You are on page 1of 24

Review

Recent Advances in Reactive Extrusion Processing of Biodegradable Polymer-Based Compositions


Jean-Marie Raquez, Ramani Narayan, Philippe Dubois*

This review reports on recent advances in the design of biodegradable polymers built from petroleum and renewable resources using reactive extrusion processing. Reactive extrusion represents a unique tool to manufacture biodegradable polymers upon different types of reactive modication in a cost-effective way. Partially based on our ongoing research, ringopening polymerization of biodegradable polyesters will be approached as well as the chemical modication of biodegradable polymers, particularly natural polymers. The development of environmentally friendly polymer blends as well as (nano)composites from natural polymers, including natural bers and nanoclays, through reactive extrusion, as an efcient way to improve the interfacial adhesion between these components, will be also discussed.

Introduction
Reactive extrusion is an attractive route for cost-effective polymer processing, which enhances the commercial viability and cost-competitiveness of these materials, in

J.-M. Raquez, P. Dubois Center of Innovation and Research in Materials & Polymers (CIRMAP), Laboratory of Polymeric and Composite Materials, Materia Nova & University of Mons-Hainaut, Place du Parc 20, B-7000 Mons, Belgium E-mail: philippe.dubois@umh.ac.be R. Narayan Department of Chemical Engineering & Material Science, 2527 Engineering Building, Michigan State University, East Lansing, MI-48824, USA J.-M. Raquez ` Departement Technologie des Polymeres et Composites, Ecole des Mines de Douai, Rue C. Bourseul 941 B.P. 10838, 59508 Douai Cedex, France
Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

order to carry out melt-blending, but also various chemical reactions including polymerization, grafting, branching and functionalization as well.[1,2] In addition, reactive extrusion processes involve introducing reactive agents at optimum points in the reaction sequence, homogenizing the ingredients, and allowing sufcient time for the completion of the reactions. In a typical reactive extrusion process, the reactants are fed into the extruder, usually through a feed hopper. However, various liquid or gaseous reactants can be introduced at specic points in the reaction sequence by using injection along the extruder barrel. The reactive mixture is conveyed through the extruder, and the reaction is driven to the desired degree of completion. At this point, and after removing any volatile by-products, the molten polymeric product is pumped out through a die and subsequently quenched, solidied, and pelletized. Therefore, production and processing can be integrated in one-stage processing.[3,4] The recent environmental regulations, societal concerns and growing environmental understanding throughout

DOI: 10.1002/mame.200700395

447

J.-M. Raquez, R. Narayan, P. Dubois

Jean-Marie Raquez was born in Paris (France) in 1977. He graduated in Chemistry from the University of Mons-Hainaut (Belgium) in 1999. He received his PhD degree in Polymer Chemistry under the supervision of Professor Philippe Dubois (University of Mons-Hainaut) on the topic dealing with the controlled synthesis of a biodegradable poly(ester-alt-ether), poly(1,4dioxan-2-one), through ring-opening polymerization. After a postdoctoral stay with Professor Ramani Narayan (Michigan State University, USA), he moved back to University of Mons-Hainaut as assistant research. In 2007 he was appointed Associated Professor at Ecole des Mines de Douai in France. His research work focuses on the chemical modication and synthesis of polymer-based (nano)composites issued from renewable resource. Ramani Narayan is University Distinguished Professor at Michigan State University in the Department of Chemical Engineering & Materials Science. He has 115 refereed publications in leading journals to his credit, 18 patents, edited three books and one expert dossier in the area of bio-based polymeric materials. Under his supervision, 20 students have obtained their Masters degree, ten students their Ph.D. degrees and six are working towards their Ph.D. He has major research programs with industry and serves as consultant for several companies. He has won several awards: Awarded the University Distinguished Professor title in 2007 highest and very selective honor that can be bestowed on a faculty member by the university. Those selected for the title have been recognized nationally and internationally for the importance of their teaching, research and public service achievements; Governors (State of Michigan) University Award for commercialization excellence; University Distinguished Faculty Award, 2006; Withrow Distinguished Scholar award, 2005 awarded to one faculty in the MSU College of Engineering based on exemplary research accomplishments, national & international recognition; Fulbright Distinguished Lectureship Chair in Science & Technology Management & Commercialization (University of Lisbon; Portugal); the William N. Findley Award for signicant contributions to the application of new technologies within the scope of ASTM Committee D20 on Plastic; Award of Excellence from ASTM committee D 20 on Plastics for exemplary technical contributions, sustained participation, and valued leadership; 2006. The James Hammer Memorial Lifetime Achievement Award, 2006 for outstanding leadership, and research accomplishments in the eld of Degradable Polymers from the BioEnvironmental Polymer Society (BEPS). Research and Commercialization Award sponsored by ICI Americas, Inc. & the National Corn Growers Association. He is on the Board of Directors of ASTM International-a premier international standards setting organization. He chairs the ASTM committee on Environmentally Degradable Plastics and Biobased Products (D20.96) and the Plastics Terminology committee D20.92. He is also the technical expert for the USA on ISO TC 61 on Plastics specically for Terminology, and Biodegradable plastics. Dr. Narayan also chairs the scientic committee of Biodegradable Products Institute (BPI), North Americaa biodegradable and biobased plastics trade industry organization (www.bpiworld.org). He is a successful entrepreneur having been responsible for commercializing several technologies. Philippe Dubois graduated in Chemical Sciences from the Facultes Universitaires Notre-Dame de la Paix (FUNDP, Namur, ` Belgium) in 1987. He received his PhD degree in Chemical Sciences from University of Liege (ULg) in 1991. In the same year, he worked as a postdoctoral fellow for Dow Chemical (Terneuzen, Holland) and the Laboratory of Macromolecular Chemistry and Organic Catalysis directed by Prof. Ph. Teyssie at ULg. Then he joined the National Fund for Scientic Research (FNRS) at ULg till 1997. In 1994, he worked as visiting scientist at the Chemical Research Engineering Department of the Michigan State University (MSU). In Oct. 1997, he moved to University of Mons-Hainaut (UMH) where he obtained the chair of macromolecular chemistry and created/directed the Laboratory of Polymeric and Composite Materials (now ca. 35 people). He has co-authored over 300 publications in international journals, 180 personal communications at conferences and is co-inventor of 40 patents. He co-edited 6 books. He is full professor at UMH and invited professor at FUNDP, MSU, and Faculte Polytechnique de Mons (FPMs). He is Scientic Director at the Materia Nova Research Center in Mons and Director of the Center of Innovation and Research in Materials & Polymers CIRMAP (with ca. 80 members). He is currently Past-President of the Belgian Royal Chemical Society (he was the President in 2006/7).

the world have triggered renewed efforts in plastic industry to develop new products and processes compatible with our environment.[510] The design of biodegradable plastics is an appropriately eco-efciency approach to enhance the environmental quality for many products. Biodegradable plastics can be converted into useful and friendly environmental products to minimize the waste disposed in landlls. Different markets are found in the realm of the biodegradable polymers, including packaging (trash bags, wrappings, loose-ll foam, food containers and laminated papers), disposable non-woven (engineered fabrics), hygiene products (diaper back sheets and cotton swabs), consumer goods (fast-food tableware, containers,
Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

egg cartons, razor handles and toys) and agricultural tools (mulch lms and planters). Unfortunately, the utilization of biodegradable polymers as bulk commodity materials is still restricted to few applications because of the strong cost-competition with cheaper petroleum-based polymers, and their limited thermo-mechanical properties.[8,11] To permit their commercial scale-up, an appropriate, inexpensive and easy process to manufacture biodegradable polymers is highly desirable for their commercial viability and costcompetitiveness. Besides, while maintaining their overall biodegradability, melt-blending through a reactive way these biodegradable polymers with inorganic/organic

448

DOI: 10.1002/mame.200700395

Recent Advances in Reactive Extrusion Processing . . .

llers as well as other biodegradable polymeric materials may represent an additional way of reducing the overall cost for the resulting materials, but also to modify their thermo-mechanical properties and (bio)degradation rates efciently. Developing such biodegradable polymeric melt-blends/composites with satisfactory overall thermomechanical behavior however requires the ability to control interfacial energy, to generate dispersed phases of limited size and strong interfacial adhesion, and to improve the stress transfer between the component phases.[12] This can be effectively completed using proper interface compatibilization between these different components during their reactive processing.[13] Undoubtedly, reactive extrusion (coined REX) serves on all these issues in the manufacture of high-performance and inexpensive biodegradable polymeric materials using a one-stage continuous reactive processing. Hence, this review aims at highlighting the recent developments of novel biodegradable polymeric materials using REX as an efcient processing technique, partially based on our ongoing research over the past few years. The production of biodegradable polymers through REX will be described, as well as the reactive modication and melt blending of biodegradable polymers in the preparation of useful and environmentally friendly products.

Reactive Extrusion Processing


Reactions that previously required heavy equipments, particularly with batch operations, can be completed in a more efcient continuous way through REX. Extruders have been used to resolve heat and mass transfer problems that arise when dramatic viscosity of the reaction medium (when monomer is converted to polymer) increases within a magnitude order of 105 in batch polymerization processes.[14,15] In a batch reactor, as the polymerization proceeds, the viscosity increases and after a certain point the material becomes unmanageable in terms of mixing and heat transfer. In this respect, REX shows to be a promising technique for polymer processing. The ability of these extruders to create new thin surface layers continuously can increase the degree of mixing

and minimize temperature gradients within the polymer being processed. Another factor that can be controlled via operating conditions and geometrical specications of screw extruders is the residence time in the system. Generally, the residence time is substantially lower as compared to that required in a batch reactor for the same reaction; reducing therefore long exposure to high temperatures that can cause polymer degradation. The ability of an extruder to handle materials of high viscosities without any solvents results in a dramatic cost reduction in raw materials and in solvent recovery equipments, and polymers produced are in a ready-to-use form. Finally, for polymer modication reactions, reactive extrusion processes offer natural means for polymerization, chemical cross-linking and grafting via the inherent capability of stage feeding of reactive agents (Figure 1). This way, one can tailor-make specialty polymers that are uneconomical to produce in large scale operations. The design of extruders for REX applications involves the manipulation and integration of information and knowledge from several distinct areas. Extruders for reactive processing must deal with the continuously changing nature of the reactive melt. Mixing phenomena are considerably delayed when the reactive mixture is highly viscous, leading to important gradient in chemical composition and temperature that affect the product quality. Mixing is an important factor when using an extruder as a reactor. Especially longitudinal mixing must

Figure 1. Schematic representation of a reactive extrusion process including a typical screw for polymerization of CL under inert atmosphere.

Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.mme-journal.de

449

J.-M. Raquez, R. Narayan, P. Dubois

be given signicant consideration for radical reactions since back-mixing can inuence the residence time distribution, that is the course of the reaction, and can affect the molecular weight distribution (MWD) of the nal product.[20] Although both single and twin-screw extruder congurations are used in REX processes, twin-screw ones are increasingly being favored over the single-screw ones. The main reasons for this are the extended control of residence time distribution and mixing, but also their superior heat and mass transfer capabilities. The single-screw extruder is currently utilized for simpler jobs like melting, plasticizing, and discharging melt for the production of lms, pipes, proles, etc. The twin-screw extruders, according to their specic characteristics, can tackle the more complex tasks such as homogenizing, dispersing pigments and additives, alloying, reactive compounding, concentrating, devolatilizing, polymerizing, etc. The major difference between single- and twin-screw extruders is the conveying mechanism. Although in single-screw machines, it depends on frictional forces in the solids conveying zone and viscous forces in the melt-pumping zone, in twinscrew extruders it is largely dependent on the screw geometrical conguration, and it is of a positive displacement character. The relative merits and the performance of twin-screw extruders have been appraised. Depending on the direction of rotation of the two screws, twin-screw extruders can be distinguished in co-rotating and counter rotating machines. In simple terms, it can be said that co-rotating screws have a radial and counter-rotating have an axial shearing and plasticizing effect. Although each type of twin-screw extruder has a certain uniqueness regarding ingredients, type of reaction, and polymer produced, and although no machine design (counter- or co-rotating) provides the complete solution, co-rotating intermeshing twin-screw extruders have been found to be suitable for many continuous REX processes. Compared with the counter-rotating twin-screw extruders where additional radial forces are present, the two screws for co-rotating intermeshing twin-screw extruders are set side by side with minimum clearance between them. The crest of one screw completely wipes the ights and the root of the other one. This self-wiping feature eliminates dead regions where material can stagnate during processing. Due to the co-rotating design of the screws, high speeds, thus strong shearing forces, and high outputs can be obtained, enabling the material to be transferred from one screw to the other under a constant mixing. In addition, the intensive and constant surface renewal creates favorable degassing conditions. The modular design and assembly arrangement of the screw and barrel sections of twin-screw machines, along with the use of special feeding and venting ports provide adequate exibility for specic reactive extrusion tasks.
Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

The sequence of screw elements is of prime importance for the control of the lling ratio inside the extruder. Three types of elements are widely used in co-rotating intermeshing twin-screw extruders: kneading blocks, mixing gears, and conveying screw elements. Kneading blocks are primarily used for the dispersion of partially melted polymer particles and solid additives or reactive agents into the melt. Mixing gears are superior to kneading blocks for the thorough distribution of nely divided particles, for achieving isothermal conditions at a given location in the barrel, and for the homogenization of two or more feed streams. Conveying elements are the ones that move the material through the extruder. The arrangement of the screw elements is what determines the residence time distribution in twin-screw extruders. The maximum practical residence time capability of a typical co-rotating intermeshing 45:1 Length/Diameter twin-screw extruder is about 7 min, at low screw speeds and with shortpitch elements. Designs with even longer residence times (1075 min) are also available. Recently, microcompounding using reactive extrusion technology has emerged as a efcient way of handling small amount of materials, leading to the preparation of expensive materials such as (nano)composites, biopolymers or pharmaceuticals at limited cost.[21] According to its unique characteristic features, the reactive extrusion processing technology has provided different types of chemical reactions:[14,2031] a. Free radical, anionic, cationic, condensation, and coordination polymerizations of monomers or oligomers to high molecular weight polymers. b. Controlled degradation and cross-linking of polymers by means of a free radical initiator for preparing a product with controlled molecular weight distribution. c. Functionalization of commodity polymers for producing materials to be used in grafting applications. d. Polymer modication by grafting of monomers or mixture of monomers onto the backbone of existing polymers for improving various properties of the starting materials. Free radical initiators and ionizing radiation can be used to initiate the grafting reactions. e. Interchain copolymer formation: Usually, this type of reaction involves combination of reactive groups from several polymers to form a graft copolymer. f. Coupling reactions that involve reaction of a homopolymer with a polyfunctional coupling agent/ller in the preparation of high-performance products. Many authors have largely reviewed different synthetic technology for the preparation of biodegradable polymers, as well as their types of applications ranging from daily applications to biomedical ones.[68,32] Within the scope of this review, we will focus on the use of reactive extrusion

450

DOI: 10.1002/mame.200700395

Recent Advances in Reactive Extrusion Processing . . .

in the design of biodegradable polymeric materials with useful properties. First, the reactive extrusion synthesis of biodegradable polymers, particularly aliphatic polyesters obtained through ring-opening polymerization (ROP) of cyclic (di)esters will be discussed. These synthetic biodegradable polymers have attracted much attention owing to their physico-chemical properties that can be readily tailored for more specic applications. Besides, chemical modication and reactive melt blending of synthetic biodegradable polymers, as well as those derived from renewable resources will be discussed, again carried out through REX for the manufacture of useful and fully biodegradable products. Finally, a special attention will be paid to biodegradable polymeric composites, and particularly the (nano)composites. The reinforcement of biodegradable polymers using (nano)llers, particularly layered silicates, have recently emerged as high-performance materials having interesting mechanical and barrier properties achieved at low ller content (less than 5 wt.-%).[33]

Reactive Extrusion (Co)polymerization of Cyclic (Di)esters


Among the synthetic biodegradable polymers, aliphatic polyesters such as poly(e-caprolactone) (PCL) and polylactides (PLAs) have drawn a lot of interest from both the academic and industrial media, whose potential applications cover such widely different elds as packaging for industrial products, mulching lms in agriculture, bioresorbable materials for hard tissue replacement and controlled drug delivery devices.[6,34] PCL has been thoroughly investigated because of the possibility of blending this aliphatic polyester with a number of miscible commercial polymers such as PVC and bisphenol A polycarbonate. PCL is a highly hydrophobic, biodegradable and semi-crystalline polyester with melting and glass transition temperatures of ca. 60 8C and 60 8C, respectively.[3537] The homopolymer of L-lactide (or D-lactide), poly(L-lactide) (or poly(D-lactide)), is semi-crystalline with melting and glass transition temperatures of ca. 175 8C and 60 8C, respectively. Poly(L-lactide) (PLA) is biocompatible, and degrades by hydrolytic scission to lactic acid, which is a natural intermediate in the carbohydrate metabolism. High tensile strength and low ultimate elongation make that poly(L-lactide) is rather used for producing porous scaffolds and load-bearing applications such as in orthopedic xations and sutures.[38,39] Interestingly enough, poly(1,4-dioxan-2-one) (PPDX) appears to be an attractive candidate as a biodegradable substitute for commodity polymers. This aliphatic poly(ester-alt-ether) copolymer offers a good compromise between its processing temperature and the service
Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

temperature range with a melting and glass transition temperatures of 110 8C and 10 8C, respectively.[4044] Aliphatic polyesters can be produced by two different synthetic pathways: step polycondensation of a,v-hydroxyacids and ROP of cyclic (di)esters. However, the traditional polycondensation usually requires high temperatures, long reaction time and a continuous removal of water to recover quite low molecular weight polymers with poor mechanical properties nally. In contrast, ROP provides a direct and easy access to the corresponding high molecular weight polyester within a few minutes. Both aluminum mono- and trialkoxides have shown to be very effective initiators[4551] to promote the ROP of various (di)lactones such as e-caprolactone (CL) with high selectivity (restricted occurrence of termination and transfer reactions). This allows the preparation of high molecular weight polyesters and a huge range of novel macromolecular architectures both in solution and in bulk (i.e., in absence of any solvent). Next to the aforementioned aluminum derivatives, tin (II, IV) alkoxides and carboxylates (the latter coupled with alcohols as co-initiators), and more particularly tin(II) bis(2-ethylhexanoate) (Sn(Oct)2)[52] have been widely used. Such commercial catalysts can be more readily handled (i.e., do not require high vacuum equipment), and are relatively easy to purify (at least down to ca. 2 mol% of proton containing impurities) by distillation for semi-quantitative synthetic work,[49] Furthermore, Sn(Oct)2 has been approved as a food additive by FDA. The most advocated mechanism[50,51] involves a direct catalytic action of Sn(Oct)2. Actually, Sn(Oct)2 has been rst proposed to activate the monomer forming a donor-acceptor complex, which further participates directly in the propagation step. Sn(Oct)2 is liberated in every act of propagation. It follows from this mechanism that Sn(II) atoms are not covalently bound to the polymer at any stage of polymerization. Recently, Penczek et al. have proposed another more likely scheme proceeding via the active chain-end mechanism. This last one involves the in-situ formation of Sn-alkoxide bonds at the chain-ends, as observed by MALDI-TOF and fully conrmed by kinetic studies.[49,52] Thus, through a rapid exchange equilibrium, Sn(Oct)2, and most probably any other covalent metal carboxylates, are rst converted by reaction with protic compounds (ROH) into tin (or other metal) alkoxides as active centers for polymerization (Figure 2). The polymerization involves a coordinationinsertion mechanism similarly to the previously discussed mechanism for covalent metal alkoxides and dialkylaluminum alkoxides (Figure 3). It is worth noting that we have recently and succinctly reported on ROP of cyclic (di)esters, and to some extent, to the chemical modication of biodegradable aliphatic polylactones through reactive extrusion processing.[53] However, the scope of that mini-review was restricted

www.mme-journal.de

451

J.-M. Raquez, R. Narayan, P. Dubois

Figure 2. Proposed activation mechanism for catalyzed ROP of e-caprolactone promoted by Sn(Oct)2.

only to our own expertise, without referring to the most relevant results obtained elsewhere. By contrast, the following section will summarize in a more systematic way the main advances in REX obtained by the different research groups that are involved in the eld. Reactive Extrusion Ring-Opening (Co)polymerization of e-Caprolactone PCL is an aliphatic polyester currently prepared by ROP of CL catalyzed with stannous octanoate (Sn(Oct)2) in the presence of heavy alcohol (initiator) such as 1-dodecanol,[54] using a batch process with a maximum number-average (Mn ) of 80 kg mol1, and marketed under the trade names TONE and CAPA by Dow Chemicals and Solvay, respectively. However, these PCL polymers suffer from poor processing characteristics like low melt-strength, and for certain applications, inadequate thermo-mechanical properties like tear strength.

Figure 3. Coordination-insertion mechanism of the ROP of CL.


Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Kim at al. reported bulk ROP of CL carried out both in a laboratory internal mixer (Brabender Plasticorder) and in a modular intermeshing co-rotating twin-screw extruder.[5557] Various initiators such as titanium n-butoxide, aluminum triisopropoxide, and sodium hydride were rst used to polymerize CL in an internal mixer. High conversion in PCL could be effectively obtained when aluminum triisopropoxide initiated the polymerization of CL. This polymerization was then investigated through reactive extrusion processing for various ratios of monomer to initiator using aluminum triisopropoxide (Al(OiPr)3) as initiator as well as under a range of different processing conditions, including barrel temperature proles, throughput, and screw speed. GPC analyses demonstrated that high molecular weight PCL, together with signicant quantities of PCL oligomers were produced at different reaction temperatures. Increasing screw speed and decreasing the throughput caused a severe reduction in the PCL molecular weight once the maximum molecular weight was obtained after the rst mixing segment. Higher molecular weight PCL produced by increasing ratios of monomer to initiator resulted in a more severe reduction of the molecular weight during reactive extrusion. This was ascribed to both specic mechanical energy (i.e. the energy consumed per unit of mass of the material extruded) and the molecular weight.[57,58] A more detailed study about rheological developments for reactive processing of PCL upon both experimental and modeling aspects has been reported elsewhere,[1619] and has recently been reviewed in International Polymer Processing.[6063] Recently, we have reported that aluminum sec-butoxide (Al(OsecBu)3) was a more suitable initiator in the reactive extrusion ROP of CL. Although Al(OiPr)3 has shown to efciently initiate the ROP of CL, it requires being sublimated rst, and then dissolved in an organic solvent like toluene in order to control the ROP of CL. In contrast, Al(OsecBu)3 is commercially available as a pure liquid, and therefore does not require any previous purication step before use. In addition, Al(OsecBu)3 has shown to be an efcient initiator in the bulk ROP of CL carried out in small reactors.[64,65] Interestingly, when Al(OsecBu)3 was used as initiator in reactive extrusion ROP of CL, a well-controlled synthesis of PCL in terms of molecular weight and polydispersity (Mw =Mn 1.7) was successfully obtained at a temperature ranging from 130 to 180 8C in an intermeshing twin-extruder with the screw conguration made up with only conveying elements. The conveying elements were selected to avoid as

452

DOI: 10.1002/mame.200700395

Recent Advances in Reactive Extrusion Processing . . .

stiffness of lactam part, lactone-lactam terpolymers having block and random caprolactam structure, i.e. P(LLa-bCLa-b-CL) and P(LLa/CLa-b-CL) have also been prepared through the same procedure. For seat belt applications, new bers derived from poly(ethylene terephthalate)block-PCL block copolymers have been developed through REX.[69] Such bers provide the desired load-limiting perforFigure 4. Three-arm star-shaped PCL as prepared by ROP of CL initiated with Al(OsecBu)3 mance for the design of safety seat belt. in reactive extrusion. The synthesis of poly(ethylene terephthalate)-block-PCL block copolymers was carried out by ROP of CL initiated by hydroxylmuch as possible the undesirable thermal degradation terminated poly(ethylene terephthalate) (PET) together reaction that occurs when kneading elements are used. with Sn(Oct)2. A block copolymer with minimal transesLike Al(OiPr)3, the polymerization of CL promoted by Al(OsecBu)3 proceeds through the so-called coordinationterication reactions could be obtained in reactive extrusion at 290 8C because of the fast distributive mixing insertion mechanism, yielding polyester chains endcapped by a growing aluminum alkoxide bond (see of CL into the high melt-viscosity PET and the short reaction time. After ber spinning, these PET-b-PCL block Figure 3).[64,65] Due to the trifunctionality of Al(OsecBu)3, copolymers exhibited high degree crystalline orientation. three-arm star shaped PCL with Mn of each arm as high as 200 kg mol1 could be prepared with monomer conversions larger than 95% and within residence times of less than ve minutes in the extruder (Figure 4). In blown lm Reactive Extrusion Ring-Opening (Co)polymerization applications, the resulting PCL displayed signicant better of L,L-Lactide dart and tear properties than commercially available linear PCL. A tremendous demand is raising on the synthesis of PLA as As potential compatibilizing agents in polyamides/PCL obtained by ring-opening polymerization of L,L-lactide (LA) blends, lactam-CL block copolymers were prepared because LA derives from renewable resources, avoiding the through reactive extrusion as well.[66,67] Continuous depletion of petrochemical resource, and the emission of green gas (CO2). The cyclic dimer of lactic acid is recovered copolymerization of CL with e-caprolactam (CLa) and v-lauryl lactam (LLa) were carried out in a modular after fermentation of corn or sugar beets followed by a intermeshing co-rotating twin-screw extruder. Sodium free-solvent polymerization/depolymerization process hydride (NaH) and N-acetyl caprolactam were employed, (Figure 5). respectively, as co-initiators in the synthesis of lactamInterestingly, a continuous one-stage reactive extrulactone copolymers. In the presence of N-acetyl caprolacsion[7074] has been reported by some of us on the tam, it is considered that CLa is unable to act as a faster manufacture of economically viable PLA via ROP of LA activator for polymerization of CL than N-acetyl capropromoted by Sn(Oct)2. This technique requires that the lactam as described by Goodman et al.[68] In this respect, bulk polymerization be close to completeness within a very short time (57 min), which is predetermined by the NaH was added as a co-initiator together with N-acetyl residence time distribution of the extrusion system, and caprolactam for polymerization of CLa before the reaction that the PLA stability is high enough at the high processing of sodiocaprolactam with CL. It has been demonstrated that the order of addition is of prime importance for the temperature. Although Sn(Oct)2 can promote quite fast formation of blocky copolymer. Simultaneous feeding of polymerization of LA, it is well-known to provide adverse both monomers with NaH and N-acetyl caprolactam in the effects on both the molecular weight and properties of PLA rst hopper of the twin-screw extruder produces a mixture as a result of back-biting and intermolecular transesterof homopolymers. In contrast, both high molecular weight ication reactions, not only during the polymerization of P(CLa-b-CL) and P(LLa-b-CL) block copolymers have been LA, but also during any further melt-processing.[75,76] In successfully achieved by adding the lactam (LLa and CLa) this respect, an equimolar amount of triphenylphosphine into the rst hopper and the CL sequentially into (P(C6H5)3) has been added to Sn(Oct)2 in order to enhance the second hopper. The respective block lengths of the the rate of LA polymerization signicantly, but also to copolymer could be adjusted by controlling the feed rate of suppress (or at least delay) any degradation reactions such each monomer during reactive extrusion. To modulate the as transesterication reactions.[70,71] This kinetic effect has
Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.mme-journal.de

453

J.-M. Raquez, R. Narayan, P. Dubois

Figure 5. PLA production via prepolymer and lactide.

been accounted for the coordination of the Lewis base onto the tin atom, making easier the insertion of the monomer into the metal alkoxide bond of the initiator/propagation active species. This tin alkoxide bond is formed in situ by reaction of alcohol and the tin(II) dicarboxylate, and proceeds through the aforementioned coordinationinsertion mechanism.[50,51] The addition of one equivalent of P(C6H5)3 onto Sn(Oct)2 allows reaching an acceptable balance between propagation and depolymerization rates, so that the polymerization is fast enough to be performed through a continuous one-stage process in an extruder. Using a closely intermeshing co-rotating twin-screw extruder (with a suitable processing and screw concept), the equimolar Sn(Oct)2/P(C6H5)3 complex was used as a catalyst system in ROP of LA yielding high molecular weight PLA within a residence time of ca. 7 min at high monomer conversions (ca. 98%) and at a temperature of about 180185 8C. Adding alcohol as (co)initiating system could easily adjust molecular weights of as-recovered PLA. In addition, by reducing the amount of the catalytic complex ([LA]0/[Sn] 5 000), the resulting PLA exhibited good melt-stability during further melt processing such as melt-spinning. For more specic applications, the melt-stability for the resulting PLA[77,78] could be further enhanced by adding stabilizers (Ultranox 626) during reactive extrusion, without inuencing the course of the polymerization reaction. Interestingly enough, some of us demonstrated that reactive extrusion processing could enhance much more the kinetics of ROP of LA in bulk than the conventional batch polymerization technology such as glass reactors. Indeed, although the conversions in LA reached the same equilibrium values (ca. 98%) whichever the polymerization system, the time required for reaching
Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

this monomer conversion was approximately 40 min using glass ampoules as reactor, compared to ca. 7 min in the reactive extrusion process. At low monomer conversion, the relative enhancement on polymerization kinetics can be ascribed to the molten temperature of reaction medium inside the extruder, being a key-parameter in reactive processing.[59] In batch polymerization, depending on the extent of polymerization, there is a temperature gradient with a bad thermal transfer for the reaction medium, leading to longer reaction times. By contrast, the heat transfer is better due to high mixing phenomena in extruders.[60,61] However, at high monomer conversion, the rate of the reaction gets limited not only by the molten temperature and the reactivity of the chemicals, but also by the diffusion of the monomers (and other low molecular weight compounds) inside the high viscous melt to nd a reactive partner, when high molecular weight polymer is reached. This physical movement is limited to the Brownian movement in the glass ampoule, but is well supported in the twin-screw extruder by the mixing elements and by the shearing of the polymer inside the intermeshing zone. The preparation of block copolymers based on PLA has also been carried out in a co-rotating twin-screw extruder.[73,74] Different lengths of v-hydroxylated prepolymers such as PCL or poly(ethylene oxide) (PEO) as macroinitiators enable to prepare a multitude of possible block polymers with the same processing concept and equipment. Other authors have attempted to prepare multi-block copolymers based on PLA through a controlled number of transesterications.[79] In the rst attempt, various catalysts (nBu3SnOMe, Sn(Oct)2, Ti(OBu)4, Y(Oct)3, and para-toluene sulfonic acid) were added to promote these transesterication reactions of PCL against PLA and PEO prepolymers. In blends of PLA and PCL (50:50 by weight), the use of nBu3SnOMe was reported to catalyze the transesterication reactions between PLA and PCL. If 2 wt.-% in nBu3SnOMe was added to the blend, some transesterication reactions occurred during reactive extrusion, but substantial PLA degradation took place also as evidenced by the formation of large amounts of LA monomer (%12 wt.-%). Smaller amounts in nBu3SnOMe were not effective in promoting transesterication under these reaction conditions. The use of the other studied catalysts, i.e. Sn(Oct)2, Ti(OBu)4, Y(Oct)3, and para-toluene sulfonic acid, revealed that no sufcient transesterication reactions occurred, and in all cases the PCL was recovered

454

DOI: 10.1002/mame.200700395

Recent Advances in Reactive Extrusion Processing . . .

unaffected, while PLA degradation was observed as indicated by the formation of LA and fast decrease in molecular weight. In this respect, Stevels et al. have preferred to effectively prepare multi-blocky PLA-bPCL-b-PLA and PLA-b-PEO-b-PLA terpolymers with moderate molecular weights at 180 8C by reactive extrusion ROP of LA initiated with a,v-hydroxyl PCL and PEO in presence of Sn(Oct)2. High yield (more than 90 wt.-% in LA converted into polymer) could be achieved within a few minutes.

Reactive Extrusion Ring-Opening (Co)polymerization of 1,4-Dioxan-2-one Poly(1,4-dioxan-2-one) (PPDX) obtained by ROP of 1,4-dioxan-2-one (PDX) appears to be another attractive candidate as a biodegradable substitute for commodity polymers. This aliphatic poly(ester-alt-ether) copolymer offers a good compromise between its processing temperature and its service temperature range because of a melting and glass transition temperatures of 110 8C and 10 8C, respectively.[40,41] Furthermore, PPDX has proven to be tougher than polylactides and even HDPE with a tensile strength close to 50 MPa for an ultimate elongation ranging from 500% to 600%. The use of PPDX materials as bulk materials is however restricted as a result of its low ceiling temperature (265 8C). This favors the unzipping depolymerization reactions from the hydroxyl end-group in the molten state (Figure 6). Again, like other cyclic esters such as CL, aluminum alkoxides can efciently promote ROP of PDX, which offers the possibility for the synthesis of high molecular weight PPDX through a fast and continuous process in a twinscrew extruder.[41] Besides these kinetic considerations, developing PPDX as biodegradable thermoplastics requires reducing its thermal degradation essentially due to unzipping reactions, while preserving as much as possible the semi-crystalline properties of PPDX (high melting temperature slightly above 100 8C). Therefore, simulta-

neous bulk (co)polymerization of PDX promoted by Al(OsecBu)3 (([PDX]0 [CL]0)/[Al] 2 500) was carried out with low molar fractions in CL ranging from 0 to 16 mol-% in a co-rotating twin-screw extruder at 130 8C. Indeed, random distribution of a few units of CL all along the PPDX backbone[43] represents the best way to prevent or at least limit undesirable unzipping reactions of PPDX chains. As far as the homopolymerization of PDX is concerned, the polymerization conversion did not exceed more than 65%, corresponding to the equilibrium value obtained at 130 8C, in a good agreement with previously published data performed in bulk and in sealed glass ampoules.[41] Interestingly, with a limited amount in CL, the PDX (co)polymerization yield reached the completion, attesting for the formation of a blocky structure for the resulting P(PDX-co-CL) copolymers. Both 13C and 1H NMR spectroscopy indicated that multiple short PPDX homosequences separated from each other by CL unit(s) composed the resulting copolyester chains. As shown by TGA and DSC experiments, this allowed enhancing signicantly the thermal stability of PPDX chains, without preventing the crystallization of the resulting copolymers with a melting temperature as high as 95 8C at a CL molar fraction close to 11 mol-%. These relevant features open the door to the manufacture of low cost PPDX-based materials at large scale using a continuous one-step process.

Reactive Extrusion Modication of Biodegradable Polymers


Chemical modication is usually employed to enlarge the potential applications of biodegradable polymers. For instance, to better suit their properties to specic applications, chemical modication of starch is often required because of the dominant hydrophilic character, and unsatisfactory mechanical properties, particularly in wet environment of starch. Typical examples of reactive extrusion modication of biodegradable polymers will be outlined in the following sections.[18]

Grafting onto Reactions of Various Monomers onto Functional Biodegradable Polymers Over the past decades, starch, an anhydroglucose polymer, has attracted considerable attention as an interesting structural platform for the manufacture of sustainable and biodegradable plastic packaging due to its natural abundance and low cost.[7]

Figure 6. Unzipping depolymerization mechanism of v-hydroxyl PPDX chains.


Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.mme-journal.de

455

J.-M. Raquez, R. Narayan, P. Dubois

Starch is a polymer made from anhydroglucose (C6H10O5) units attached either by a-1,4 linkages or by a-1,6 linkages.[80] The a-1,4 linkages yield a linear homopolymer called amylose with molecular weights ranging from 100 000 to 500 000 g mol1, while the a-1,6 linkages are at the origin of the branching points in the polysaccharide structure called amylopectine with molecular weights higher than millions (Figure 7). It is worth noting that the proportion of both amylose and amylopectine is genetically established and is relatively constant for each species of plant. Due to its high functionality in hydroxyl functions, various types of monomers have been grafted onto unmodied starch through reactive extrusion. In papermaking and textile applications, cationic starches have been successfully prepared through reactive extrusion in the presence of 3-chloro-2-hydroxypropyltrimethylammonium chloride (CHPTMA) as monomer, and of an alkaline catalyst, yielding a quaternary ammonium cationic starch ether [starch O CH2 CHOH CH2N(CH3)3].[81] Using a Clextral BC 45 twin-screw extruder as a reactor, a reaction efciency of up to 82% was obtained in extrusion processing with wheat starch within only a few minutes. Carrs et al. proposed a similar study[82] with high reaction efciency of up to 90% or more for the system of unmodied starch reacting with CHPTMA (molar ratios monomer/starch of 1:2) using NaOH as catalyst in a ZSK 30 twin-screw extruder. Reaction temperature was kept at 70 8C for all experiments. The combined effect of the high temperature (90 8C), intensive mixing, high-starch solids (65%), and appropriate level of catalyst contributed to the unusually high reaction efciency values (90%), exceeding maximum values previously reported using laboratorybatch reaction procedures. Carr et al. reported grafting of acrylic acid (AA) and acrylamide (AC) onto starch free-radically promoted by aqueous ceric ammonium nitrate (CAN), carried out using a twin-screw extrusion.[83] Presumably, free-radicals are formed at carbon atoms 2 or 3 in starch that are able to initiate the polymerization with acrylic compounds as

Figure 7. Structure of amylose (a) and amylopectine (b).


Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

evidenced by electron spin resonance. In the presence of 35 wt.-% in water, and at 80 8C, levels of add-on (wt.-% of total synthetic polymer) of the starch grafted with polyacrylic compounds (St-g-PAC) were of 2744% for the extrusion process. Interestingly, the average residence times in the extruder were about 23 min, largely below two hours, which are currently required in a batch process. When AA was used as monomer, the saponication of resulting St-g-PAC led to highly water-absorbent materials for hygienic and cosmetic applications. Cellulose, a polysaccharide constituted by anhydroglucose (C6H10O5) units attached via b-1,4 linkages, is the most abundant biomass on the surface of the earth. Cellulose can be converted into a wide range of derivatives with desired properties and applications. The most important derivative from cellulose is cellulose acetate used in many common applications including toothbrush handles and adhesive tape backing. The common synthesis of cellulose acetate is the processing of high-grade cellulose in the presence of a mixture of methylene chloride and acetic anhydride.[84] The resulting cellulose acetate is processed into strands, sheets and lms by addition of liquid additives using the extrusion technique. Cellulose acetate corresponds to a thermoplastic with good barrier properties to grease and oil, and is used for the packaging coating of food. Until the mid-1990s, plastic-grade cellulose acetates were believed to be non-biodegradable due to their high substitution degree (SD).[8487] Between two and three hydroxyl groups of the glucose repeat unit are most often acetylated. However, it has been shown that cellulose acetates with SD up to 2.5 are biodegradable in stimulated composting.[87] A decrease in SD from 2.5 to 1.7 results in a large increase in the rate of their biodegradation. In this respect, thermoplastic cellulose acetates are regaining interest as potentially biodegradable plastics for composting of plastic waste without encountering the water-solubility problems typical for starch-based materials. Low molecular weight plasticizers like phthalates, glycerol, triacetine, or cyclic lactones are preferred for easy processability.[88] However, their main disadvantage is related to plasticizer migration that can account for loss of mechanical properties. Novel families of thermoplastic cellulose acetate starting from cellulose-2,5-acetate were produced through reactive extrusion technology that grafted cyclic lactones, simultaneously onto polysaccharide, hydroxyfunctional plasticizer, optionally also hydroxyfunctional llers.[89] Organosolv ligin, cellulose, starch, and chitin were added to reinforce the polymer matrix. It was

456

DOI: 10.1002/mame.200700395

Recent Advances in Reactive Extrusion Processing . . .

demonstrated that cyclic lactones such as CL, glycolide, and lactide could be successfully grafted onto the rigid polysaccharide backbone, and hydroxyfunctional plasticizers in a twin-screw extruder system (length/diameter ratio was of 48) at 190 8C. Low molecular weight hydroxyfunctional plasticizers such as glycerol were in situ converted into non-migrating higher molecular weight oligopolyester plasticizer. The simultaneous grafting of cellulose-2,5-acetate and hydroxyfunctional plasticizer substantially improved compatibility between these two components. These resulting oligolactone-modied cellulose-2,5-acetate exhibited high compatibility with llers such as organosolv ligin, cellulose, starch, and chitin when these latter were added in a subsequent downstream reactive processing.

Acid/acetyl Derivatization Carried out onto Functional Biodegradable Polymers Acid derivatization of starch is a well-known technique to obtain lower viscosity products, which are dispersible at higher solids than one made from the native starch and one whose dispersions are still able to be pumped and handled.[9093] Miladinov et al. reported the reactive extrusion preparation of starch-fatty acid esters containing 0.010.03 mol levels in co-organic acid anhydrides (acetic, propionic, heptanoic, and palmitic anhydrides) with regard to the degree of substitution of starch, in the presence of sodium hydroxide as catalyst.[94] Some molecular weight reduction of the amylopectin fraction could be detected that lowered the specic mechanical energies for REX, particularly when heptanoic and palmitic anhydrides were used as co-organic acid anhydrides. Reactive extrusion preparation of starch esters has been carried out using maleic anhydride (MA) as cyclic dibasic acid anhydrides, yielding a free carboxylic group. Such a free carboxylic group has shown to be valuable to promote acid-catalyzed esterication reactions with biodegradable

poly[(butylene adipate)-co-terephthalate], leading to the formation of a graft copolymer in subsequent downstream blending operations.[53,95] In situ reactive modication of starch by 08 wt.-% in MA as esterication agent and in the presence of 20 wt.-% glycerol as plasticizer has been carried out through reactive extrusion at 150 8C. When 2.5 wt.-% MA was used, the recovery yield for the resulting maleated thermoplastic starch (MTPS), as determined by Soxhlet extraction, was almost complete. This attests for the reaction of some glycerol moieties to starch backbone during the extrusion process. Increased MA content decreased the recovery yield of resulting MTPS due to its partial solubilization in the solvent used for Soxhlet extraction (acetone). Besides the in situ esterication of TPS, intrinsic viscosity, FT-IR and 2D liquid-phase NMR spectroscopy measurements proved the occurrence of some hydrolysis and glucosidation reactions as promoted by MA moieties grafted onto the starch backbone. Such reactions reduced the relative molecular weight of MTPS (Figure 8). Therefore, the resulting MTPS had improved processability due to its reduced melt viscosity. Indeed, TPS has found to display a gel-like viscoelastic behavior. This is related to the formation of a crystalline elastic network produced by the complexation of amylose molecules with lipids/plasticizers, and the physical entanglement of starch chains caused by its high molecular weight. Such a physical entanglement is responsible for the incomplete homogenization in the melt blend of TPS with other polymers such as PCL, affecting the nal properties for the resulting melt-blends.[96] WAXS diffraction analyses conrmed the complete disruption of the granular structure of native starch in MTPS during the reactive extrusion processing. Tomasik et al.[90] reported a similar chemical modication of cornstarch using MA and the like, and by varying amounts of water (18, 20 and 30%) as plasticizer, the whole process was carried out by extrusion. Carbonate buffer, either at pH 8 or pH 9, was added during extrusion. Extrusion of starch with cyclic anhydrides in alkaline medium represented a facile method for the preparation of

Figure 8. Hydrolysis (a) and glucosydation (b: for sake of clarity, only the reaction between starch and the hydroxymethylene function from glycerol is represented) reactions present during the in situ maleation of TPS.
Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.mme-journal.de

457

J.-M. Raquez, R. Narayan, P. Dubois

anionic thermoplastic starches. However, such modied hydrophilic thermoplastic contained large amounts in water, e.g. rendering difcult their subsequent melt blending with hydrophobic biodegradable polymers. Soy proteins are the rst biopolymers derived from agriculture, which have been used for the manufacture of molded materials. Among plant protein sources, soy protein is of relatively low cost with vast available supplies. Soy proteins are available in three different forms as soy our, soy isolate, and soy protein concentrate. Soy proteins are complex macromolecules containing 20 amino acids that supply enough available sites to react with coupling agents. As biomaterials, soy proteins can be masterfully converted into soy protein plastics through extrusion with a plasticizers.[97,98] Common plasticizers used in the manufacture of soy protein plastics include glycerol, ethylene glycerol, propylene glycerol, 1,4-butanediol, 1,3butanediol, poly(ethylene glycol) sorghum wax, and sorbitol.[99] Thermoplastic processing of proteins with plasticizers induces marked changes in the connectivity of the protein network, which limits the processing window. Cross-linking of proteins occurs in situ as a temperaturecontrolled phenomenon, but high shear conditions can signicantly decrease the activation energy of this reaction. Interestingly, Vaz et al. used the high functionality of soy proteins in the reactive extrusion for the design of biodegradable soy matrix drug delivery systems.[100,101] Glyoxal was used as a cross-linker, which has dialdehyde functionality able to react with the free e-amine groups of the lysine (hydroxylysine) residues of soy protein (Figure 9). In situ cross-linking reactions of soy protein were carried out at ca. 130 8C under rotation speed at 100 RPM, in the presence of glyoxal and of an encapsulated drug (theophylline) through one-stage reactive extrusion process at pH 4 and 7. The resulting drug delivery systems offer several advantages as follows: (i) ease of production, (ii) suitability for a large variety of polymeric matrices, (iii) applicability to different types of drugs, and (iv) biodegradability. The drug release patterns could be adjusted during processing by cross-linking, changing the net charge (effect of pH), and by a ller reinforcement such as hydroxyapatite. Corn gluten meal chemically modied with citric derivatives has also been prepared in a continuous reactive extrusion process.[102] Corn gluten meal is a mixture of corn starch, ber and corn protein obtained, as a by-product, from the wet-milling of corn in the ethanol

industry. In this study, citric anhydride was reacted with this corn product in reactive extrusion to generate valueadded, acid-insoluble reaction products with enhanced metal-binding properties for the treatment of industrial wastewaters. Pendant carboxylic groups were so obtained from citric anhydride that reacted with the different nucleophilic groups of corn gluten meal from the part of starch/bers (hydroxyl functions), and corn protein (hydroxyl, sulfhydryl, and amino groups). Interestingly, these derivatives exhibited high metal binding ability for, at least nine metals (Cd2, Co2, Cu2, Fe2, Pb2, Mn2, Ni2, Ag, and Zn2). Respirometry testing revealed that biodegradability of corn gluten meal was unaffected after treatment with citric anhydride.

Free-Radical Grafting of Functional Groups onto Biodegradable Polymers through Reactive Extrusion

In order to incorporate functional pendant groups all along the polymer backbone, free-radical grafting of unsaturated monomers such as maleic anhydride, and acrylic acid and its derivatives has received much attention in the reactive extrusion technology.[2426,30,31] The most preferred unsaturated monomer is actually maleic anhydride (MA) that is useful for further compatibilization reactions with, e.g., hydrophilic llers like starch granules.[103] The reason of this interest is that MA does not readily polymerize under the conditions employed in grafting reactions, and is therefore grafted at high efciency without the accompanying formation of homopolymer. A number of methods are available in the literature to produce such a graft polymer that includes melt-grafting, solid state grafting, solution grafting, suspension grafting in aqueous or organic solvents, and redox-induced grafting.[104] The most widespread method is the melt state process performed by REX, which can alleviate the difculties owing to diffusion-controlled grafting reactions of high molten viscosities polymeric matrix in bulk. Tang et al. investigated thoroughly the grafting of MA onto biodegradable aliphatic/aromatic (co)polyesters through reactive extrusion using dicumyl peroxide, benzoyl peroxide, and di-tert-butyl peroxide as free-radical initiators.[1] The effect of various factors such as freeradical initiator concentration, MA concentration, and reaction temperature on the percent grafting of MA onto the copolyesters was investigated. The structures of the grafted (co)polyesters were characterized using FT-IR and NMR spectroscopy. It has been demonstrated that MA can be grafted onto any copolyesters. The grafting reaction takes place selectively at aliphatic dicarboxylic Figure 9. Acetyl functionalization of soy proteins from e-amine groups of the lysine acid units of the copolyesters (Figure 10). Minimal degradation of the polyester (hydroxylysine) residues.

458

Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

DOI: 10.1002/mame.200700395

Recent Advances in Reactive Extrusion Processing . . .

Figure 10. Free-radical grafting of PLA backbone.

chains was also observed from intrinsic measurements. The desired graft content can be controlled by the aforementioned factors, but there is, however, an optimum radical concentration, which depends on the [freeradical initiator]/[MA] ratio to promote grafting efciency, beyond which the chain scission and termination reactions become predominant. In PLA, low grafting efciency was observed due to its limited availability of free-radical sites on the polymer backbone for grafting. Some of us have carried out a more consistent study on the free-radical grafting reaction of MA onto the PLA backbone using a co-rotating intermeshing twin-screw extruder (L/d ratio of 14) at two reaction temperatures (180 and 200 8C).[104] For all experiments, 2 wt.-% MA was used, while varying the free-radical initiator concentration, i.e. 2,5-dimethyl-2,5-di-(tert-butylperoxy)hexane (Luperox L101), between 0 and 5 wt.-% by PLA. Triple-detector size exclusion chromatography (TriGPC), melt-ow index, and thermal gravimetric analysis were used as main characterization tools. Whichever the processing temperature, increasing free-radical initiator content increases the content of MA moieties grafted onto PLA chains, as determined by back titration of an excess of morpholine with HCl. However, the molecular weight for the resulting MA-g-PLA decreased with further addition of Luperox 101 during the reactive extrusion processing. Such a behavior is more likely due to a competition between the molecular weight increase through chain branching, and the molecular weight decrease by b-chain scission triggered by grafted MA (Figure 11). In contrast, the simple addition of Luperox 101 to the extruded PLA allows increasing its molecular weight to some extent through a free-radical self-branching reaction. However, highly branched PLA chains until the formation of microgel could occur at higher temperature as shown by TriGPC. This was more

likely due to the hydrogen radical abstraction in a-position of the carbonyl groups followed by radical addition onto the carboncarbon double bond of the enolate forms in equilibrium within the polyester chains. Additional free-radical chain scissions might occur and participate in the chain branching process as well. Interestingly, when these lowmaleated PLA (0.7 wt.-% MA) were melt-blended with granular cornstarch (30 and 40 wt.-%) again by reactive extrusion, improvement in the interfacial adhesion of PLA-based composites could be successfully achieved. These grafted reactive functions can therefore react with the hydroxyl groups of starch macromolecules to form covalent bonds, and thus, they provide better control of the size of phase and strong interfacial adhesion.

Reactive Extrusion Oxidation Reactions of Biodegradable Polymers Commercial oxidized starches are batch-prepared at room temperature conditions and low (3%) concentrations of oxidant (usually hypochlorite) for adhesive applications.[105] During product isolation (ltration and aqueous washing), some of the product is dissolved (due to molecular breakdown) and gets lost. Hypochlorite has been the oldest and most frequently used oxidant. Other oxidants such as permanganate, hydrogen peroxide, persulfate, periodate and dichromate have also been used. The different oxidation procedures result in variations in molecular structure and properties. Wing and Willett[106] used the reactive extrusion to prepare oxidized starches. Three types of cornstarches (waxy cornstarch, pearl cornstarch, and amylomaize) were oxidized by a reactive extrusion-drum drying process with hydrogen peroxide and a ferrous-cupric sulfate catalyst. A ZSK 30 corotating twin-screw extruder, which had a temperature prole of 88/115/88/77/71/65/54/48 8C (feed to die) and a screw speed of 110 rpm, with a starch feed of 180 g min1 (10% moisture) was used. Thermo-chemical oxidation of starch by reactive peroxide extrusion-drum drying represents a rapid, continuous procedure for making water-soluble products containing high carboxyl and carbonyl contents. Increasing the peroxide level increased the carboxyl and the carbonyl content, whereas increasing the amylose content decreased the solubility. The residual granule structure was still present in high amylose starch

Figure 11. Free-radical b-scission of PLA backbone.


Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.mme-journal.de

459

J.-M. Raquez, R. Narayan, P. Dubois

extrudates with or without peroxide. Solution viscosities indicate a signicant molecular degradation by reactive extrusion.

Reactive Extrusion Melt Blending of Biodegradable Polymers


Polymer melt-blending is a well-used technique whenever modication of properties is required, because it uses conventional technology at low cost.[107] The objective for preparation of novel melt-blends between two or more polymers is not to change the properties of the entire components drastically, but to capitalize on the maximum possible performance of the blend. Substantial efforts have been realized in the preparation of biodegradable starchbased melt-blends using the reactive extrusion technology, due to the abundance, renewability and low cost production of starch.[108122] Interestingly, hydroxyl groups cause starch to behave as an alcohol during chemical reactions generally. This property of starch is important when considering reactive melt blending of starch with synthetic polymers. The presence of such large numbers of hydroxyl groups affords starch hydrophilic properties, and therefore adds afnity for moisture and dispersability in water. However, hydrophilicity is undesirable in many plastic packaging applications, and hence it is a major limitation in using starch as a homopolymer. Meltblending starch with moisture resistant and biodegradable polymer having good mechanical properties represents the best method to prepare useful products, while maintaining the biodegradability of overall products. Besides, llers such as talc and kaolin are frequently incorporated in thermoplastics to reduce the costs of molded products.[123] These llers improve the properties of the polymers such as strength, rigidity, durability, and hardness. Particularly, worldwide for at least twenty years, there has been a new and intense desire to tailor the structure and composition of materials applying sizes about the nanometer. Nanollers having at least one of their dimension in the nanometer scale, exhibit high specic surface areas able to signicantly improve the properties of polymeric (nano)composites at low content in contrast to microllers.[33] Developing such melt-blends/(nano)composites with satisfactory overall physico-mechanical behavior however requires the ability to control interfacial tension, to generate a dispersed phase of limited size and strong interfacial adhesion, and to improve the stress transfer between the component phases.[12] Compatibilization is therefore called upon. Usually, this is achieved by adding or creating in situ during the blending process, a third component, often called an interfacial agent, emulsier, or compatibilizer. The latter can be a graft or a block
Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

copolymer. Effective compatibilizers must be located at the interface between the phase domains of the immiscible blend. Most importantly, the degree of compatibilization in a particular system depends on the reactivity of the compatibilizer used. It has been found that a compatibilizer is most effective when its sections are of higher molecular weight than the corresponding blend components. Several theories have attempted to explain the role of compatibilizers, of which two mechanisms are considered plausible. The rst mechanism is thermodynamic in nature in that the compatibilizer reduces the interfacial tension between the phases. The second mechanism is kinetic in nature in that the presence of the compatibilizer at the interface reduces the agglomeration of domains by steric stabilization.[1,124] Recently, Macosko et al. have reported an excellent review about reactions at the polymerpolymer interface for blend compatibilization.[125] It has been shown that the major factors inuencing the interfacial reaction, and therefore blend compatibilization are: inherent reactivity of functional polymers used as compatibilizer, thermodynamic interactions between different polymeric partners, functional group location along the compatibilizer chain, and the effect of processing ows. Concerning the last factor, ow in melt mixers is well-known to accelerate the coupling rate tremendously, by changing the concentration prole of functional groups from compatibilizer at the interface and/or increase the collision probability. However, in this case, it is often unclear which of the mechanisms is predominately involved under simple model ows. Reactive Extrusion Starch-Based Melt-Blending Starch is largely used as ller for environmentally friendly plastics since about two decades. Although starch may be added as ller, its more interesting and technologically challenging uses have been in the area of using starch as a binder, as a thermoplastically processable constituent within thermoplastic polymer blends, and as a thermoplastic material by itself.[126] While native starch does not typically behave as a thermoplastic material by itself, it is a thermoplastic in the presence of a plasticizer when heated and sheared.[127,128] Glycerol and water are the most widely used plasticizers. The role of plasticizers is to destructurize by cleaving hydrogen bonds between the starch macromolecules, and by inducing partial depolymerization of starch polymers. It contributes to lower the melting and the glass transition temperatures below its decomposition temperature (230 8C).[13,129,130] However, the physical properties of polyol-plasticized starch tend to be modied after being stored for a long period because of its re-crystallization (retrogradation), with migration of plasticizers. To limit the retrogradation phenomenon, amide compounds as plasticizers such as urea and

460

DOI: 10.1002/mame.200700395

Recent Advances in Reactive Extrusion Processing . . .

formamide were added in the plasticized starch preparation, but the resulting materials were very rigid and brittle.[131,132] The starch structure may be modied by, e.g., acetylation to reduce the hydrophilic character of the macromolecules. This hydrophobic starch acetate was shown to be useful in starch-based extruded foams.[133135] However, this chemical process results again in inferior mechanical properties and greater product cost for these starchy materials.[136] Therefore, some authors have preferred to melt-blend TPS with biodegradable hydrophobic polymers such as PCL and cellulose acetate in order to manufacture environmentally friendly products.[108119,137,138] As excepted, melt-blending TPS with these polyesters resulted in a signicant improvement in the properties of plasticized starch. However, although a protective polyester skin layer is formed at the surface of most blends during injection molding, the moisture sensitivity of TPS can be not fully addressed. Averous et al. prepared different compositions of wheat thermoplastic starch and PCL through a two-stage extrusion in order to determine the thermo-mechanical properties of resulting melt-blends.[139] A large range of blends was analyzed with different glycerol (plasticizer):starch content ratios (0.14:0.54) and various PCL concentrations (up to 40 wt.-%). Whatever the composition, it resulted a phase-separation occurring in the melt-blend as observed by DSC and DTMA. Depending on the content in glycerol, two distinct behaviors could be obtained. When the starch matrix has a glassy behavior (low content in glycerol), extrusion blending with PCL resulted in a decrease of the material modulus, but the impact resistance was improved. On the other hand, when the starch has a rubbery behavior, PCL increased the modulus of the materials. However, ageing studies showed a structural evolution for the resulting melt-blends after processing during several weeks. A signicant increase of Youngs modulus and of the maximum strength was observed for all melt-blends. This is due to post-crystallization and water evolution inside melt-blends. Other efforts have been employed to develop a coating of TPS with hydrophobic biodegradable polyester through reactive extrusion. Multilayer coextrusion has been widely used in the past decades to combine the properties of two or more polymers into one single multi-layered structure. Martin et al. prepared multilayer lms based on plasticized wheat starch (PWS) and various biodegradable aliphatic polyesters through at lm coextrusion.[140] PLA, polyesteramide (PEA), PCL, poly(butylene succinate adipate) (PBSA), and poly(hydroxybutyrate-co-valerate) (PHBV) were chosen as the outer layers of the stratied polyester/ PWS/polyester lm structure. Different levels of peel strength were found, depending on the compatibility of plasticized starch with the respective polyesters. In
Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

particular, PEA presented the best adhesion to the PWS layer, probably due to its polar amide groups. PCL and PBSA showed medium adhesion values, and both PLA and PHBV were the least compatible polyesters. The same trend in the magnitude of adhesion strength was observed whichever the multilayer techniques. However, some inherent problems rise due to the multilayer ow conditions encountered in co-extrusion such as an encapsulation and interfacial instabilities phenomena because of the difference of hydrophilic balance between TPS and the polyesters. For formation of in situ bonding, electron irradiation carried out TPS layers has also been attempted, but the results were quite unsatisfactory. Anhydride functionalization of the biodegradable hydrophobic polymers represents another method of compatibilization largely utilized in the reactive extrusion preparation of starch-based melt-blends with useful end-properties.[1,9093,138,141144] For instance, MA and dicumyl peroxide (DCP) were used as cross-linking agent and initiator respectively for blending plasticized starch with a biodegradable aliphatic/aromatic (co)polyesters, called EnPol1, using a two-stage reactive extrusion process (L/D 40:1).[145] The rst step was the preparation of maleated polyester, and the resulting polyester was then melt-blended with 40 wt.-% in starch plasticized with glycerin as a plasticizer. It has been demonstrated that at a peroxide initiator constant varying MA content higher than 2 wt.-% were found to be unsuitable for compounding with starch as observed on the tensile properties for the melt-blends. In contrast, improvement in Youngs modulus and stress at break were achieved in all blends containing less than 1.5 wt.-% of MA. Melt-blends with 1.0 wt.-% in MA showed larger improvement in break elongation. This could be explained by the improvement in interfacial adhesion between plasticized starch and maleated polyesters as shown by SEM analyses. In melt-blends between TPS and polyesters, some authors[146] have tried to develop chemically modied TPS as a valuable coupling agent with polyesters through a reactive extrusion free-radical grafting process using ricinoleic oxazoline maleate derivatives (Figure 12). Direct insertion of the fatty acid oxazoline moiety into the starch backbone represents a suitable way of hydrophobic modication for the polysaccharide, but also the bifunctionality of the grafting agent provides the opportunity for reactive coupling of the modied starch with hydrophobic polymers such as polycondensates with small amounts of reactive groups, e.g. free carboxylic acid functions. In a mini-scale extruder, the grafting reaction was promoted by bis(a,a-dimethylbenzyl) peroxide as free-radical initiator in high grafting yields (between 1 and 30 wt.-% to starch) for potato starch plasticized with 20 wt.-% in glycerin. From torque measurement, the grafting was completed within 15 min. The presence of oxazoline

www.mme-journal.de

461

J.-M. Raquez, R. Narayan, P. Dubois

Figure 12. Reaction scheme for the radical grafting of starch with ricinoleic oxazoline maleate.

yield an environmentally sound polymer while, at the same time, fullling desired mechanical and cost criteria, such a combination has been difcult to achieve. The reason for this is that the emphaFigure 13. Proposed coupling reaction between oxazoline groups and free acid groups in sis has been on nding the optibiodegradable PBAT. mal polymer or mixture of polymers and other admixtures in moieties onto starch backbone was proved by 1H NMR and order to optimize the properties of the starch/polymer blend thereby. FT-IR spectroscopy. Furthermore, a major issue that most authors have not Subsequently, coupling reactions were carried out from addressed yet, is that the morphology, particularly the size these TPS derivatives with poly[(butylene adipadate)co-terephthalate] (PBAT), biodegradable a,v-hydroxyof dispersed domain, and thus the mechanical properties carboxy-polycondensate, again in a laboratory extruder for the resulting melt-blend of both immiscible partners, is (Figure 13). PBAT is a biodegradable aromatic/aliphatic strongly affected by their difference in melt-viscosity. It copolyester obtained by polycondensation from tereis very critical since starch,[141] and also TPS, exhibits high phthalate acid, adipic acid, and 1,4-butanediol, wherein melt-viscosity related to its high molecular weight the maximum amount of terephthalate acid in copolyester (ranging from 100 000 to 500 000 g mol1 and more than is close to 40 wt.-%, enhancing its mechanical strength, millions, respectively, for the amylose and amylopectin). while retaining the biodegradability of the resulting As a consequence, at least 20% plasticizer is required in the copolyester (Figure 14). extrusion of TPS-based melt-blends at temperatures of ca. Interestingly, a uniform blend could be obtained with 130 8C.[127] TPS chemically modied by 10% ricinoleic oxazoline In contrast to this approach, we have originally prepared maleate derivatives as shown by SEM. In contrast, the through reactive extrusion a novel in situ chemically polymer obtained from untreated TPS exhibits a wellmodied TPS, so called maleated TPS (MTPS) with separated phase. Interestingly, the tensile properties of improved processing and high reactivity in blown lm lms obtained from the PBAT/TPS grafted with 10% applications.[142] As aforementioned, MTPS was prepared, ricinoleic oxazoline maleate derivatives melt-blends through reactive extrusion processing of starch in the (50:50) were enhanced. Biodegradation studies did not presence of glycerol as plasticizer, and of MA as show an impairment of the degradation values compared esterication agent. In addition to derivatization of starch to the unmodied TPS/PBAT melt-blend. backbone with MA, the MA moieties grafted onto the However, although many have attempted for years to starch backbone could promote some hydrolysis and discover the perfect starch/polymer blend that would glucosidation reactions that reduced the relative molecular weight of MTPS. This reduction in molecular weight is a major issue in the melt blending of TPS with its polymeric partner that most of the authors have not addressed. In addition to reduced melt-viscosity, the interest in using MTPS is the presence of free carboxylic acid groups, i.e. MA moieties grafted onto the starch backbone, which could promote Figure 14. Synthesis of PBAT.
Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

462

DOI: 10.1002/mame.200700395

Recent Advances in Reactive Extrusion Processing . . .

acid transesterication reactions with polyester chains such as PBAT, leading to graft copolymers[147149] (Figure 15). We selected PBAT as a good candidate for melt-blending with starch-based products due to its interesting thermal and mechanical properties.[138] As far as MTPS/PBAT melt-blends are concerned, effect of polyester and MA contents was studied on the physico-chemical parameters for the resulting meltblends, again prepared through a downstream extrusion operation. For high polyester fractions, i.e., 60 and 70 wt.-%, PBAT-g-MTPS graft copolymers were obtained through transesterication reactions promoted by the MA-derived acidic moieties grafted onto the starch backbone. This was determined by selective Soxhlet extraction experiments and FT-IR spectroscopy analyses. At lower polyester content, no signicant reaction occurred more likely due to an inversion in phase morphology between PBAT and MTPS. Interestingly, tensile properties of blown lms derived from the PBAT-g-MTPS graft copolymer containing 70 wt.-% polyester, were much higher than those of the melt-blend of TPS and PBAT performed in the presence of MA. This difference in mechanical performances resulted from the low melt-viscosity of MTPS, yielding a ner morphology of the dispersed phase in the continuous PBAT matrix, together with an increased interfacial area for the grafting reaction. This was attested by ESEM. Increased MA content in the preparation of MTPS did not affect the tensile values, suggesting that the entanglement of PBAT and MTPS chains, responsible for these values, did not change after reactive extrusion melt blending. Moreover, WAXS diffraction analyses evidenced that the MTPS crystalline structure was completely disrupted in the PBAT-g-MTPS reactive blends suggesting the grafting

reaction/homogenization of the MTPS in the polyester continuous phase.

Biodegradable Polymeric (Nano)composites Prepared through Reactive Extrusion

By denition, a composite material is formed by the combination of different phases, which have distinct structural and chemical compositions, leading to a synergy of physical, chemical and/or mechanical properties compared to each component taken separately. As reinforcing elements, one can distinguish bers, particles, clays, minerals, and so on. In addition, the matrix can include various types of materials: organic, mineral, and metallic. The classication of composites is based on either the shape of the ller (bers or particles) or on the matrix. In recent years, the use of natural/bio-ber reinforced composites has rapidly expanded due to the availability of natural/bio-ber derived from annually renewable resources, as reinforcing bers in both thermoplastic and thermosetting matrix composites as well as for the positive environmental benets gained by such materials.[150,151] Most of research is also being conducted on the potential of natural bers as reinforcement for polymers, because natural bers have low density, acceptable specic strength properties, easy preparation, and biodegradability. Fully green biocomposites were prepared from PLA and recycled cellulose bers (from newsprint) by extrusion followed by injection molding processing.[17] The physicomechanical and morphological properties of the resulting composites were investigated by varying the amount of cellulose bers. Compared to the neat resin, the tensile and exural moduli of the composites were signicantly higher because of the high modulus provided by cellulose. Increase in stiffness of resulting composites was also conrmed by DMA analyses. The authors claimed that PLA/ natural ber composites have mechanical properties of sufcient magnitude to compete with conventional thermoplastic composites. However, a signicant decrease in tensile strength at high amount Figure 15. Proposed mechanism of transesterication reactions between MTPS and PBAT promoted of bers could be observed, because of the lack of interacby the MA-derived acidic moieties grafted onto the starch backbone.

Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.mme-journal.de

463

J.-M. Raquez, R. Narayan, P. Dubois

tion between cellulose and PLA. Cellulose has a strong hydrophilic character due to three hydroxyl groups per monomeric unit in contrast to the rather hydrophobic PLA. Wood ber was used as a reinforcing ber in the reactive extrusion preparation of corn gluten meal-based composites.[152] Corn gluten meal mainly contains oil (4%), starch (20%), zein protein (6070%), which is an alcohol-soluble protein extracted from corn that has low water resistance. Homogeneous dispersion of wood ber in corn gluten meal could be obtained through reactive extrusion in the presence of glycerol, ethanol, and water as plasticizers. The mixture of ethanol/water was an excellent plasticizer to disrupt the intermolecular interaction of zein, leading to the improvement of melt-mobility. Corn gluten mealbased composites were successfully prepared with 10 50 wt.-% wood ber. It is worth pointing out that the melt-viscosity of the medium increased with increasing wood ber content, and with a decreasing water content, which led to a decrease of melt-mobility. From the exural testing, it has been demonstrated that the exural strengths of these biocomposites increased after the addition of 1030 wt.-% wood ber, but decreased by the addition of 4050 wt.-% wood ber. These results are ascribed to the difference in interfacial tension between wood ber and corn gluten meal. Morphological studies revealed that breaking occurred in the matrix for these bio-composites at high content of wood ber (30 wt.-%), breaking occurred at the interface of the ber and the matrix. Chemically pre-treated natural bers were also employed in reactive extrusion in order to enhance the interfacial adhesion between natural bers and polymers, and hence the properties of resulting composites. Indian grass ber reinforced soy based biocomposites were accordingly fabricated using twin-screw extrusion and injection molding technology.[153] Using extrusion technology, soy protein can be masterfully converted to soy protein plastics.[154] However, soy protein plastic products tend to have lower strength and higher moisture absorption. Currently, biodegradable polymers in the melt-blend of soy protein plastic are used to overcome these drawbacks, including polyester amide and PCL, whose processing windows match that of soy protein plastic. Liu et al. incorporated PBAT with the soy protein polymer to form a soy-based bioplastics.[153] To get higher strength and modulus materials from soy based bioplastics, raw Indian grass and alkaline-treated Indian grass were added as natural bers. Alkaline treatment removes the fraction of lignin contained in this ber, and reduces its size. Tensile and exural properties as well as the heat deection temperature of soy-based bioplastics were improved, but the impact strength of the biocomposites did not improve after reinforcement with raw Indian grass ber. The impact fracture of raw Indian grass ber
Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

reinforced biocomposites was found to occur on the outer surface of the ber, due to intrinsic differences in the morphological structure between the outer and inner surfaces of the grass ber as shown by ESEM. Interestingly, the alkali solution treated Indian grass ber signicantly increased the tensile strength and impact strength as well as the exural strength due to the improved dispersion of the ber in the matrix, and the enhanced aspect ratio of the ber. Other authors have preferred to melt-blend biodegradable polymers and natural bers by adding maleic anhydride-modied polymers as a compatibilizer. Nitz et al. reported the melt-compounding of PCL with wood our and lignin in a Werner & Pfeiderer twin-screw extruder.[155] Wood our contains about 25 wt.-% lignin, which together with cellulose forms the structural component of trees and various plants. As a cheap phenolic biopolymer, lignin offers attractive potential as ller and additive, especially with respect to the modication of biodegradable polymers. Because of the presence of phenolic groups in lignin, it is expected that lignin llers can inuence both oxidative stability and biostability of such compounds. Reactive extrusion technology was carried out to prepare several families of thermoplastic PCL compounds containing wood our and lignin in the presence of MA-grafted PCL as a compatibilizer. In the rst step, grafting of MA onto PCL was performed with a reactive extrusion process in the presence of Luperox 101 as free-radical initiator. Appropriate MA/peroxide ratios led to MA-grafted PCL with a MA content close to 1.44 wt.-% as determined by titration. When a low amount of this MA-grafted PCL (less than 2.5 wt.-% of overall product) was added, attractive properties could be obtained in these PCL-based composites reinforced with wood our and lignin. For a PCL composites containing 2.5 wt.-% in MA-g-PCL and 40 wt.-% wood our, Youngs modulus increased by 450%, and the tensile strength increased by 115% in comparison with the properties of neat PCL. The mechanical properties of the wood our composites were much better as those of the lignin-based compounds. More than 70 wt.-% lignin was added without mechanical properties being impaired, while compositions containing 40 wt.-% lignin showed break elongation exceeding 500%. According to morphological studies, very effective lignin dispersion was achieved within the PCL matrix due to better compatibilization provided by MA-g-PCL. Interestingly, biodegradation studies revealed that the addition of lignin enhanced the biostability of PCL compounds, affording to enhance the lifetime of PCL-based compounds in outdoor applications. Inorganic llers have also been utilized in the preparation of biodegradable polymeric composites. Talc is a common ller used for the improvement of properties of the polymers such as strength, rigidity, durability, and

464

DOI: 10.1002/mame.200700395

Recent Advances in Reactive Extrusion Processing . . .

hardness. Talc has a plate-like geometry in which an edge-shared octahedral sheet of Mg(OH)2 is sandwiched between tetrahedral sheets of silica (SiO2). A bacterial polyester, i.e. PHBV has been melt-compounded with different talc weight contents (1550 wt.-%) through extrusion combined with an injection molding process.[156] PHBV belongs to the family of polyhydroxyalkanoates, which are biocompatible and biodegradable thermoplastics with potential applications in different elds like agriculture, marine, and medicine. The synthesis of polyhydroxyalkanoates occurs normally, when there is an excess of carbon and energy and limitation of at least one nutrient (N, P, Mg, Fe, and so on) needed for microorganism growth. When PHBV was melt-blended with talc, moderate to signicant improvements in the tensile, exural, and storage moduli of talc-lled PHBV were obtained compared to those of neat PHBV. This can be explained by the poor ller dispersion and ller-matrix adhesion as revealed by scanning electron microscopy in the talc-lled PHBV composites. We reported the preparation of new biodegradable and high-performance talc/PBAT hybrid materials through reactive extrusion in blown lms applications.[157] In the rst step, the polyester backbone was reactively modied through free-radical grafting of MA in order to improve the interfacial adhesion between PBAT and talc. The resulting MA-g-PBAT was then reactively melt-blended with talc through esterication reaction of MA moieties grafted onto the polyester chains with the silanol present at the edge surface of talc.[158,159] Sn(Oct)2 and 4-dimethylaminopyridine (DMAP) were studied as esterication catalysts (Figure 16). The interfacial adhesion between both partners was substantially enhanced as evidenced by SEM and selective extraction of the polyester part. As a result, the biaxial tensile properties measured on blown lms prepared from these compatibilized composites were considerably improved, as compared to those of the conventional PBAT-talc melt-blends. Extrapolation to a one-step reactive extrusion process was successfully achieved by preparation of in situ chemically modied PBAT-talc compositions containing up to 60 wt.-% talc. Interestingly, the highest tensile properties were obtained

by melt-blending 50 wt.-% of native PBAT and 50 wt.-% of such a chemically modied PBAT/talc hybrid lled with 60 wt.-% of talc, therefore used as a masterbatch. Such an approach allowed reducing the degradation of native polyester chains through undesirable reactions such as b-scission and transesterication reactions promoted, respectively, by the MA free-radical treatment and esterication catalysts used along the reactive extrusion process. Finally, X-ray photoelectron spectroscopy measurements carried out on the reactively modied PBAT-talc compositions, e.g. containing 60 wt.-% talc, attested for the formation of covalent ester bonds between the silanol functions available at the edge surface of talc particles, and the maleic anhydride moieties grafted onto the polyester backbones. Recently, polymer-layered silicate nanocomposites have emerged as a new class of organic-inorganic materials that have shown unexpected properties such as large increase in the thermal stability, mechanical strength, and impermeability to gases such as water and oxygen. Such improvements in their properties achieved at low content in layered silicate (<5 wt.-%) are relied on the interactions between the clays and polymers, which can yield intercalation and/or exfoliation structures. In the intercalated hybrid structure, a monolayer of extended polymer chains is sandwiched between the silicate sheets, resulting in a well-ordered multilayer of alternating polymer and inorganic sheets. In the exfoliated (or delaminated) nanostructure, the silicate nanolayers are individually dispersed in the polymer matrix. Exfoliation of the silicate layers usually provides nanocomposite materials with the highest improvement in properties aforementioned.[160167] Most of works have focused on the development of (nano)composites based on aliphatic polyesters. Aliphatic polyesters are among the promising materials for the production of high-performance, and environmentfriendly biodegradable plastics. However, gas-barrier properties, melt-viscosity for further processing, etc. are not often sufcient for various end-use applications. Okamoto et al. reported detailed studies about the structure-property relationship in designing desired properties for layered silicate nanocomposites PLA.[168170] These PLA/ layered silicate nanocomposites were prepared by melt-extrusion from montmorillonite organically modied with octadecylammonium (MMT-C18), wherein silicate layers of the clay were intercalated and randomly distributed in the matrix. MMT-C18 was prepared by cation-exchange beFigure 16. Surface grafting mechanism of MA moieties onto the silanol functions at the edge tween octadecylammonium and the naturally occurring Cloisite Na. of talc.

Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.mme-journal.de

465

J.-M. Raquez, R. Narayan, P. Dubois

Incorporation of very small amounts of oligoPCL as compatibilizer led to better parallel stacking of the silicate layers, and also much stronger occulation due to the hydroxylated edge-edge interaction of the silicate layers.[168] The PLA/layered silicate nanocomposites exhibited a remarkable improvement of materials properties in both solid (loss and storage moduli) and melt-states (melt-viscosity) compared to the matrix without clay. Other types of organically modied layered silicates (synthetic uorine mica, synthetic uorine mica modied with N-alkyl-N,N- [bis(2-hydroxyethyl)-N-methylammonium, and smectite) have also been developed by Okamoto et al. in reactive extrusion.[169,170] It has been demonstrated that intercalated, exfoliated, and a mixture of both structures were achieved in the PLA/layered silicate nanocomposites. Interestingly, better dispersion and thus better barrier properties for PLA/layered silicate nanocomposites could be achieved when smectite was used as nanoclays. In addition, all nanocomposites exhibited a remarkable improvement of various materials properties with simultaneous improvement in biodegradability compared to neat PLA. Same efforts were provided by the authors in the preparation of polybutylene succinate (PBS)/layered silicate nanocomposites through an extrusion process.[171173] PBS is chemically synthesized by polycondensation of 1,4-butanediol with succinic acid. In all (nano)composites, intercalated structures were obtained, with a remarkable improvement in tensile properties, thermal stability, and biodegradation of resulting PBS-based (nano)composites. New (nano)hydrids were developed from PCL through reactive extrusion. Organically layered double hydroxide was used as nanoclay.[174] Although a very low content of ller was added into the PCL matrix, no good dispersion, particularly exfoliation was obtained. However, some mechanical and physical properties were improved with respect to neat PCL. Nanocomposites were made from natural biodegradable polymers chemically modied or not. For instance, to limit the retrogradation phenomenon of plasticized starch, green (nano)composites were successfully prepared through reactive extrusion from activated montmorillonite and thermoplastic cornstarch.[175] The thermoplastic cornstarch was plasticized with urea, ethanolamine as plasticizers (one equivalent with respect to starch), but also with natural montmorillonite activated by ethanolamine. Exfoliated structures could be so obtained as attested by WAXS analyses. TEM and SEM images showed that the resulting nanocomposites presented reticulating ber structure after rapid cooling in liquid nitrogen. The mechanical properties of (nano)composites evidently improved such as tensile stress, and Young modulus, but also their thermal stability and water resistance.
Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Green nanocomposites from cellulose acetate were also prepared through reactive extrusion. Eco-friendly triethyl citrate was used as a plasticizer, while Cloisite 30B as organically modied montmorillonite with a methyl tallow bis(2-hydroxyethyl) quaternary ammonium and maleic anhydride grafted cellulose acetate butyrate (CAB-g-MA) were used as organoclay and as compatibilizer, respectively.[176,177] CAB-g-MA was previously prepared through reactive extrusion by grafting MA onto cellulose acetate butyrate as promoted by Luperox 101. The objective of adding CAB-g-MA was that the MA moieties react with the hydroxyl functions from Cloisite 30B. Nanocomposites containing from 0 to 7.5 wt.-% in Cloisite 30B prepared with 5 wt.-% of CAB-g-MA (MA content was of 0.86 wt.-%) exhibited the best morphology, i.e. the complete exfoliation of nanoclay within the matrix, but also the best mechanical properties in terms of tensile and exural properties. In melt-blends derived from hydrophilic plasticized starch and hydrophobic biodegradable polyester, layered silicates were also added to improve the compatibility between both partners. Ikeo at al.[178] developed nanoclay reinforced biodegradable plastics of PCL/plasticized starch blends through reactive extrusion. Starch was rst plasticized with glycerin and water, following by its melt blending with PCL. Maleic anhydride grafted PCL as recovered through reactive extrusion was added as a compatibilizer, enabling to improve the compatibility between thermoplastic starch and PCL. Signicant improvement could be however obtained when natural montmorillonite (Cloisite Na) was added as nanoclay. Interestingly, the nanoclay acted positively in these melt-blends after their electron irradiation as attested by the increase in their modulus and yield stress by more than 50 wt.-% without any reduction in their elongation. In situ REX processing was performed in the preparation of plasticized starch/PCL nanocomposites.[179,180] Native wheat starch (ca. 60 wt.-%), PCL (ca. 40 wt.-%), glycerin as plasticizer, organo-modied montmorillonite, and Fentons reagent (H2O2 and Fe2 from ferrous sulfate) were extruded in a conical twin-screw micro-extruder at 120 8C, and injection-molded at 150 8C. Native starch was partially oxidized by the peroxide, enabling ester groups from PCL to cross-link with carbonyl and/or carboxyl groups as generated from oxidized starch through a peroxide-initiated free process. Ferrous ion catalyzes the decomposition of H2O2 into highly reactive hydroxyl radical that initiates this free-radical chain reaction process. Addition of 3 wt.-% organo-modied clay (MMT-C18) in this chemically modied starch/PCL blends increased elongation almost fourfold over that of unmodied starch/ PCL blends. Better solvent-resistance properties were also achieved.

466

DOI: 10.1002/mame.200700395

Recent Advances in Reactive Extrusion Processing . . .

Biodegradable nanoscale-reinforced starch-based products were also prepared from an in situ chemically modied thermoplastic starch and PBAT through reactive extrusion. Four nanoclays were employed in this study: hydrophilic Cloisite Na, organophilic Cloisite 30B, Bentone 111, and Bentone 166.[181,182] First, thermoplastic starch was in situ chemically modied in the presence of nanoclay previously swollen in glycerol as plasticizer together with MA. Melt blending of these nanoscalereinforced MTPS with PBAT was carried out in the subsequent downstream blending operation. It is worth noting that the swelling treatment was benecial in order to get a better dispersion of nanoclays within the starch-based melt-blends. As shown previously, the MA moieties thus grafted onto the starch backbone were able to promote some glucosidation and hydrolysis reactions during the preparation of MTPS, reducing the molecular weight of native starch. This allows a better interpenetration of the resulting starch ester into the nanoclay galleries, together with a benecial swelling pre-treatment of nanoclay in glycerol. Interestingly enough, the resulting formulations exhibited superior tensile strength and high elongation at break, particularly with Cloisite 30B in blown lm applications. In this case, the presence within the clay galleries of a quaternary ammonium ion bearing two primary hydroxyl groups could form strong hydrogen bond interactions with the MA-derived acidic moieties grafted onto the starch backbone in the MTPS and to some extent, to react with MTPS and PBAT through transesterication reactions promoted by MA. Within the PBAT-g-MTPS graft copolymer-cloisite 30B nanocomposite, WAXS and TEM analyses attested for the partial exfoliation of some clay platelets. As a result, water vapor and oxygen barrier properties of nanoscale-reinforced MTPS-gPBAT nanocompositions were enhanced as compared to the precursors.

Integration of other extrusion streams along with the polymerization process Environmental concerns have triggered many efforts in the development of environmentally friendly plastics, i.e. biodegradable polymers to minimize the waste disposed in landlls. However, the use of these biodegradable polymers as bulk materials is still restricted by their relatively high production cost and poor mechanical properties compared with commodity plastics such as polyethylene. This review has hence outlined the substantial research and the development that have been undertaken in the realm of biodegradable polymers using high-performance continuous reactive extrusion. When combined to both an adapted chemistry such as the right selection of the catalytic system, and a ne selection of extrusion parameters, reactive extrusion has demonstrated a remarkable ability in the synthesis of biodegradable polymers through ring-opening polymerization using effective initiating systems, chemical modication of biodegradable polymers and reactive melt-blending of natural polymers with aliphatic (co)polyesters. Biodegradable polymers-based (nano)composites have effectively been prepared from common llers such as wood ber, but also using layered silicates as nanoclays through reactive extrusion. Whatever reactive melt-blending, biodegradable polymeric melt-blends/composites with satisfactory overall physicomechanical behavior however require the ability to control interfacial tension, to generate a dispersed phase of limited size and strong interfacial adhesion, and to improve the stress transfer between the component phases, i.e. the reinforcement of their interface through the formation of strong covalent bonds. This can be effectively completed using proper interface compatibilization between different components, as reactive extrusion is able to provide during the reactive processing of these biodegradable polymeric melt-blends/composites. The reactive extrusion technology serves on the sustainability and future growth of biodegradable polymers, particularly in the realm of compatibilizing mechanisms, surface modications, and advanced processing techniques, and it is through an understanding of these points that they are expected to replace more and more commodity plastics.

Concluding Remarks and Outlook


Reactive extrusion is a versatile tool for cost-effective polymer processing, which enhances the commercial viability and cost-competitiveness of these materials, in order to carry out melt-blending, and various chemical reactions including polymerization, grafting, branching and functionalization as well. For instance, the obvious advantages of reactive extrusion polymerization process are as follows: Solvent free-melt process Continuous processing, starting from monomer, and resulting in polymer or nished product Control over residence time and residence time distribution
Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Abbreviations
AA AC Al(OiPr)3 Al(OsecBu)3 CAN CHPTMA acrylic acid; acrylamide; aluminum isopropoxide; aluminum sec-butoxide; ceric ammonium nitrate; 3-chloro-2-hydroxypropyltrimethylammonium chloride;

www.mme-journal.de

467

J.-M. Raquez, R. Narayan, P. Dubois

CL CLa DCP DMA DMAP DSC DTMA ESEM FT-IR LA LLa Luperox 101 MA MMT-C18 NaH NMR REX ROP PBAT PBS PBSA P(C6H5)3 PCL PEA PEO PHBV PLA PDX PPDX SD SEM Sn(Oct)2 TriGPC WAXS

e-caprolactone; e-caprolactam; dicumylperoxyde; dynamic mechanical analyzer; 4-dimethylaminopyridine; differential scanning calorimetry; dynamic thermo-mechanical analysis; environmental scanning electronic microscopy; Fourier Transform Infrared; L,L-lactide; v-lauryl lactam; 2,5-dimethyl-2,5-di-(tert-butylperoxy)hexane; maleic anhydride; montmorillonite organically modied with octadecylammonium; sodium hydride; nuclear magnetic resonance; reactive extrusion; ring-opening polymerization; poly[(butylene adipate)-co-terephathalate]; poly(butylene succinate); poly(butylene succinate adipate); triphenyl phosphine; poly(e-caprolactone); polyesteramide; poly(ethylene oxide); poly(hydroxybutyrate-co-valerate); polylactide; 1,4-dioxan-2-one; poly(1,4-dioxan-2-one); substitution degree; scanning electronic microscopy; tin (II) bis(2-ethylhexanoate); triple-detection gel-permeation chromatography; wide angle X-ray scattering.

Acknowledgements: This research was partly funded by Corn Products International. The authors are very grateful to Re gion Wallonne and the European Community (FEDER, FSE) for general support in the frame of Objectif 1-Hainaut: Materia Nova. This work was partly supported by the Belgian Federal Government Ofce of Science Policy (SSTC-PAI 6/27). Received: December 6, 2007; Revised: January 30, 2008; Accepted: January 30, 2008; DOI: 10.1002/mame.200700395 Keywords: biodegradable polymer; chemical modication; meltblending; nanocomposites; reactive extrusion; ring-opening polymerization

[1] R. Mani, M. Battacharya, J. Tang, J. Polym. Sci., Part A: Polym. Chem. 1999, 37, 1693. [2] W. Michaeli, H. Hocker, U. Berghaus, W. Frings, J. Appl. Polym. Sci. 1993, 48, 871. [3] H. Azizi, I. Ghasemi, Polym. Test. 2004, 23, 137. [4] G. Hu, Y. Sun, M. Lambla, Polym. Eng. Sci. 1996, 36, 676. [5] M. Vert, J. Feijen, A.-C. Albertsson, G. Scott, E. Chiellini, Biodegradable Polymers and Plastics, Royal Chemistry Society, Redwood Press, Melksham, Wiltshire 1992. [6] R. Chandra, R. Rustgi, Prog. Polym. Sci. 1998, 23, 1273. [7] C. Reddy, R. Ghai, V. Rashmi, Bioresour. Technol. 2003, 87, 137. [8] R. Gross, Science 2002, 297, 803. [9] S. Karlsson, A.-C. Albertsson, Polym. Eng. Sci. 1998, 38, 1251. [10] ASTM standards on Environmentally Degradable Plastics, ASTM Publication Code Number (PCN): 1993, # 003-420093-19. [11] W. Amass, A. Amass, B. Tighe, Polym. Int. 1998, 47, 89. [12] J. W. Barlow, D. R. Paul, Polym. Eng. Sci. 1984, 24, 525. [13] P. L. Nayak, J. Macromol. Sci. Rev. Macromol. Chem. Phys. 1999, C39, 481. [14] C. Tzoganakis, Adv. Polym. Tech. 1989, 9, 321. [15] S. B. Brown, C. M. Orlando, Reactive Extrusion, in: Encyclopedia of Polymer Science and Engineering, Vol. 14, H. F. Mark, N. M. Bikales, C. G. Overberger, G. Menges, J. I. Kroschwitz, Eds., Wiley, New York 1998, p. 169. [16] P. Cassagnau, V. Bounor-Legare, F. Fenouillot, Internat. Polym. Process. 2007, XXI, 3. [17] J. L. White, B. J. Kim, S. Bawiskar, J. M. Keum, Polym. -Plast. Technol. Eng. 2001, 40, 385. [18] M. Xanthos, Process Analysis from Reaction Fundamentals: Examples of Polymerization and Controlled Degradation in Extruders, in: Reactive Extrusion, M. Xhanthos, Ed., Hanser Publishers, New York 1992, Chapter 14. [19] B. Vergnes, F. Berzin, C. R. Chimie 2006, 9, 1409. [20] S. A. Nied, H. Budman, C. Tzoganakis, Control Eng. Pract. 2000, 8, 911. [21] M. S. Huda, A. K. Mohanty, L. T. Drzal, E. Schut, M. Misra, J. Mat. Sci. 2005, 40, 4221. [22] F. Xie, L. Yu, H. Liu, L. Chen, Starch/Starke 2006, 58, 131. [23] S. Balakrishnan, M. Krishnan, P. Dubois, R. Narayan, Polym. Eng. Sci. 2006, 46, 235. [24] K. E. Oliphant, K. E. Russel, W. E. Baker, Polymer 1995, 36, 1597. [25] K. Kelar, B. Jurkowski, Polymer 2000, 41, 1055. [26] D. Shi, J. Yang, Z. Yao, Y. Wang, H. Huang, W. Jing, J. Yin, G. Costa, Polymer 2001, 42, 5549. [27] M. Coltelli, E. Passaglia, F. Ciardelli, Polymer 2006, 47, 85. [28] L. Wu, Y. Jia, S. Sun, G. Zhang, G. Zhao, L., Mater. Sci. Eng. 2007, A454455, 221. [29] J.-P. Puaux, P. Cassagnau, G. Bozga, I. Nagy, Chem. Eng. Process. 2006, 45, 481. [30] S. S. Pesetskii, B. Jurkowski, Y. M. Krivoguz, K. Kelar, Polymer 2001, 42, 469. [31] G. Moad, Prog. Polym. Sci. 1999, 24, 81. [32] E. Chiellini, P. Cinelli, F. Chiellini, S. Imam, Macromol. Biosci. 2004, 4, 218. [33] M. Alexandre, P. Dubois, Mater. Sci. Eng. 2000, 28, 1. [34] G. Kale, T. Kijchavengkul, R. Auras, M. Rubino, S. Selke, S. Singh, Macromol. Biosci. 2007, 7, 255. [35] M. Yasin, B. Tighe, Biomaterials 1992, 13, 9. [36] P. Matzinos, V. Tserki, P. Gianikouris, C. Panayiotou, Eur. Polym. J. 2002, 38, 1713.

468

Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

DOI: 10.1002/mame.200700395

Recent Advances in Reactive Extrusion Processing . . .

[37] E. Corradini, L. Mattoso, C. Guedes, D. Rosa, Adv. Polym. Technol. 2004, 15, 340. [38] J. Seppala, A. O. Helminen, H. Korhonen, Macromol. Biosci. 2004, 4, 208. [39] M. Widmer, P. Gupta, L. Lu, R. Meslenyi, G. Evans, K. Brandt, T. Savel, A. Gurlek, C. Patrick, A. Mikos, Biomaterials 1998, 19, 1945. [40] J.-M. Raquez, P. Degee, R. Narayan, P. Dubois, Macromol. Rapid Comm. 2000, 21, 1063. [41] J.-M. Raquez, P. Degee, R. Narayan, P. Dubois, Macromol ecules 2001, 34, 8419. [42] J.-M. Raquez, P. Degee, R. Narayan, P. Dubois, Macromol. Chem. Phys. 2004, 205, 1764. [43] J.-M. Raquez, P. Degee, R. Narayan, P. Dubois, Polym. Degrad. Stab. 2004, 86, 159. [44] R. Bezwada, D. Jamiolkowski, K. Cooper, Handbook of Biodegradable Polymers, A. Domb, J. Kost, D. Wiseman, Eds., Harwood academic publishers, Newark 1997, p. 29. [45] D. Mecerreyes, R. Jerome, Macromol. Chem. Phys. 1999, 200, 2581. [46] C. Jacobs, P. Dubois, R. Jerome, P. Teyssie, Macromolecules 1991, 24, 3027. [47] P. Dubois, R. Jerome, P. Teyssie, Polym. Bull. 1989, 22, 475. [48] I. Barakat, P. Dubois, R. Jerome, P. Teyssie, E. Goethals, J. Polym. Sci.,Part A: Polym. Chem. 1993, 31, 505. [49] S. Penczek, A. Duda, A. Kowalski, J. Libiszowski, M. Katarzyna, T. Biela, Macromol. Symp. 2000, 157, 61. [50] Y. Du, P. Lemstra, A. Nijenhuis, H. Van Aert, C. Bastiaansen, Macromolecules 1995, 28, 2124. [51] H. Kricheldorf, I. Kreiser-Saunders, C. Boettcher, Polymer 1995, 36, 1253. [52] A. Kowalski, A. Duda, S. Penczek, Macromol. Rapid Comm. 1998, 19, 567. [53] J.-M. Raquez, P. Degee, Y. Nabar, R. Narayan, P. Dubois, C. R. Chimie 2006, 9, 1370. [54] D. Perrin, J. English, Handbook of biodegradable polymers, Harwood academic publishers, Newark 1997, p. 291. [55] B. J. Kim, J. L. White, J. Appl. Polym. Sci. 2004, 94, 1007. [56] B. J. Kim, J. L. White, Int. Polym. Process 2002, 17, 33. [57] L. Zhu, K. Narth, K. Hyun, Int. J. Heat Mass Transfer 2005, 48, 3411. [58] S. Choulak, F. Couenne, Y. Le Gorrec, C. Jallut, P. Cassagnau, A. Michel, Ind. Eng. Chem. Res. 2004, 43, 7373. [59] A. Poulesquen, B. Vergnes, P. Cassagnau, J. Gimenez, A. Michel, Int. Polym. Process. 2001, 16, 31. [60] A. Gaspar-Cunha, A. Poulesquen, B. Vergnes, J. A. Covas, Int. Polym. Process. 2002, 17, 201. [61] J. Gimenez, M. Boudris, P. Cassagnau, A. Michel, Polym. React. Eng. 2000, 8, 135. [62] P. Cassagnau, J. Gimenez, V. Bounor-Legare, A. Michel, C. R. Chimie 2006, 9, 1351. [63] P. Cassaganau, V. Bounor-Legare, F. Fenouillot, Intern. Polym. Process. 2007, XXI, 3. [64] D. Mecerreyes, R. Jerome, P. Dubois, Adv. Polym. Sci. 1999, 147, 1. [65] H. Von Schenck, M. Ryner, A. Albertsson, M. Svensson, Macromolecules 2002, 35, 1556. [66] B. J. Kim, J. L. White, J. Appl. Polym. Sci. 2005, 96, 1875. [67] B. J. Kim, J. L. White, J. Appl. Polym. Sci. 2003, 88, 1437. [68] I. Goodman, R. Vachon, Eur. Polym. J. 1984, 20, 529. [69] W. Tang, S. Murthy, F. Mares, M. McDonnell, S. Curran, J. Appl. Polym. Sci. 1999, 74, 1858. [70] P. Degee, P. Dubois, R. Jerome, S. Jacobsen, H. G. Fritz, Macro mol. Symp. 1999, 144, 289.

[71] P. Degee, P. Dubois, R. Jerome, S. Jacobsen, H. G. Fritz, Macro mol. Symp. 2000, 153, 261. [72] P. Degee, P. Dubois, R. Jerome, S. Jacobsen, H. G. Fritz, Polymer 2000, 41, 3395. [73] S. Jacobsen, P. Degee, H. G. Fritz, P. Dubois, R. Jerome, Polym. Eng. Sci. 1999, 39, 1311. [74] S. Jacobsen, P. Degee, H. G. Fritz, P. Dubois, R. Jerome, Ind. Crops Prod. 2000, 11, 265. [75] A. Sodergard, J. H. Nasman, Polym. Degrad. Stab. 1994, 46, 25. [76] S. Gogolewski, M. Javanovic, S. M. Perren, J. G. Dillon, M. K. Hughes, Polym. Degrad. Stab. 1993, 40, 313. [77] G. Schmack, B. Tandler, R. Vogel, R. Beyreuther, S. Jacobsen, H. G. Fritz, J. Appl. Polym. Sci. 1999, 73, 2785. [78] G. Schmack, B. Tandler, R. Vogel, R. Beyreuther, S. Jacobsen, H. G. Fritz, J. Biotech. 2001, 86, 151. [79] W. M. Stevels, A. Bernard, P. Van de Witte, P. Dijkstra, J. Feijen, J. Appl. Polym. Sci. 1996, 62, 1295. [80] A. K. Mohanty, M. Misra, G. Hinrichsen, Macromol. Mater. Eng. 2000, 276/277, 1. [81] G. Della, P. Valle, J. Taybeb. Colonna, Starch/Starke 1991, 43, 300. [82] M. E. Carr, J. Appl. Polym. Sci. 1994, 54, 1855. [83] M. E. Carr, S. Kim, J. Yoon, K. D. Stanley, Cereal Chem. 1992, 69, 70. [84] A. Ach, J. M. S. -Pure Appl. Chem. 1993, A30, 733. [85] Y. Miyashita, T. Suzuki, Y. Nishio, Cellulose 2002, 9, 215. [86] R. Stilwell, M. Marks, L. Saferstein, D. Wiseman, Handbook of biodegradable polymers, Harwood academic publishers, Newark 1997, 291. [87] C. Buchanan, R. Gardner, R. Komarek, J. Appl. Polym. Sci. 1993, 47, 1709. [88] A. K. Mohanty, A. Wibowo, M. Misra, L. T. Drzal, Polym. Eng. Sci. 2003, 43, 1151. [89] H. Warth, R. Mulhaupt, J. Schatzle, J. Appl. Polym. Sci. 1997, 64, 213. [90] P. Tomasik, P. Wang, J. Jane, Starch/Starke 1995, 47, 96. [91] US Patent 2,461,139 (1949), C. G. Caldwell, F. Hills. [92] US Patent 2,661,349 (1953), invs.: C. G. Caldwell, F. Hills, O. B. Wurzburg. [93] J. Radley, Industrial Uses of Starch and Its Derivatives, Applied Science Publishers, London 1976, p. 1 976. [94] V. D. Miladinov, M. A. Hanna, Ind. Crops Prod. 2000, 11, 51. [95] US patent application 2006111511 (2006), invs.: Y. Nabar, J.-M. Raquez, P. Degee, R. Narayan, P. Dubois. [96] F. J. Rodriguez-Gonzalez, B. A. Ramsay, B. D. Favis, Carbohydrat. Polym. 2004, 58, 139. [97] X. Mo, X. Sun, J. Am. Oil Chem. Soc. 2002, 79, 197. [98] S. Wang, H. Sue, J. Jane, J. Macromol. Sci. Pure Appl. Chem. 1996, A33, 557. [99] P. Chen, L. Zhang, Biomacromolecules 2006, 7, 1700. [100] C. Vaz, P. Van Doeveren, G. Ylmaz, L. de Graaf, R. Leis, A. Cunha, J. Appl. Polym. Sci. 2005, 97, 604. [101] C. Vaz, P. Van Doeveren, R. Leis, A. Cunha, Biomacromolecules 2003, 4, 1520. [102] D. Sessa, R. Wing, Ind. Crops Prod. 1999, 10, 55. [103] I. Taniguchi, W. A. Kuhlman, A. M. Mayes, L. G. Grifth, Polym. Int. 2006, 55, 1385. [104] D. Carlson, P. Dubois, L. Nie, R. Narayan, J. Appl. Polym. Sci. 1999, 72, 477. [105] O. B. Wurzburg, Converted starch in Modied Starch: Properties and Uses, O. B. Wurzburg, Ed., CRC Inc. Boca Raton, FL 1986. [106] R. E. Wing, J. L. Willet, Ind. Crops Prod. 1997, 7, 45. [107] L. Yu, K. Dean, L. Li, Prog. Polym. Sci. 2006, 31, 576.

Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.mme-journal.de

469

J.-M. Raquez, R. Narayan, P. Dubois

[108] US Patent 5,288,765 (1994), invs.: C. Bastioli, V. Bellotti, L. Del Giudice, R. Lombi, A. Rallis, G. Del Tredici, A. Montino, R. Ponti. [109] US Patent 5,360,830 (1994), invs.: C. Bastioli, V. Bellotti, L. Del Giudice, R. Lombi, A. Rallis. [110] US Patent 5,736,586 (1998), invs.: C. Bastioli, V. Bellotti, G. Del Tredici, A. Montino, R. Ponti. [111] US Patent 5,801,207 (1998), invs.: C. Bastioli, V. Bellotti, G. Del Tredici, A. Rallis. [112] US Patent 5,185,382 (1993), invs.: P. E. Neumann, P. A. Seib. [113] US Patent 5,208,267 (1993), invs.: P. E. Neumann, P. A. Seib. [114] US Patent 5,665,786 (1997), invs.: W. Xu, W. M. Doane. [115] US Patent 5,854,345 (1998), invs.: W. Xu, W. M. Doane. [116] R. L. Shogren, Carbohydrate. Polym. 1996, 29, 57. [117] L. Wang, R. L. Shogren, Proceedings 6th Annual Meeting of the Bio/Environmental Degradable Polymer Society, The Society: St. Paul, MN 1997. [118] J. L. Willett, R. L. Shogren, Polymer 2002, 43, 5935. [119] Q. Fang, M. A. Hanna, Cereal Chemistry 2000, 77, 779. [120] Q. Fang, M. A. Hanna, Bioresource Technology 2001, 78, 115. [121] US Patent 6,184,261 (2001), invs.: G. Biby, M. A. Hanna, Q. Fang. [122] Q. Fang, M. A. Hanna, Ind Crops Prod 2001, 13, 219. [123] V. Khunova, J. Hurst, I. Janigova, V. Smatko, Polym. Test. 1999, 18, 501. [124] D. J. Lohse, S. Datta, Polymeric Materials Encyclopedia, CRC Press, Boca Raton FL 1996. [125] C. W. Macosko, H. K. Jeon, T. R. Hoye, Prog. Polym. Sci. 2005, 30, 939. [126] R. F. Septo, Macrompl. Symp. 2003, 201, 203. [127] J. Van Soest, K. Benes, D. de Witt, Polymer 1996, 37, 3543. [128] P. Dubois, R. Narayan, Macromol. Symp. 2003, 198, 233. [129] R. de Graaf, A. Karman, L. Janssen, Starch/Starke 2003, 55, 80. [130] R. Souza, C. Andrade, Adv. Polym. Tech. 2002, 21, 17. [131] J. H. Yang, J. G. Yu, X. F. Ma, Starch/Starke 2006, 58, 330. [132] X. Ma, J. Yu, Starch/Starke 2004, 56, 545. [133] J. Guan, M. Hanna, Ind. Eng. Chem. Res. 2006, 45, 3991. [134] V. Miladinov, M. A. Hanna, Ind. Eng. Chem. Res. 1999, 38, 3892. [135] J. Guan, A. Hanna, Biomacromolecules 2004, 5, 2329. [136] J. J. G. Van Soest, N. Knooren, J. Appl. Polym. Sci. 1997, 64, 1411. [137] Y. Nabar, D. Draybuck, R. Narayan, J. Appl. Polym. Sci. 2006, 102, 58. [138] Y. Nabar, J.-M. Raquez, P. Dubois, R. Narayan, Biomacromolecules 2005, 6, 807. [139] L. Averous, L. Moro, P. Dole, C. Fringant, Polymer 2000, 41, 4157. [140] O. Martin, E. Schwach, L. Averous, Y. Couturier, Starch/Starke 2001, 53, 372. [141] J. Wu, W.-C. Lee, W.-F. Kuo, H.-C. Kao, M.-S. Lee, J.-L. Lin, Adv. Polym. Tech. 1995, 14, 47. [142] J.-M. Raquez, Y. Nabar, P. Dubois, R. Narayan, Polym. Eng. Sci., 2008, in press. [143] P. Dais, A. Perlin, Carbohydrate Res. 1982, 100, 103. [144] M. Gidley, Carbohydrate Res. 1985, 139, 85. [145] R. B. Maliger, S. A. McGlashan, P. J. Halley, L. G. Matthew, Polym. Eng. Sci. 2006, 46, 248. [146] B. Kosan, F. Meister, T. Liebert, T. Heinze, Cellulose 2006, 13, 105.

[147] C. K. Chong, J. Xing, D. Philipps, H. Corke, J. Agric. Food Chem. 2001, 49, 2702. [148] M. Carr, J. Appl. Polym. Sci. 1991, 42, 45. [149] A. Sarko, C. Chen, B. J. Hardy, ACS Symposium series 1990, 430, 345. [150] A. K. Mohanty, M. Misra, T. Drzal, J. Polym. Env. 2002, 10, 19. [151] P. Kulpinski, J. Appl. Polym. Sci. 2007, 104, 398. [152] Q. Wu, H. Sakabe, S. Isobe, Ind. Eng. Chem. Res. 2003, 42, 6765. [153] W. Liu, A. K. Mohanty, L. T. Drzal, M. Misra, Ind. Eng. Chem. Res. 2005, 44, 7105. [154] D. Graiver, L. H. Waikul, C. Berger, R. Narayan, J. Appl. Polym. Sci. 2004, 92, 3231. [155] H. Nitz, H. Semke, R. Landers, R. Mulhaupt, J. Appl. Polym. Sci. 2001, 81, 1972. [156] A. Whaling, R. Bhardwaj, A. K. Mohanty, Ind. Eng. Chem. Res. 2006, 45, 7497. [157] J.-M. Raquez, Y. Nabar, P. Dubois, R. Narayan, Macromol. Mater. Eng., DOI:10.1002/mame.200700352 [158] J. Temuujin, K. Okada, T. S. Jadambaa, K. J. D. Mackenzie, J. Amarsanaa, J. Mat. Sci. Let. 2002, 21, 1607. [159] V. C. Farmer, The infrared Spectra of Minerals, Mineralogical Society, London 1974. [160] R. Schollhorn, Chem. Mater. 1996, 8, 1747. [161] P. B. Messersmith, S. I. Stupp, J. Mater. Res. 1992, 7, 2599. [162] A. Okada, A. Usuki, Mater. Sci. Eng. 1995, C3, 109. [163] E. Giannelis, Adv. Mater. 1996, 8, 29. [164] Z. Wang, J. Massam, T. Pinnavaia, Epoxy - Clay Nanocomposites, in: Polymer-Clay Nanocomposites, T. J. Pinnavaia, G. Beall, Eds., Wiley, Indiana 2000, pp. 127. [165] R. Bharadwaj, Macromolecules 2001, 34, 9189. [166] D. Chaiko, A. Leyva, Chem. Mater. 2005, 17, 13. [167] J. Lange, Y. Wyser, Packag. Technol. Sci. 2003, 16, 149. [168] S. S. Ray, P. Maiti, M. Okamoto, K. Yamada, K. Ueda, Macromolecules 2002, 35, 3104. [169] S. S. Ray, K. Yamada, M. Okamoto, A. Ogami, K. Ueda, Chem. Mater. 2003, 15, 1456. [170] P. Maiti, K. Yamada, M. Okamoto, K. Ueda, K. Okamoto, Chem. Mater. 2002, 14, 4654. [171] S. S. Ray, K. Okamoto, M. Okamoto, Macromolecules 2003, 36, 2355. [172] S. S. Ray, K. Okamoto, M. Okamoto, J. Polym. Sci., Part B: Polym. Phys. 2003, 41, 3160. [173] S. S. Ray, K. Okamoto, M. Okamoto, Macromol. Mater. Eng. 2005, 290, 759. [174] R. Pucciariello, L. Tammaro, V. Villani, V. Vittoria, J. Polym. Sci., Part B: Polym. Phys. 2007, 45, 945. [175] M. Huang, J. Yu, J. Appl. Polym. Sci. 2006, 99, 170. [176] H. Park, X. Lang, A. K. Mohanty, M. Misra, L. Drzal, Macromolecules 2004, 37, 9076. [177] H. Park, M. Misra, L. Drzal, A. K. Mohanty, Biomacromolecules 2004, 5, 2281. [178] Y. Ikeo, K. Aoki, H. Kishi, S. Matsuda, A. Murakami, Adv. Polym. Tech. 2006, 17, 940. [179] S. Kalambur, S. Rizvi, Polym. Int. 2004, 53, 1413. [180] S. Kalambur, S. Rizvi, J. Appl. Polym. Sci. 2005, 96, 1072. [181] J.-M. Raquez, Y. Nabar, P. Dubois, R. Narayan, Int. Polym. Process. 2007, XXI, 5. [182] J.-M. Raquez, Y. Nabar, P. Dubois, R. Narayan, Polymer 2008, submitted.

470

Macromol. Mater. Eng. 2008, 293, 447470 2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

DOI: 10.1002/mame.200700395

You might also like