Вы находитесь на странице: 1из 11

Recent Developments in Biopolymers

Vicki Flaris,1 Gurpreet Singh2 1 Department of Chemistry, BCC of CUNY, University Ave. and West 181st St., Bronx, New York 10453-3102
2

Department of Polymer Engineering and Technology, Institute of Chemical Technology, University of Mumbai, [UICT, formerly UDCT], Mumbai-400019, India

Society has been reaping the benets of industrial polymers for a long time. Polymers have entered every market in a very inuential manner, from the packaging industry to the construction business. The very properties that made polymers commercially viable are posing great environmental problems for our future generations. Also, the starting material for most of the commercial polymers is crude oil. Thus, environmental issues coupled with decreasing crude oil reserves have forced the polymer industry to nd new sources. These problems having been taken into consideration, biopolymers have emerged as a promising eld. This paper takes into consideration the sources of renewable materials, such as starch, lignocellulosic biomass, vegetable oils, proteins, etc.; the synthesis of polymers such as polylactic acid and monomers such as furfural, ethane, propanediol, etc., from renewable materials; and the recent developments in this eld. J. VINYL
ADDIT. TECHNOL., 15:111, 2009. 2009 Society of Plastics Engineers

products have proven disastrous for the natural polymers market. It is only after a lapse of almost 50 years that the signicance of eco-friendly materials has been realized once again. Polymers from renewable resources have attracted an increasing amount of attention over the last two decades, predominantly for two major reasons: rst, environmental concerns, and second, the realization that our petroleum resources are nite [6]. The idea of being able to convert plants, crops, and other natural things around us into polymers used to make everything from packaging and consumer goods to bers for apparel and furnishings will not only reduce our dependence on petroleum but also will reduce pollution [7]. Generally, polymers from renewable resources can be classied into three groups [8]:
1. Natural polymers, such as starch, protein, and cellulose. 2. Synthetic polymers from natural monomers, such as poly(lactic acid) (PLA). 3. Polymers from microbial fermentation, such as poly (hydroxybutyrate) (PHB).

INTRODUCTION The environment is being overwhelmed by nonbiodegradable, petroleum-based polymeric materials. The increasing demands for such materials have increased the dependence on crude oil and left the highways, beaches, and landlls overowing with these nonrenewable, indestructible materials. In contrast, current interest in cheap, biodegradable polymeric materials has recently encouraged the development of such materials from readily available, renewable, inexpensive natural sources, such as carbohydrates, starch, and proteins [14]. The study and utilization of natural polymers is an ancient science. Typical examples, such as paper, silk, skin, and bone arts, can be easily found in museums around the world. However, the availability of petroleum at a lower cost [5] and the biochemical inertness of petroleum-based
Correspondence to: Vicki Flaris; e-mail: vicki.aris@bcc.cuny.edu DOI 10.1002/vnl.20171 Published online in Wiley InterScience (www.interscience.wiley.com). 2009 Society of Plastics Engineers

THE NEED FOR POLYMERS FROM RENEWABLE SOURCES The discovery of petroleum in the 1930s in the United States revolutionized the chemical industry. Today, chemistry based on petroleum is well-understood and the petroleum industry is a well-geared machine. However, with the ever-increasing demand for petroleum-based products, it has been estimated that we will deplete petroleum reserves by the end of the present century and see a signicant shortage and signicant increase in crude oil costs as early as 2040 [9]. Now, from an environmental perspective, even though the vast majority of fossil resources consumed today is used for energy purposes, such as for heating and transport, a signicant share, 1112% of the crude oil, is used for nonenergy applications, such as in the production of polymers [4, 7]. This consumption leads to a similar

JOURNAL OF VINYL & ADDITIVE TECHNOLOGY 2009

transfer of carbon from fossil deposits to the atmosphere as does the combustion of fossil fuels for energy. Through replacing fossil raw materials with renewable raw materials, the transfer of carbon from the fossil deposits could be reduced. Intermediate chemicals play an integral role in the world economy. Organic chemicals are synthesized primarily from petroleum for the production of numerous nonfuel industrial products, such as plastics, plastics additives, clothing, synthetic bers, and paints [6, 10, 11]. Biobased feedstocks such as trees, grasses, agricultural crops, agricultural residues, animal wastes, and municipal solid waste can be converted into these products. This conversion will reduce or even eliminate our dependence on the depleting petroleum reserves. Recently, hybrid resins that are blends of polypropylene (PP) and starch have been commercialized by Cereplast. Essentially, the carbon footprint of these resins is much better than that of 100% fossil-fuel-based PP.

FIG. 1.

a-Amylose.

Materials from biomass which can be used as renewable raw materials are given below:

Agricultural residues: Straws, corn stalks and cobs, bagasse, cotton gin trash, palm oil wastes. Crops grown specically for their biomass: Grasses, sweet sorghum, fast-growing trees. Paper: Recycled newspaper, paper mill sludges, sorted municipal solid waste. Wood wastes: Prunings, wood chips, sawdust. Green wastes: Leaves, grass clippings, vegetable and fruit wastes.

SOURCES OF RENEWABLE RAW MATERIALS The biological sciences, when combined with recent and future advances in process engineering, can become the foundation for producing a wide variety of industrial products from renewable plant resources. These opportunities begin with the plant sources for raw materials. Modern principles of molecular biology and genetic engineering can be used to create agricultural crops that contain desired chemical polymers or polymer intermediates. Additionally, trees and grasses could be genetically engineered to have a structural composition that facilitates and enhances the effectiveness and efciency of subsequent conversion into desired products [10]. Renewable raw materials are in this context dened as products derived from the agricultural and forestry sectors being used for other purposes than nutritionally. Thus, the organic carbon used in the manufacturing of products is regenerated each year. The carbon that is released to the atmosphere when the products are disposed of has been xated from the atmosphere during plant growth. This is what is called closing the carbon cycle. Thus, renewable raw materials are a wide-ranging and diverse group of raw materials but can broadly be divided into oil, carbohydrate, and ber crops. Their common denominator is that they are derived from the agricultural and forestry sectors and are applied in the manufacturing industry. They renew themselves, annually or over longer time perspectives, and thus, provided that they are cultivated and harvested in a responsible manner, provide more sustainable means to cover our needs for raw materials than do the fossil raw materials. Biomass, which is composed of carbohydrates, can be used to produce the products that are commonly manufactured from petroleum and natural gas, or hydrocarbons [10, 12]. Both resources contain the essential elements of carbon and hydrogen.

Of all these materials, some were recognized for their important properties, and they are described below with respect to their composition and conversion issues, if any. Biomass can be differentiated as follows [10]:

Starch. Lignocellulosic biomass. Vegetable oils and proteins.

Starch Starch is composed of glucose, but it is a mixture of aamylose and amylopectin. a-Amylose (Fig. 1) is a straight chain of glucose molecules joined by a-1,4-glycosidic linkages. Amylopectin (Fig. 2) is similar to amylase, except that short chains of glucose molecules branch off from the main chain (backbone). Starches found in nature are 1030% a-amylose and 7090% amylopectin. The a1,4-glycosidic linkages are relatively bent and thus prevent the formation of sheets and the subsequent layering of polymer chains. As a result, starch is soluble in water and relatively easy to break down into utilizable sugar units. Corn grain serves as the primary feedstock for starch used to manufacture the biobased products of today [8, 10]. Lignocellulosic Biomass The nongrain portion of biomass (e.g., cobs, stalks), often referred to as agricultural stover or residues, and energy crops such as switchgrass also contain valuable components, but they are not as readily accessible as starch. These lignocellulosic biomass resources (also called cellulosic) are comprised of cellulose, hemicellulose, and lignin. Generally, lignocellulosic material contains 3050% of cellulose, 2030% of hemicellulose, and 2030% of lignin. Some exceptions to this composition are cotton (98% of

2 JOURNAL OF VINYL & ADDITIVE TECHNOLOGY 2009

DOI 10.1002/vnl

FIG. 2. Amylopectin.

cellulose) and ax (80% of cellulose). Lignocellulosic biomass is perceived as a valuable and largely untapped resource for the future bioindustry. However, recovering the components in a cost-effective way represents a signicant technical challenge [10, 12, 13]. The components of lignocellulosic biomass are described below: Cellulose. Cellulose (Fig. 3) is one of the polymers of nature and is composed of glucose, a six-carbon sugar. The glucose molecules are joined by b-1,4-glycosidic linkages which allow the glucose chains to assume an extended ribbon conformation. The only structural difference from linear starch is b-1,4 links instead of a-1,4 links, but it makes a world of difference in properties. Cellulose has very good strength. Its links are broken by cellulase enzymes that are far less common in nature than the amylase enzymes that split starch. Hydrogen bonding between chains leads to the formation of at sheets that lie on top of one another in a staggered fashion, similar to the way staggered bricks add strength and stability to a wall. As a result, cellulose is very stable chemically and insoluble, and it serves as a structural component in plant walls [8, 10, 12, 14, 15]. Lignocellulosic materials as they are found in nature are much more resistant than starch to acid and enzymatic hydrolysis owing to the high degree of crystallinity of cellulose. As the core structural component of biomass, cellulose is protected from environmental exposure by a sheath of lignin and hemicellulose. Harnessing the sugars of lignocellulosics, therefore, involves a pretreatment stage to reduce the resistance of the biomass to cellulose

hydrolysis. The effectiveness of the pretreatment stage is the key to successful conversion into sugars [10]. Hemicellulose. Hemicellulose (Fig. 4) is a polymer of ve different sugars. It contains ve-carbon sugars (usually D-xylose and L-arabinose) (Fig. 5), six-carbon sugars (D-galactose, D-glucose, and D-mannose), and uronic acid. A short-chain polymer is formed which interacts with cellulose and lignin to form a matrix in the plant wall, thus strengthening it. Hemicellulose is easily hydrolyzed. Most of the hemicellulose in lignocellulosic materials is solubilized and hydrolyzed into pentose and hexose sugars during the pretreatment stage. Some of the hemicellulose is too intertwined with the lignin and is therefore unrecoverable [10]. Lignin. Lignin (Fig. 6) is an amorphous aromatic natural biopolymer that provides rigidity to plant bers [16]. Lignin helps to bind the cellulose/hemicellulose matrix while adding exibility to the mix. It is considered as the greatest store of aromatic molecules in nature. The molecular structure of lignin polymers is very random and disorganized and consists primarily of carbon-ring structures (benzene rings with methoxy, hydroxy, and propyl groups) interconnected by polysaccharides (sugar polymers). The ring structures of lignin have great potential as valuable chemical intermediates. However, separation and recovery of the lignin are difcult [10].

FIG. 3. Cellulose.

FIG. 4. Hemicellulose.

DOI 10.1002/vnl

JOURNAL OF VINYL & ADDITIVE TECHNOLOGY 2009 3

rials for lubricants, hydraulic uids, polymers, and a host of other products [10]. Vegetable Oils. Vegetable oils are composed primarily of triglycerides, also referred to as triacylglycerols. To use these compounds as starting materials for polyurethane synthesis, it is necessary to functionalize them to form polyols. Epoxidation and ring-opening reactions with haloacids or alcohols, ozonolysis, and hydration are some of the common methods for the functionalization of unsaturated vegetable oils. Triglycerides contain a glycerol molecule as the backbone with three fatty acids attached via the glycerol hydroxy groups. Fatty acids differ in their chain length as well as their degree of unsaturation. The fatty acid prole and the double bonds present determine the properties of the oil. These features can be manipulated to obtain certain performance characteristics. In general, the greater the number of double bonds, the lower the melting point. Chemicals which can be obtained from vegetable oils are

FIG. 5.

D-xylose

and L-arabinose.

Oils and Proteins Oils and proteins are offered by the seeds of certain plants. Oils and proteins have great potential for bioproducts. They are found in the seeds of plants such as soybeans, castor beans, cashew nuts, etc., and can be extracted in a variety of ways. Plants raised for this purpose include soy, corn, sunower, safower, rapeseed, and others. A large portion of the oil and protein recovered from oilseeds and corn is processed for human or animal consumption, but they can also serve as raw mate-

FIG. 6. Lignin.

4 JOURNAL OF VINYL & ADDITIVE TECHNOLOGY 2009

DOI 10.1002/vnl

FIG. 8. Cross section of a corn kernel.

Conversion of Starch (Corn) into PLA


FIG. 7. Synthetic route for production of various polymers from vegetable oils.

glycerol, alkyd resins, fatty oils, fatty acids, polyurethanes, and many others. One of the synthetic routes by which such products can be produced from vegetable oils is shown in Fig. 7. Vegetable oils useful for the polymer industry include soybean, sunower, safower, rapeseed, crambe, castor, lesquerella, jojoba, and meadow foam oil [10]. Proteins. Proteins have very unique crystalline structures that can provide unmatched performance as materials. Spider silk, for example, has ber strength and elasticity properties that no synthetic ber approaches. Certain aquatic organisms produce proteins with unmatched tenacious adhesive properties. Utilizing and modifying natural proteins for novel materials holds tremendous promise but is an area that has barely been explored [10].

Corn is the rst crop to be used commercially in the manufacture of polymers. This process is commercialized by Cargill Dow, which makes PLA from corn under the trade name NatureWorks. The cross section of a corn kernel is shown in Fig. 8 [17]. Standard eld corn (maize) is used today to make PLA because of its cost and abundance. In the future, PLA and many other polymers could be manufactured from any plant material from which sugar can be harvested. The typical properties of PLA are compared with those of some common synthetic polymers in Table 1. Let us now take a look at the process of converting corn sugar into PLA. Production of Starch by Plants. Plants trap solar energy through chloroplasts. Cells inside chloroplasts create sugar and oxygen. This sugar is used by plants as fuel. The unused sugar is stored as starch in the kernel. Conversion of Corn into PLA. Harvested corn is sent for milling in a milling plant. The plant is cooked, a process which causes it to swell and soften. The corn cooks for 30 to 40 hours at 122 degrees Fahrenheit. Leftover water is used in the production of animal feed. Machines grind and screen the corn mixture to isolate the starch. Now the starch is converted into sugar. Microorganisms

SYNTHESIS OF POLYMERS/MONOMERS FROM RENEWABLE RAW MATERIALS Let us take a look at how some of these raw materials are converted into polymers or their monomers or other chemicals that are useful for polymer synthesis.

TABLE 1. Typical properties of PLA and other synthetic polymers. Property Specic gravity (g/cc) MFI (g/10 min) Flexural modulus (MPa) Tensile strength (MPa) Tensile elongation (%) Notched Izod impact (J/m) PP-homopolymer 0.8990.948 0.2749 9261850 31.140.8 1.026 5.3447.9 PET 1.321.41 15603280 23250 2.5210 14.080.4 PC 1.161.21 3.826 19002660 52.570 0120 0961 PLA 1.231.26 3.025 3363830 13.0112 2.010.0 12.829.0

DOI 10.1002/vnl

JOURNAL OF VINYL & ADDITIVE TECHNOLOGY 2009 5

FIG. 9. Mechanism for polymerization of PLA from corn starch by anionic groups.

convert the sugar into lactic acid through fermentation. Lactic acid molecules link to form rings called lactide monomer. The lactide rings open and link together to form a long chain of polylactide polymer. The plastic is then formed into pellets and can be used in a wide range of products, including packaging and bers [8, 18]. This ring-opening polymerization can be promoted by anionic as well as cationic groups. The mechanisms for both are given in Fig. 9 and Fig. 10, respectively [19].

comfort-stretch and recovery, dyeability, and easy care are desired [20].

Production of Furfural from Hemicellulose The production of furfural and its derivatives is a wellestablished technology. The major raw materials are hemicelluloses from annual crops. Although any agricultural product containing high-pentose-content hemicelluloses could be used in theory, only a few (corn cobs, oat hulls, rice hulls, and bagasse) are available in tonnage quantities within economical hauling distance from furfural production plants. Furfural production involves loading raw materials into a digester containing a strong inorganic acid. The hemicelluloses are hydrolyzed into pentoses, and these pentoses are then cyclodehydrated into furfural. High-pressure steam provides the necessary temperature, and the furfural is steam-distilled. Various subsequent chemical modications can produce furan derivatives, such as adipic acid, hexamethylenediamine, 1,4-butanediol, and adiponitrile [21].

Conversion of Starch (Corn) to Propanediol Through metabolic engineering of biochemical pathways, researchers engineered a microorganism that can use sugars from corn and corn biomass in a fermentationbased process to manufacture 1,3-propanediol. The method incorporates a microorganism containing a dehydratase enzyme which is contacted with a carbon substrate, and 1,3-propanediol is isolated from the growth media. The single organism may be a wild type or a genetically altered organism harboring a gene encoding a dehydratase enzyme. The glucose obtained by the breaking down of corn starch is fed down pipes in a three-story vat (a large container) containing genetically engineered organisms, water, and some vitamins and minerals. The genetically modied organism eats the glucose and excretes the 3-carbon molecule 1,3-propanediol. 1,3-Propanediol, together with terephthalic acid, is used to produce polytrimethylene terephthalate (PTT). The PTT is a polymer with remarkable stretch-recovery properties and is used in apparel, upholstery, specialty resins, and other applications where properties such as softness,

Production of Ethylene from Biomass Ethylene is perhaps the most important petrochemical because of the value of its numerous derivatives, such as polyethylene (PE), ethylene dichloride, vinyl chloride, ethylene oxide, styrene, ethanol, vinyl acetate, and acetaldehyde. Today, biobased ethylene production based on ethanol derived from corn stover still is not cost-competitive with petroethylene sources. In the near term, ethylene based on lignocellulose fermentation could move into the

FIG. 10. Mechanism for polymerization of PLA from corn starch by cationic groups.

6 JOURNAL OF VINYL & ADDITIVE TECHNOLOGY 2009

DOI 10.1002/vnl

margin of competition against petrochemical sources. Corn starch fermentation yields ethanol, which then can be dehydrated for the production of ethylene [22]. Production of Aromatics from Lignin More research needs to be done to develop new product potential for the low-molecular-weight (LMW) lignin. The LMW lignin is soluble in common organic solvents such as ethanol and acetone, thus making structural transformation easier. Key intermediates made from LMW lignins will be low-priced, because much of the cost of generating them will be covered by the value of the pure cellulose and hemicellulose produced during the steam explosion process. Phenol, a source of antioxidant additives for polymers, and benzene can be made from LMW lignin in direct substitution for petrochemical processes [22]. RECENT DEVELOPMENTS Some recent developments in the production of polymers from renewable sources are given below. Biodegradable Laminate Films from Naturally Occurring Carbohydrate Polymers Coextruded laminate lms have become increasingly important for many applications, especially in the food industry. They are mainly applied in the packaging of products such as fresh pasta, meats, and cut vegetables to extend the shelf life of the goods. Commercial multilayer lms currently comprise a number of layers of different polymers. The advantage of multilayer lms is that different lms forming different layers offer different advantages, and as a whole, the properties of these lms are far superior to those of monolayer lms. In most applications, the outer layers consist of cheap, water-barrier polymers with good mechanical properties (PE, polystyrene), while the inner layers consist of more expensive materials that offer good gas-barrier properties (polyvinylidene chloride, polyethylene terephthalate). However, these are not biodegradable. Conventional plastics, manufactured from fossil fuels, not only consume nonrenewable and nite resources, but also impact heavily on waste disposal. Laminates based on modied starch and PLA offer the following advantages:
1. Good water- and gas-barrier properties. 2. Good processability. 3. Good biodegradability at the end of the product life.

native starch to form esters (at substitution levels of degree of substitution 1.53.0) gives a material that can be extruded, processed, and shaped in the same way as traditional plastic products. Plasticizers such as glycerol triacetate and diethyl succinate which are completely miscible with starch esters can be used to improve processability. The water resistance of such starch esters is greatly improved over that of the unmodied starch. The method comprises gelatinization of the candidate starch in hot water, followed by the introduction of an organic solvent and preferential removal of the water. The resulting anhydrous suspension is heated with lithium chloride, which renders the gelatinized starch soluble in the organic solvent. In this homogeneous form, conventional esterication reactions are performed, and the modied starches are isolated after appropriate workup [23]. Biodegradable Polyurethane Networks Based on Castor Oil Among vegetable oils, castor oil represents a promising raw material because of its low cost, low toxicity, and availability as a renewable agricultural resource. Its major constituent, ricinoleic acid (12-hydroxy-cis-9-octadecenoic acid), is an hydroxyl-containing fatty acid. So, castor oil can be used directly as a raw material for the preparation of polyurethanes without any future modication. Polyurethane networks based on castor oil as a renewable resource polyol and poly(ethylene glycol) (PEG), with tunable biodegradation rates, are potential candidates for biomedical implants and tissue engineering and were synthesized through the reaction of epoxy-terminated polyurethane prepolymers (EPUs) with a 1,6-hexamethylenediamine (HMDA) curing agent. The EPUs were prepared by the reaction of glycidol with isocyanate-terminated polyurethane prepolymers made from castor oil or PEG and 1,6-hexamethylene diisocyanate, and they are called EPU1 and EPU2, respectively. The synthetic route for EPU1 production is shown in Fig. 11. Degradation rate and mechanical properties of the nal products could be controlled by the ratio of the PEG- to castor oil-based EPUs in these products. Increasing the PEG-based EPU content causes an increase in hydrolytic degradation rate and improved mechanical properties. The synthetic route for polyurethane from PEG follows the same route as above, except that the triglyceride is replaced by ethylene glycol. Crosslinked polyurethane networks (CPUs) were prepared via the reaction of HMDA with EPUs. Upon curing, the N groups of HMDA react with the epoxy H groups of the EPUs, and a crosslinked network is produced. The amounts of HMDA and EPUs are adjusted in a way such that a 1:1 molar ratio of epoxy and NH groups is established. The curing reactions could be performed at room temperature. However, to increase the rate, they were carried out at elevated temperatures. During the testing of these polyurethane networks, various blends of EPU1 and EPU2 were prepared by varying the

Chemical modication of starch is needed to overcome issues of thermoplasticity and water sensitivity. Properties of modied starch can be varied, depending on the nature of the modication and the degree to which modication is made. Modication of the abundant hydroxyl groups of

DOI 10.1002/vnl

JOURNAL OF VINYL & ADDITIVE TECHNOLOGY 2009 7

and improve synthesis in the host bacterium by mutating (modifying) the existing biosynthetic pathway. Another is to move the genes for PHB synthesis into other bacteria, plants, or yeasts for increased production. At Michigan State University, Sommerville and colleagues pioneered this approach by using the common weed Arabidopsis as a bioreactor. The researchers manipulated genes for PHB synthesis in Arabidopsis and showed that the plant produced PHB at a low level. Moving the genes to the target expression in a different part of the plant cell (i.e., from the cytoplasm to the chloroplast) dramatically increased PHB production. Increased understanding of the basic science underlying the plant and bacterial metabolic and biosynthetic pathways has made possible another exciting development; i.e., new polymer structures can now be engineered by manipulating the PHB metabolic pathway in various plants and microbes. For example, less brittle plastics may result if low amounts of poly(hydroxyvalerate) are coproduced in bacterial or plant cells that manufacture PHB. These polymers have properties that make them a suitable substitute for petrochemical-derived thermoplastics [2, 10, 25, 26].
FIG. 11. Synthetic route for preparation of polyurethane from castor oil.

amount of EPU2. It was observed that the mechanical properties of the blends were improved by increasing the EPU2 content. Surface and bulk hydrophilicities and consequently the hydrolytic degradation rate increased with the content of EPU2 [24].

Production of PHB and its Variants from Microorganisms PHB and its variants, generally known as poly(hydroxyalkanoates), are a family of natural polymers produced by many bacterial species for carbon and energy storage. They are extremely versatile and can be used in a broad range of applications. The PHB is a biodegradable plastic material that is naturally and efciently degraded into carbon dioxide and water by many common soil bacteria. Under appropriate growth conditions, bacteria such as Alcaligenes eutrophus may contain 90% of their total cell mass as PHB. The PHB is derived from acetyl-CoA, a component of primary metabolism, by a process involving three enzymes. The enzyme b-ketothiolase condenses two molecules of acetyl-CoA to yield acetoacetyl-CoA. This compound is reduced to b-hydroxybutyryl-CoA by acetoacetyl-CoA reductase and then condensed to a nascent polymer chain by PHB polymerase. The ICI Corporation has commercialized biodegradable PHB plastics for shampoo bottles and other higher-cost disposables. Commercial production by fermentation is currently under way for poly-3-hydroxybutyrate-3-hydroxyvalerate (PHB-V), a PHB that has characteristics similar to those of PP or PE. A possible approach to improve production is to increase

PHB Versus PP. Production of 1000 kg of PP releases 3530 kg of CO2 , which is over 80% more than PHB proeq duction. Ozone layer depletion is over 50 times lower in PHB production, thus representing the greatest reduction in all categories. Terrestrial toxicity is almost 10 times higher in PP production. The decrease in these impacts in PHB production is mainly due to the greatly reduced requirement for crude oil, which in PP production is needed for both energy and propylene monomer unit production [25, 26]. PHB Versus PE. In PE production, the impacts of abiotic depletion, freshwater toxicity, terrestrial toxicity, human toxicity, photochemical oxidation (HDPE), and ozone layer depletion are 1.7, 1.61.9, 34, 3, 22, and 4 10 times greater, respectively, than those for PHB production, depending on the type of PE produced [25, 26]. Composites Based on Copolymer of Epoxidized Allyl Soyate and Commercial Epoxy Resins Unsaturated triglyceride oils, such as soybean, crambe, linseed, and castor oil, constitute one major class of renewable resources. The main components of these triglyceride oils are saturated and unsaturated fatty acids. The common method by which soybean seed can be converted into soy oil or soybean meal is shown in Fig. 12. They can be polymerized to form elastomeric networks and promise alternative material resources to petrochemical-derived resin [27]. The major unsaturated fatty acids in soybean oil triglycerides are linolenic acid (7%); linoleic acid (51%); and oleic acid (23%). The largest category of industrial soybean oil use is in plastics and resins.

8 JOURNAL OF VINYL & ADDITIVE TECHNOLOGY 2009

DOI 10.1002/vnl

Synthesis of 4-[(4-Hydroxy-2-pentadecenylphenyl)diazenyl] phenol (HPPDP) from Cashew Nut Shell Liquid (CNSL) The CNSL contains four major components: 3-pentadecenylphenol (cardanol), 5-pentadecenylresorcinol (cardol), 6-pentadecenylsalicylic acid (anacardic acid), and 2methyl-5-pentadecenylresorcinol (2-methylcardol). Recently, it has been reported that a series of polyurethanes has been synthesized from HPPDP and a polyether by modifying cardanol with epichlorohydrin and polymerizing through ring-opening polymerization. The monomeric HPPDP can be synthesized from cardanol, which acts as a renewable resource in the synthesis of the polymer. The monomer synthesis can be carried out as shown in Fig. 13. The red dye obtained (HPPDP) can be puried by column chromatography on silica gel using chloroform as eluent. Solvents are removed, and the product is recrystallized from a methanolwater mixture. Yields as high as 80% can be obtained for HPPDP with a reddish-green color and a melting point of 1301388C [9].

FIG. 12. Techniques to convert soybean seed into soy oil and deated rubber.

Epoxidized soybean oil (ESO) is currently being used mainly as a plasticizer or stabilizer to modify the properties of plastic resins such as PVC. The ESO can be used as a reactive modier and diluent of the epoxy resin system. Several researchers have investigated the curing and conversion of the soybean oil into exible, semiexible, and rigid crosslinked polyester by using various approaches. These studies show the potential for the synthesis of new polymers derived from renewable soybean oil. So far, the wide structural applications of ESO are limited because of its low crosslink density and mechanical performance. Soy-based epoxy resin, namely, epoxidized allyl soyate (EAS), has been developed from unsaturated soybean oil. It is expected that modied soy-based epoxy resins will possess higher reactivity and therefore provide denser crosslinking that yields materials which are stronger than the materials obtained from commercially available unmodied ESO. The EAS can be synthesized through a two-step process from food grade soybean oil. First, the large molecular triglycerides are transesteried to yield fatty acid methyl ester and allyl ester by using methyl alcohol and allyl alcohol, respectively. In the second step, the fatty acid esters are epoxidized to yield soyate epoxy resins. The epoxidation of the carbon-carbon double bond of the unsaturated vegetable oil is known to proceed without molecular rearrangement. The copolymerization of EAS with a commercial epoxy resin yields a viable lowcost, high-performance thermoset product and is ideal for the pultrusion process. Mechanical tests show that pultruded glass-ber-reinforced composites with soy-based co-resin systems possess comparable or improved structural performance characteristics such as exural strength, modulus, and impact resistance. The lubricity of soybean oil signicantly reduces the pull force in the pultrusion process. The epoxidized soy-based resin system holds great potential for environmentally friendly and low-cost raw materials in the fabrication of epoxy composites for structural applications [5].

Synthesis of Thermosetting Polymers from Fish Oil Fish oil can be copolymerized with maleic anhydride to produce maleated sh oils, which have been used industrially as drying oils, heat-resistant varnishes, and plasticizers in poly(vinyl chloride). New polymers have been prepared from both native and conjugated sh oil (FO and CFO). These new polymers are prepared through cationic copolymerization, and boron triuoride diethyl etherate is used as the initiator. Divinylbenzene, norbornadiene, and dicyclopentadiene are employed as comonomers. The advantages of these plastics include low cost and the renewable nature of the sh oil. The CFO can be produced from FO through the use of Wilkinsons catalyst. Generally, sh oil has a triglyceride structure with saturated, monounsaturated, and multiunsaturated fatty acid side chains. The high unsaturation in the FO structure is expected to provide reactive sites for the cationic polymerization of the FO molecules. However, the simple

FIG. 13. Synthetic route for preparation of 4-[(4-hydroxy-2-pentadecenylphenyl)diazenyl]phenol.

DOI 10.1002/vnl

JOURNAL OF VINYL & ADDITIVE TECHNOLOGY 2009 9

homopolymerization of FO by boron triuoride diethyl etherate does not lead to solid polymeric materials. The reactivity of the sh oil is increased by conjugating the carbon-carbon double bonds. Therefore, CFO is also used. These polymers are found to be typical thermosetting materials. They have glass transition temperatures ranging from 50 to 1108C, to an extent that depends on the original composition of the reactants. The modulus at ambient temperature of the CFO polymers can be as high as 1.1 3 109 Pa, which is comparable to those of conventional polymers, such as polyolens [3].

FIG. 15. Epoxidized soybean oil.

Production of Polycaprolactone (PCL) from Saccharides Polycarprolactone (Fig. 14) is used mainly in thermoplastic polyurethanes, resins for surface coatings, adhesives, and synthetic leather and fabrics. It also serves to make stiffeners for shoes and orthopedic splints, as well as fully biodegradable compostable bags, sutures, and bers. Although the polymer is now being prepared from fossil resources, it can also be made from renewable resources via the chemical treatment of saccharides. In the rst step, the saccharides are converted into ethanol and acetic acid by fermentation. In the second step, ethanol is converted into cyclohexanone by the use of chromic acid. Cyclohexanone thus obtained is then converted into PCL. This is not a high-yield or a high-conversion reaction. Also, even the production of ethanol by the synthetic route has a better payback time than the ethanol production from fermentation. Therefore, at least at present, it is more desirable to produce PCL via the synthetic route. Naturally, such a product is expensive and is therefore mixed with large amounts of other natural materials in order to obtain a good biodegradable material at a low price. Fibers of PCL can be used for incorporation into a PCL matrix for craniofacial bone repair. The PCL ber can be xed tightly by melting, and the strength of this thread is found to be higher than that of a stainless-steel wire with the same cross-sectional area [8]. FUTURE PROSPECTS Genetic mapping will be one of the thrust areas of research in the future. Genetic mapping techniques allow researchers to sequence genes more quickly and to determine the relationship between structure and function [28]. These genetic tools are being used to create new and improved microorganisms which convert biomass compo-

nents into end products or intermediates that can then be thermochemically upgraded. The tools are also being used to improve biomass feedstocks by:
1. Increasing the content of desired components. 2. Decreasing the content of components such as lignin. 3. Adding the capability to produce a new component (e.g., a new fatty acid in oilseeds).

Another area where research will be needed is in the production of nonbiodegradable biopolymers, as we do not always want the biopolymer to degrade under atmospheric conditions. Some success has also been achieved in this area. One example is as follows. Recently, it was shown that some microorganisms capable of poly(hydroxyalkanoate) synthesis are also able to synthesize polymers of mercaptoalkanoic acids, which are generally referred to as polythioesters (PTEs). Although the building blocks of PHAs are covalently linked by oxoester bonds, in PTEs they are linked by thioester bonds. Both polymers are synthesized by PHA synthase, which is the key enzyme of PHA biosynthesis. It has also been observed that incorporation of biopolymers into the conventional polymers can bring about an enhancement in the properties [29, 30]. Example In epoxidized soybean oil (ESO, Fig. 15)-(reinforced diglycidyl ether of Bisphenol-A), ESO brings about an increase in the impact strength of the epoxy resin [31]. CONCLUSION Polymers are preferred over traditional materials (metals, glass, wood, clay, etc.) because of their superior properties (light weight, corrosion and shatter resistance, longevity) and lower cost. As a result, polymers have become an integral part of our lives. The development of polymers from renewable sources is best viewed in the context of diminishing crude-oil reserves. In future years, it will be largely driven to derive more carbon for chemical processes from renewable resources and to preserve the ecosystem. Also, the properties of polymers from biological sources are not at all inferior when compared with those of synthetic polymers. Typical properties of PLA

FIG. 14. Polycaprolactone.

10 JOURNAL OF VINYL & ADDITIVE TECHNOLOGY 2009

DOI 10.1002/vnl

and other synthetic polymers are shown in Table 1 [32 35]. There are various technological approaches to the production of polymers from renewable sources. The fermentation industry, supported by the developments of genetic engineering, can yield microbes that more efciently convert inexpensive raw materials into desired polymers. Such powerful tools will make the industrial production of a wide range of chemicals, including monomers, polymers, and polymer additives, very feasible. Now the technical background of the food industry permits us to develop environmentally friendly products based on starch and composite materials. Economics and environmental considerations will be the driving forces as the commercialization of polymers from renewable sources continues. REFERENCES
1. L. Zhu and R.P. Wool, J. Mol. Catal. B: Enzym., 45, 39 (2007). 2. N. Jacquel, C.-W. Lo, Y. Wei, H. Wu, and S.S. Wang, Biochem. Eng. J., 39, 15 (2008). 3. F. Li, D.W. Marks, R.C. Larock, and J.U. Otaigbe, Polymer, 41, 7925 (2000). 4. J.H. Clark, Green Chem., 8, 17 (2006). 5. R. Kumar, V. Choudhary, S. Mishra, I.K. Varma, and B. Mattiason, Ind. Crops Prod., 16, 155 (2002). 6. S. Warwel, F. Bruse, C. Demes, M. Kunz, and M.R. Gen. Klaas, Chemosphere, 43, 39 (2001). 7. G. Scott, Polym. Degrad. Stab., 68, 1 (2000). 8. M. Flieger, M. Kantorova, A. Prell, T. Rezanka, and J. Votruba, Folia Microbiol. (Prague), 48, 27 (2003). 9. H.P. Bhunia, A. Basak, T.K. Chakia, and G.B. Nando, Eur. Polym. J., 36, 1157 (2000). 10. M. Paster, J.L. Pellegrino, and T.M. Carole, Industrial Bioproducts: Today and Tommorow, US Department of Energy, Washington, DC, 189 (2003). 11. National Research Council (U.S.), Committee on Biobased Industrial Products, Biobased Industrial Products: Research and Commercialization Priorities, National Academy Press, Washington, DC, 2654 (2000). 12. National Renewable Energy Laboratory Research Review, Unraveling the Structure of Plant Life, National Renewable Energy Laboratory, Colorado, 1015 (2004).

13. M. Palm and G. Zacchi, Sep. Purif. Technol., 36, 191 (2004). 14. D.N.S. Hon and N. Shiraishi, Wood and Cellulosic Chemistry, 2nd ed., Marcel Dekker, New York, 109174 (2000). 15. W.A. Farone and J.E. Cuzens, U.S. Patent 467,222 (1997). 16. B. Shivani, A.K. Mishra, N.K. Mishra, and M.A. Khan, J. Mater. Process. Technol., 183, 273 (2007). 17. NatureWorks, http://www.natureworksllc.com/aboutnatureworks- llc/from-plants-to-plastic.aspx (last accessed June 2008). 18. Z.Y. Zhang, B. Jin, and J.M. Kelly, Biochem. Eng. J., 35, 251 (2007). 19. A.P. Gupta and V. Kumar, Eur. Polym. J., 43, 4053 (2007). 20. L.A. Laffend, V. Nagarajan, and C.E. Nakamura, U.S. Patent 09/575,638 (2000). 21. S. Hussain, A.H. Fawcett, and P. Taylor, Prog. Org. Coat., 45, 435 (2002). 22. A. Aden, J. Bozell, J. Holladay, J. White, A. Manheim, Top Value Added Chemicals From Biomass, US Department of Energy, 167 (2004). 23. J.M. Fang, P.A. Fowler, C. Escrig, R. Gonzalez, J.A. Costa, and L. Chamudis, Carbohydr. Polym., 60, 39 (2005). 24. H. Yeganeh and P. Hojati-Talemi, Polym. Degrad. Stab., 92, 480 (2007). 25. M. Akiyama, T. Tsuge, and Y. Doi, Polym. Degrad. Stab., 80, 183 (2003). 26. K.G. Harding, J.S. Dennis, H. von Blottnitz, and S.T.L. Harrison, J. Biotechnol., 130, 57 (2007). 27. K. Chandrashekhara, S. Sundararaman, V. Flanigan, and S. Kapila, Mater. Sci. Eng., A, 412, 2 (2005). 28. A. Ohta, Y. Yamagishi, and H. Suga, Curr. Opin. Chem. Biol., 12, 159 (2007). 29. D.Y. Kim, T. Lutke-Eversloh, K. Elbanna, N. Thakor, and A. Steinbuchel, Biomacromolecules, 6, 897 (2005). 30. A. Steinbuchel, Curr. Opin. Biotechnol., 16, 607 (2005). 31. F. Jin and S. Park, Mater. Sci. Eng., A, 478, 402 (2008). 32. IDES, http://www.ides.com/generics/Polyester/Polyester_ typical_properties.htm (last accessed September 2008). 33. IDES, http://www.ides.com/generics/PLA/PLA_typical_ properties.htm (last accessed September 2008). 34. IDES, http://www.ides.com/generics/PP/PP_typical_properties.htm (last accessed September 2008). 35. IDES, http://www.ides.com/generics/PC/PC_typical_properties.htm (last accessed September 2008).

DOI 10.1002/vnl

JOURNAL OF VINYL & ADDITIVE TECHNOLOGY 2009 11

Вам также может понравиться