci., (2010) in press.<br />

3. Xiuli Zhuang, Chunsheng Xiao, Kenichi Oyaizu, Natsuru Chikushi, Xuesi Chen, Hiroyuki<br />

Nishide, “<strong>Synthesis</strong> <strong>of</strong> Amphiphilic Block Copolymers bearing Stable Nitroxyl Radicals” J.<br />

Polym. Sci., Part A: Polym. Chem.,, (2010), submitted.<br />

4. Xiuli Zhuang, Haiyang Yu, Zhaohui Tang, Kenichi Oyaizu, Hiroyuki Nishide, Xuesi Chen,<br />

“Polymerization <strong>of</strong> Lactic o-Carboxylic Anhydride Using Organometallic Catalysts” Cn. J. Polym.<br />

Sci., (2010), in press.<br />

5. Xiuli Zhuang, Han Zhang, Yu Wang, Changwen Zhao, Kenichi Oyaizu, Xuesi Chen, Hiroyuki<br />

Nishide, “<strong>Synthesis</strong> <strong>of</strong> <strong>Biodegradable</strong> <strong>and</strong> TEMPO-Substitute Triblock Copolymers by<br />

Combination <strong>of</strong> ROP <strong>and</strong> RAFT Polymerization Methods” Polymer, (2010), submitted.<br />

6. Changwen Zhao, Xiuli Zhuang, Pan He, Chunsheng Xiao, Chaoliang He, Jingru Sun, Xuesi Chen,<br />

<strong>Synthesis</strong> <strong>of</strong> <strong>Biodegradable</strong> Thermo- <strong>and</strong> pH-Responsive Hydrogels for Controlled Drug Release,<br />

Polymer, 50, 4308-4316 (2009)<br />

7. Zhaopei Guo, Yanhui Li, Huayu Tian, Zhuang Xiuli, Xuesi Chen, Xiabin Jing, Self-Assembly <strong>of</strong><br />

Hyperbranched Multiarmed PEG-PEI-PLys(Z) Copolymer into Micelles, Rings, <strong>and</strong> Vesicles,<br />

Langmuir, 25, 9690-9696 (2009).<br />

8. Youliang Hong, Yanan Li, Zhuang Xiuli, Xuesi Chen, Xiabin Jing, Electrospinning <strong>of</strong><br />

Multicomponent Ultrathin Fibrous Nonwovens for Semi-Occlusive Wound Dressings, J. Biomed.<br />

Mater. Res. A, 89A, 345-354 (2009).<br />

9. Jun Hu, Lihong Huang, Le Lang, Yadong Liu, Zhuang Xiuli, Xuesi Chen,Yen Wei, Xiabin Jing,<br />

The Study <strong>of</strong> Electroactive Block Copolymer Containing Aniline Pentamer Isolated from<br />

Different Solvents, J. Polym. Sci., Part A: Polym. Chem., 47, 1298-1307 (2009).<br />

10. Yongsheng Niu, Hongchun Li, Xuesi Chen, Wanxi Zhang, Zhuang Xiuli, Xiabin Jing, Alternating<br />

Copolymerization <strong>of</strong> Carbon Dioxide <strong>and</strong> Propylene Oxide Catalyzed by Cobalt Schiff Base<br />

Complex, Macromol. Chem. Phys., 210, 1224-1229 (2009).<br />

11. Yongsheng Niu, Wanxi Zhang, Hongchun Li, Xuesi Chen, Jingru Sun, Zhuang Xiuli, Xiabin Jing<br />

Carbon Dioxide/Propylene Oxide Coupling Reaction Catalyzed by Chromium Salen Complexes,<br />

Polymer, 50, 441-446 (2009).<br />

12. Chunsheng Xiao, Huayu Tian, Zhuang Xiuli, Xuesi Chen, Xiabin Jing, Recent Developments in<br />

Intelligent Biomedical Polymers, Sci. China Ser B., 52, 117-130 (2009).


13. Xuan Pang, Xuesi Chen, Xiuli Zhuang, Xiabin Jing, Crown-Like Macrocycle Zinc Complex<br />

Derived from β-Diketone Lig<strong>and</strong> for the Polymerization <strong>of</strong> rac-Lactide, J. Polym. Sci., Part A:<br />

Polym. Chem., 46, 643–649 (2008).<br />

14. Lihong Huang, Xiuli Zhuang, Jun Hu, Le Lang, Peibiao Zhang, Yu Wang, Xuesi Chen, Yen Wei,<br />

<strong>and</strong> Xiabin Jing, <strong>Synthesis</strong> <strong>of</strong> <strong>Biodegradable</strong> <strong>and</strong> Electroactive Multiblock Polylactide <strong>and</strong><br />

Aniline Pentamer Copolymer for Tissue Engineering Applications, Biomacromolecules, 9,<br />

850-858 (2008).<br />

15. Jun Hu, Lihong Huang, Xiuli Zhuang, Xuesi Chen, Yen Wei, Xiabing Jing, New Oxidation State<br />

<strong>of</strong> Aniline Pentamer Observed in Water-Soluble Electroactive Oligoaniline-Chitosan Polymer, J.<br />

Polym. Sci., Part A: Polym. Chem.,46, 1124–1135 (2008).<br />

16. Lihong Huang, Jun Hu, Le Lang, Xiuli Zhuang, Xuesi Chen, Yen Wei, Xiabin Jing,<br />

“S<strong>and</strong>glass’’-Shaped Self-Assembly <strong>of</strong> Coil–rod–coil Triblock Copolymer Containing Rigid<br />

Aniline-Pentamer, Macromol. Rapid Commun., 29, 1242–1247 (2008).<br />

17. Li Chen, Zhigang Xie, Xiuli Zhuang, Xuesi Chen, Xiabin Jing, Controlled Release <strong>of</strong> Urea<br />

Encapsulated by Starch-g-poly(L-lactide), Carbohyd. Polym., 72, 342–348 (2008).<br />

18. Chaoliang He, Changwen Zhao, Xinhua Guo, Zhaojun Guo, Xuesi Chen, Xiuli Zhuang, Shuying<br />

Liu, Xiabin Jing, Novel Temperature- <strong>and</strong> pH-Responsive Graft Copolymers Composed <strong>of</strong><br />

Poly(L-glutamic acid) <strong>and</strong> Poly(N-isopropylacrylamide), J. Polym. Sci., Part A: Polym. Chem., 46,<br />

4140–4150 (2008).<br />

19. Chaoliang He, Changwen Zhao, Xuesi Chen, Zhaojun Guo, Xiuli Zhuang, Xiabin Jing, Novel<br />

pH- <strong>and</strong> Temperature-Responsive Block Copolymers with Tunable pH-Responsive Range,<br />

Macromol. Rapid. Commun., 29, 490–497 (2008).<br />

20. Li Chen , Xiuli Zhuang , Ge Sun , Xuesi Chen , XiabinJing, An Approach to Synthesize<br />

Poly(ethylene glycol)-b-poly(epsilon-caprolactone) with Terminal Amino Group via Schiff's Base<br />

as An Initiator, Cn. J. Polym. Sci., 26, 455-463 (2008).<br />

21. Xueyu Qiu, Yadong Han, Xiuli Zhuang, Xuesi Chen, Yuesheng Li, Xiabin Jing, Preparation <strong>of</strong><br />

Nano-Hydroxyapatite/Poly(L-lactide) Biocomposite Microspheres, J. Nanopaticle Res., 9,<br />

901-908, (2007).<br />

22. Yongsheng Niu, Wanxi Zhang, Xuan Pang, Xuesi Chen, Xiuli Zhuang, Xiabin Jing, Alternating<br />

Copolymerization <strong>of</strong> Carbon Dioxide <strong>and</strong> Propylene Oxide Catalyzed by<br />

(R,R)-SalenCo(III)-(2,4-dinitrophenoxy) <strong>and</strong> Lewis-Basic Cocatalyst, J. Polym. Sci., Part A:<br />

Polym. Chem., 45, 5050-5056 ( 2007).<br />

23. Junli Hu, Yadong Han, Xiuli Zhuang, Xuesi Chen, Yuesheng Li, Xiabin Jing, Self-Assembly <strong>of</strong> a<br />

Polymer Pair Through Poly(lactide) Stereocomplexation, Nanotechnology, 18, 185607 (2007).<br />

24. Hongzhi Du, Xuan Pang, Haiyang Yu, Xiuli Zhuang, Xuesi Chen, Dongmei Cui, Xianhong Wang,<br />

Xiabin Jing, Polymerization <strong>of</strong> rac-Lactide Using Schiff Base Aluminum Catalysts: Structure,<br />

Activity, <strong>and</strong> Stereoselectivity, Macromolecules, 40, 1904-1913 (2007).<br />

25. Chaoliang He, Jingru Sun, Ting Zhao, Zhongkui Hong, Xiuli Zhuang, Xuesi Chen, Xiabin Jing,


Formation <strong>of</strong> a Unique Crystal Morphology for the Poly (ethylene glycol)-poly( -caprolactone)<br />

Diblock Copolymer, Biomacromolecules, 7, 252-258 (2006).<br />

26. Xuan Pang, Hongzhi Du, Xuesi Chen, Xiuli Zhuang, Dongmei Cui, Xiabin Jing, Aluminum<br />

Schiff Base Catalysts Derived from β-Diketone for the Stereoselective Polymerization <strong>of</strong><br />

Racemic Lactides, J. Polym. Sci., Part A: Polym. Chem., 43, 6605-6612 (2005).<br />

27. Xiuling Xu, Xiuli Zhuang, Xuesi Chen, Xinri Wang, Lixin Yang, Xiabin Jing, Preparation <strong>of</strong><br />

Core-Sheath Composite Nan<strong>of</strong>ibers by Emulsion Electrospinning, Macromol. Rapid. Commun.,<br />

27 (19), 1637-1642 (2006).<br />

28. Suobo Zhang, Xiuli Zhuang, Jifeng Zhang, Wenqi Chen, Juzheng Liu, <strong>Synthesis</strong> <strong>and</strong> Crystal<br />

Structure <strong>of</strong> [(2,4-C7H11)2LnC≡CC6H5]2 (Ln=Gd, Er), J. Organomet. Chem, 584, 135 (1999).<br />

29. Jizhu Jin, Xiuli Zhuang, Zhongsheng Jin, Wenqi Chen, Syntheses <strong>and</strong> Crystal Structure <strong>of</strong><br />

(C8H8)Nd(C5H9C5H4)(THF)2 <strong>and</strong> [(C8H8)Gd(C5H9C5H4)(THF)2], J. Organomet. Chem., 490, C8<br />

(1995).<br />

30. Jusong Xia, Xiuli Zhuang, Zhongsheng Jin, Wenqi Chen, <strong>Synthesis</strong> <strong>and</strong> Crystal Structure <strong>of</strong><br />

[(C8H8)3(C6H5CH2C5H4)Nd2K(THF)3], Polyhedron, 15 3399-3403 (1996).<br />

31. Xiuli Zhuang, Suobo Zhang, Ninghai Hu, Wenqi Chen, Syntheses <strong>and</strong> Crystal Structure <strong>of</strong><br />

Bis(tetrahydr<strong>of</strong>urfurylcyclopentadieny) Ytterbium Chloride Chem. Res. Cn. Univ., 10, 254<br />

(1994).<br />

32. Suobo Zhang, Xiuli Zhuang, Gecheng Wei Wenqi Chen, Syntheses <strong>of</strong><br />

(C4H7OCH2C5H4)2LnCl(Ln=Nd,Gd,Dy,Yb) <strong>and</strong> Crystal Structure <strong>of</strong> (C4H7OCH2C5H4)2DyCl,<br />

Polyhedron, 13, 2867 (1994).


Acknowledgments<br />

The present thesis is the collection <strong>of</strong> the studies which have been carried out under the<br />

direction <strong>of</strong> Pr<strong>of</strong>. Dr. Hiroyuki Nishide, Department <strong>of</strong> Applied Chemistry in Waseda<br />

University, during the 2008-2010. The author expresses the greatest acknowledgement to<br />

Pr<strong>of</strong>. Dr. Hiroyuki Nishide for his invaluable suggestion, discussion, <strong>and</strong> continuous<br />

encouragement throughout this work.<br />

The author also expresses her sincere gratitude to Pr<strong>of</strong>. Dr. Xuesi Chen <strong>of</strong> Changchun<br />

Institute <strong>of</strong> Applied Chemistry, Chinese Academy <strong>of</strong> Sciences, <strong>and</strong> Assoc. Pr<strong>of</strong>. Dr.<br />

Kenichi Oyaizu <strong>of</strong> Waseda University for their valuable advice <strong>and</strong> encouragement.<br />

The author wishes to thank Pr<strong>of</strong>. Dr. Kuniki Kino <strong>of</strong> Waseda University <strong>and</strong> Timothy E.<br />

Long <strong>of</strong> Virginia Polytechnic Institute <strong>and</strong> State University for their efforts as members<br />

on judging committee for the doctoral thesis.<br />

The author expresses the special thanks to all active <strong>and</strong> energetic collaborators, Dr.<br />

Kenichiroh Koshika, Dr. Changwen Zhao, Mr. Han Zhang, Mr. Chunsheng Xiao, Dr.<br />

Yongsheng Niu, Mr. Natsuru Chikushi, Mr. Haiyang Yu, Mr. Naoki Sano for their strong<br />

assistance in the experimental work.<br />

The author deeply thanks Dr. Xuan Pang, Dr. Zhaohui Tang, Dr. Takeo Suga, Ms Shiori<br />

Furuyama, Ms Yu Zhang, Dr. Wihatmoko Waskitaji, Mr. Won-Song Choi, Dr. Fumiaki<br />

Kato, Dr. Takeshi Ibe, Mr. Satoshi Nakajima, <strong>and</strong> all members in Pr<strong>of</strong>. Nishide’s<br />

laboratory in Waseda University <strong>and</strong> Pr<strong>of</strong>. Chen’s laboratory in Changchun Institute <strong>of</strong><br />

Applied Chemistry for their fruitful discussion <strong>and</strong> kind assistance.<br />

Finally, the author expresses her deepest gratitude heartily to her parents, her husb<strong>and</strong>,<br />

<strong>and</strong> her son for their heartfelt supports.<br />

July, 2010<br />

Xiuli Zhuang

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!