Sunteți pe pagina 1din 24

Chapter 1

INTRODUCTION

1.1 Background of the Study

Plastics are ubiquitous in contemporary society. They are found in households as

kitchenware, toys, and resilient flooring to name a few, and are extensively employed in many

such key industries as automobile, aerospace, electrical, and medical industries (Richardson,

2005).

Useful as they are, plastics are disposed constantly by people, and eventually

accumulate to large piles along with other wastes. Various efforts have been addressed in

response to the problem: the creation of sanitary landfills in strategically located places,

adoption of selective collecting and recycling programs, and active government and civic

participation (Vilpoux and Averous, 2004); still, people resort to quicker, short-term solutions

such as incineration and waste burning at household levels, aggravating the effects of global

warming in the planet.

One effective procedure to address this global problem is the creation of

biodegradable plastics, which currently exist in the market. Biodegradable plastics refer to

degradable polymers that are consumed by microorganisms through polymer chain

breakdown and subsequent utilization (Garth and Kowal, 2001).

Biodegradable plastics have used starch as fillers in the 1970’s in response to the rise

of petroleum prices, the latter being the prime component of plastics manufacture (Curvelo, et

al., 2001; Garth and Kowal, 2001). Starch-based polymers are made from starch, which is a

plant storage polysaccharide made up of repeating glucose groups linked by glucosidic

linkages in the C1 and C4 positions (Whistler and Hilbert, 1984). Made up of two polymers,

the linear-chain amylose and the branched-chain amylopectin, the alpha linkage of amylose

allows it to be flexible and digestible, hence giving it the ability to be converted to starch-

based biodegradable polymers.


2

Among its advantages, starch is cheap, abundant, and renewable; moreover, it is

found in several forms due to the origin of its raw material (Lawter and Fischer, 2000).

Although easily obtained from such agricultural commodities as corn, potatoes, and tapioca,

starch production and extraction has been done in expense of food production and

reforestation (Garth and Kowal, 2001).

A superlative substitute could be sago starch from the sago palm, a crop widely

cultivated in peat soil land in tropical areas (Shibata et al., 2006). The palm exhibits large

starch productivity, even amounting to twenty-five tons/hectare-year, consequently serving as

a notable starch source, and alleviating extensively the problem of accruing atmospheric

greenhouse gases. The palm, in addition, is abundant here in the Philippines, specifically in

the island of Mindanao and some parts of the Visayan Islands.

Starch-based plastics have starch contents from 10% to 90% and can be any one of

following types: surface-modified starch additives; gelatinized starch additives; and

thermoplastic starch materials (Garth and Kowal, 2001); thermoplastic starch blends with

synthetic aliphantic polysters such as poly-ε -caprolactone (PCL), poly-lactic acid (PLA),

and polybutylene succinate adipate (PBSA); several others are also existent nowadays

(Nolan-ITU Pty, 2002). Nonetheless, of these categories, the thermoplastic starch materials

gained attention since the high amount of starch, 70 to 100 percent, engenders a material with

increased biodegradability (Garth and Kowal, 2001).

Thermoplastic starch, or TPS, can be made by cast-molding in dry drums, or a

continuous supply of thermomechanical energy. More often than not, this combination of

thermal and mechanical input can obtained by extrusion. In this process, starch, along with

plasticizers such as water, is fragmented, destructured, plasticized, melted, and partially

depolymerized through high temperature and shearing in an extruder, giving a homogenous

molten product called the extrudate (Averous, 2004).


3

1.2 Objectives of the Study

This study basically aimed to utilize sago starch as the main substrate for

thermoplastic starch material production. Moreover, it specifically aimed to:

1. produce thermoplastic sago starch from extrusion, which can be

useful for some practical application to be identified later, if

successful;

2. perform preliminary testing on tensile strength, water

disintegration and biodegradability of the resulting material.

1.3 Hypothesis

H0: There is no significant difference on the mean tensile strength of the TPS exposed

to different treatments.

1.4 Conceptual Framework

INDEPENDENT VARIABLE DEPENDENT VARIABLE

Tensile Strength

Thermoplastic starch Biodegradability

Water Disintegration

1.5 Scope and Limitation

For this study, the molten product was evaluated by its tensile strength and

degradability only. The extrusion processing part of the study took place in University of the

Philippines in Mindanao Department of Food Science and Chemistry Laboratory.


4

1.6 Significance of the Study

This study serves to find alternative ways of controlling the problem on accumulation

of plastics in the environment. The production of thermoplastic starch from sago starch can be

of great help in reducing plastics in the environment since they are composed of degradable

polymers, and will eventually degrade into substances that are less harmful to the

environment. Furthermore, this study will help in minimizing the factors that contribute to

global warming, since it is degradable unlike any other common plastics that require burning

for decomposition.

1.7 Definition of Terms

Glycerol -A syrupy, sweet, colorless or yellowish liquid, C3H8O3, obtained from fats

and oils as a byproduct of saponification and used as a solvent, an antifreeze,

a plasticizer, and a sweetener and in the manufacture of dynamite, cosmetics,

liquid soaps, inks, and lubricants.

Plasticizer -Any of various substances added to plastics or other materials to make or

keep them soft or pliable.

Starch -A naturally abundant nutrient carbohydrate, (C6H10O5)n, found chiefly in the seeds,

fruits, tubers, roots, and stem pith of plants, notably in corn, potatoes, wheat,

and rice, and varying widely in appearance according to source but

commonly prepared as a white amorphous tasteless powder.


5

Tensile strength - The resistance of a material to a force tending to tear it apart,

measured as the maximum tension the material can withstand without

tearing.

Biodegradable - Capable of being decomposed by biological agents, especially

bacteria: a biodegradable detergent.

Disintegrate -To become reduced to components, fragments, or particles.

1.8 Literature Review

1.8.1 Biodegradable Plastics

Materials that are consumed by microorganisms such as bacteria, fungi, and algae are

commonly referred to as biodegradable polymers. These microorganisms break down the

polymer chain and consume the material through several methods. Biodegradable plastics can

either be hydrolyzable (able to change chemically in water) or water-soluble (able to dissolve

in water). Some common biodegradable plastics are polyesters, polyhydroxybutyrates, and

vinyl polymers. An in-depth description of these biodegradable plastics is listed in Table 1.

Table 1. Types of biodegradable plastics (Garth and Kowal, 2001).

Type Name Abbreviat Description Uses


ion
Polyester Polyglycolic acid PGA Hydrolyzable Specialized applications;
polyhydroxy acid controlled drug releases;
implantable composites

Polylactic acid PLA Hydrolyxable Packaging and paper


polyhydroxy acid; coatings; other possible
polymers derived markets include
from fermenting sustained release
crops and dairy systems for pesticides
products; and fertilizers, mulch
compostable films, and compost
bags
Polycaprolactone PCL Hydrolyzable; low Long term items; mulch
softening and and other agricultural
melting points; films; fibers containing
compostable; long herbicides to control
time to degrade aquatic weeds; seedling
containers; slow release
6

systems for drugs

Polyhydroxybutyrate PHB Hydrolyzable; *


produced as storage
material
microorganisms;
possibly degrades in
aerobic and
anaerobic
conditions; stiff;
brittle; poor solvent
Resistance
Polyhydroxyvalerate PHBV Hydrolyzable Films and paper
copolymer; coatings; other possible
processed similar to markets include
PHB; contains a biomedical
substance to increase applications,
degradability, therapeutic delivery of
melting point, and worn medicine for
toughness; cattle, and sustained
comppostable; low release systems for
volume and costly pharnmaceutical drugs
production and insecticides
Vinyl Polyvinyl alcohol PVOH Water soluble; Packaging and bagging
dissolves during applications which
composting dissolves in water to
release products such as
laundry detergent,
pesticides, and hospital
washables
Polyvinyl acetate PVAC Water soluble; *
predecessor to
PVOH; has shown
no significant
property loss during
composting tests
Polyenylketone PEK Water soluble; *
derived from PVOH;
possibly degrades in
aerobic and
anaerobic conditions

The ‘biodegradability’ of plastics is dependent on the chemical structure of

the material and on the constitution of the final product, not just on the raw materials used

for its production. Therefore, biodegradable plastics can be based on natural or synthetic

resins. Natural biodegradable plastics are based primarily on renewable resources (such as
7

starch) and can be either naturally produced or synthesized from renewable resources. Non-

renewable synthetic biodegradable plastics are petroleum-based. As any marketable plastic

product must meet the performance requirements of its intended function, many natural

biodegradable plastics are blended with synthetic polymers to produce plastics which meet

these functional requirements (Nolan-ITU Pty, 2002).

1.8.2 Starch-based Polymers

Because the oil embargo in the early 1970s caused the price of plastics to almost

double, researchers began to look for less expensive, non-plastic filler. The result of their

search was starch-based polymers. Since that time, starch-based polymers have remained

the most commonly used and lowest-costing ingredient of all biodegradable polymers

(Garth and Kowal, 2001).

Starch-based biodegradable plastics may have starch contents ranging from 10% to

greater than 90%. Starch based polymers can be based on crops such as corn (maize), wheat

or potatoes. Starch content needs to exceed 60% before significant material breakdown

occurs. As the starch content is increased, the polymer composites become more

biodegradable and leave less recalcitrant residues. Often, starch-based polymers are blended

with high-performance polymers (e.g. aliphatic polyesters and polyvinyl alcohols) to

achieve the necessary performance properties for different applications (Averous, 2004;

Nolan-ITU Pty, 2002).

Biodegradation of starch based polymers is a result of enzymatic attack at the

glucosidic linkages between the sugar groups leading to a reduction in chain length and the

splitting off of sugar units (monosaccharides, disaccharides and oligosaccharides) which are

readily utilized in biochemical pathways (Averous, 2004).

At lower starch contents (less than 60%) the starch particles act as weak links in

plastic matrix and are sites for biological attack. This allows the polymer matrix to

disintegrate into small fragments, but not for the entire polymer structure to actually

biodegrade (Garth and Kowal, 2001; Averous, 2004).


8

There are several categories of biodegradable starch-based polymers including:

thermoplastic starch products; starch synthetic aliphatic polyester blends; starch PBS/PBSA

polyester blends; and starch PVOH blends (Averous, 2004).

1.8.3 Thermoplastic Starch Materials

Thermoplastic starch, or TPS, biodegradable plastics have starch contents greater

than 70% and are based on gelatinized vegetable starch, and with the use of specific

plasticizing solvents, can produce thermoplastic materials with good performance

properties and inherent biodegradability. Starch is typically plasticized, destructed, and/or

blended with other materials to form useful mechanical properties. Importantly, such TPS

compounds can be processed on existing plastics fabrication equipment. High starch

content plastics are highly hydrophilic and readily disintegrate on contact with water. This

can be overcome through blending, as the starch has free hydroxyl groups which readily

undergo a number of reactions such as acetylation, esterification and etherification

(Averous, 2002).

The Cooperative Research Center (CRC) for the International Food Manufacture

and Packaging Science Australia has developed its own version of TPS biodegradable

plastics. These natural vegetable starch polymers have amylose content greater than 70%.

Trials have been successfully performed using maize starch polymers as mulch film, and

the material was found to perform as well as polyethylene film, with the added advantage

that after harvest, the film can be simply ploughed into soil. These natural starch polymers

are now being commercialized through a new company called Plantic Technologies Ltd.

Based in Melbourne (Averous, 2004).

The applications of thermoplastic starch polymers are generally film, such as

shopping bags, bread bags, bait bags, over wrap, ‘flushable’ sanitary product backing

material, and mulch film. Foam loose fill packaging and injected molded products such as

takeaway containers are also potential applications. Foamed polystyrene can be substituted

by starch foams that are readily biodegradable in some loose-fill packaging and foam tray
9

applications. Foamed starch loose-fills are rather easy products to produce and this area has

become an early market for biodegradable plastics. During its, preparation, raw starch is

premixed with 25 to 50 weight percent water and fed into an extruder capable of imparting

intensive shear and operating at high temperature (higher than the boiling point of water,

i.e., 150-180°C). Under these conditions of shear and temperature, starch breaks down,

loses its crystallinity, and gets plasticized with water, resulting in a homogenous amorphous

mass. When this gelatinized starch/water mixture exits the extruder, the water that is present

in the mass that is present in the mass at a temperature higher than its boiling point expands

into steam due to a sudden drop in pressure, and the foam is formed. Generally a plasticizer

(such as glycerol) and other polymer (such as polyvinyl alcohol) impart more reproducible

properties to starch foam. Along with the biodegradation of the polymers by sugar

molecules, certain TPS grades are also fully water solube (Averous, 2004).

The TPS is a relatively new concept and, today, it is one of the main research hints

for the manufacturing of biodegradable materials (Curvelo, et al., 2001). The starch is not a

real thermoplastic, but, in the presence of a plasticizer (water, glycerin, sorbitol,etc.), high

temperatures (90-180°C) shearing, it melts and fluidizes, enabling its use in injection,

extrusion and blowing equipment, such as those for synthetic plastics. So as to obtain a

thermoplastic starch, it is necessary that the starch maintain its semi-crystal granular

structure and that it behaves in a way similar to that of a melted thermoplastic, obtained

through a mono- or twin-screw extrusion with the use of mechanical and thermal energy

(Lourdin et al., 1999).

The water added to formula plays two roles: it is an agent that breaks the structure

of the native granule, breaking the bonds of hydrogen chains, and it is a plasticizer.

However, it is necessary to add an additional plasticizer besides water, such as polyol,

which will be only slightly influenced by the atmospheric conditions in the sorption-

desorption mechanism, and that will allow a melting phase at a temperature lower than that

of the starch degradation (Averous, 2002). During starch extrusion, the combining of

shearing, temperature and plasticization allows the obtaining of a melted thermoplastic


10

material (Averous et al., 2001). Afterwards this material can be transformed by means of

thermoforming or injection molding. The low resistance to water and variations in

mechanical properties under humid conditions affect the use of starch. Starch derivatives

present a high permeability to moisture and degrade rapidly for many types of applications.

Changes to remedy these problems make the final product more expensive, what limits any

solutions (Lawter and Fischer, 2000). According to Averous et al. (2001), besides being

susceptible to moisture, starch biopolymers have mechanical properties that change as time

goes by, apart from their low resistance to impacts. Moreover, in the case of thermoformed

products, the thinness of formed products limits the use of starch derivatives. According to

Levesque (2001), the use of starch and its derivatives as the only component of

biodegradable materials was object of a number of surveys. Techniques on starch

plasticization were developed to allow the manufacturing of objects (usually massive

objects), which, however, evolve rapidly due to the atmospheric humidity. It is necessary to

use a large quantity of plasticizers or destructurating compounds with hydrophilic

properties to plastify the starch. This material can be processed in the same way as synthetic

materials: extrusion, injection, etc. According to Lourdin et al. (1999), a plasticizer such as

glycerol and sorbitol is added in a ratio ranging from 20% to 40% to the starch weight. The

plasticizer content is directly related to the mechanical properties and glassy transition of

the material. The use of plasticizer also allows for a lower water activity, limiting the

growth of microorganisms (Weber, 2000).

The properties of plasticized starches depend a lot on moisture. As water has

plasticizing power, the material behavior changes according to the relative humidity of the

air (Averous, 2002), through a sorption-desorption mechanism. The higher the plasticizer

content (polyol), the higher is the moisture content. At the same time, the properties of the

material evolve as time goes by, even when moisture and temperature are controlled,

translating into a lower elongation breaking and higher rigidness. So as to improve the

impermeability of starch biodegradable products, Averous (2002) suggests the possibility of

incorporating biodegradable, renewable or synthetic polymers. These polymers can be


11

blended with the starch or applied to multilayered films, with an internal layer of starch and

two external layers of impermeable polymer. According to Averous et al. (2001) and Martin

et al. (2001), starch films can complecting with thermoplastic polyesters, such as

polycaprolactone (PCL), and then processed through blowing or calandering, as with

traditional plastics. As regards biodegrabadle films, the industrial process reaches a yield

equivalent to 80-90% of that obtained for low-density polyethylene (Bastioli, 2000).

The multilayered structure is more difficult to be produced than the blended one

(Averous, 2001). The main manufacturing technique is the co-extrusion (with a flat die) for

the production of a film that can be thermoformed for the manufacturing of, for example,

trays. Hot compression can also be applied to the films, but, in this case, there are fewer

adherences among the layers. Not all materials are compatible among themselves, and they

can come unstuck. In this case, compatibilizers are added (Averous, 2002). In the case of

multilayered products, delamination is measured by means of a peeling test at 90°C.

According to Martin et al. (2001), the adherence among the films diminishes through the

addition of more than 20% of glycerol.

According to Curvelo et al. (2001), the resistance to moisture can be enhanced by

adding synthetic polymers, cross-linked agents, such as Ca and Zn, salts, or lignin. TPS

properties vary according to the type of starch used, emphasizing the amylose content. The

almost linear increase of breaking forces and elongation according to the amylose content

can be explained through the larger agglomeration capacity of linear macromolecules while

ramified structures of amylopectin originate nodules. The amylose-rich corn starch (70% of

amylose) or a specific pea starch (35% of amylose is indicated to enhance the

manufacturing of these materials. However, even that the values of maximum forces for

breaking are acceptable (30 to 60 MPa), materials are fragile due to their low deformation

(approximately 6%). This property explains the reason why it is necessary to add other

components, such as plasticizers, able to enhance the plastic behavior (Lourdin et al., 1999).

.According to Lourdin et al. (1995 and 1999), the effect of the amylase content is

considered a favorable factor when there is no plasticizer. When plasticizers are added,
12

there is a reverse effect, since films with a larger content of amylase content can be

preferable. Besides the blend of starch with other polymers and the use of plasticizers,

many surveys have been conducted regarding the possibility of modifying the starch so that

the TPS can acquire the desirable properties with regards to both mechanical and moisture

resistance (Averous et al,. 2000).

Starch-based films found in the market are made mainly from starch complected

with 6thermoplastic polyesters, such as poly-epsilon caprolactone (PCL), to form

biodegradable products (Bastioli, 2000). When used for the manufacturing of bags for the

recycling of organic waster (composting), packaging, hygienic items and agriculture, the

properties of these films are similar to those of low density polyethylene. The more

common products of starch-based films are: purchase bags; consumer goods packaging;

food packaging; composting bags; hygiene and cosmetics; and funerary goods.

Purchase bags were introduced in the market in 1999 and started being used in

many supermarkets in Scandinavia and in the Mediterranean Coast. They were introduced

in places where collecting of organic wastes already existed and where they were accepted

as biodegradable compost bags. Bastioli (2000) stated that these bags have a behavior

similar to the bags made with LDPE with regards to the maximum weight and most of its

performances.

In consumer goods packaging, the main market is that of silk paper, but there are

markets for magazines’ wrapping and bubble films, mainly for electronic goods (Basioli,

2000).

It was further added that in food packaging, the main markets are those of bags for

fruit, vegetables and bakery products. The high cost of biodegradable products limits this

market, but starch films have an advantage against traditional plastics, that is to say, they

allow for a better breathing of the products.

Meanwhile, composting bags are used in the selective collecting of organic waste,

which will be treated to produce a compound. Million of Europeans already use these bags

(Bastioli, 2000)
13

In hygiene business lines, biodegradable polymers are found in diapers, swabs and

even in toothpicks. Moreover, biodegradable plastics can be used to wrap corpses, in

compliance with the rules on the use of biodegradable materials (Thouzeau, 2001).

According to Van Tuil et al. (2000), multi layered biodegradable films can be used

as food packaging with either controlled atmosphere or gas. Starch packaging boast the

characteristics of being an excellent barrier to O2 and CO2, facing problems only with water

and water vapor. The application of layers of biopolymers resistant to water in the ratio of

10% allowed Van Tuil et al. (2000) to obtain films resistant to water with a thickness from

30 to 100 μm. This material is currently commercialized by AVEBE under the name of

Paragon® and can be processed by means of blowing and used in multilayer’s with PLA or

PHBV (Weber, 2000).

1.8.4 Starch

Starch is the major carbohydrate reserve in plant tubers and seed endosperm where it is found

as granules (Buléon et al., 1998). It consists of two types of molecules: amylose (normally 20-

30%) and amylopectin (normally 70-80%). Both consist of polymers of α-D-glucose units in

the 4C1 conformation. In amylose these are linked -(14)-, with the ring oxygen atoms all on

the same side, whereas in amylopectin about one residue in every twenty or so is also linked -

(16)- forming branch-points. The relative proportions of amylose to amylopectin and -

(16)- branch-points both depend on the source of the starch, for example, amylomaizes

contain over 50% amylose whereas 'waxy' maize has almost none (Li and Yeh, 2001; Singh et

al., 2003). The representative partial structures of the two components of starch are shown in

Figure 1.
14

A
15

Fig. 1. Representative partial structures of (A), amylose and (B), amylopectin (Lifted from
Chaplin, 2008).

Amylose and amylopectin are inherently incompatible molecules; amylose having

lower molecular weight with a relatively extended shape whereas amylopectin has huge but

compact molecules. Although α-(14) links are capable of relatively free rotation, hydrogen

bonding between the O3' and O2 oxygen atoms of sequential residues tends to encourage a

helical conformation. These helical structures are relatively stiff and may impart adjacent

hydrophobic surfaces (Chaplin, 2008).

Amylose molecules consist of single mostly-unbranched chains with 500-20,000 α-

(14)-D-glucose units dependent on source (a very few α-16 branches and linked

phosphate groups may be found (Hoover, 2001), but these have little influence on the

molecule's behavior (Buléon et al., 1998). Hydrogen bonding between aligned chains causes

retrogradation and releases some of the bound water (syneresis). The aligned chains may then

form double stranded crystallites that are resistant to amylases. These possess extensive inter-
16

and intra-strand hydrogen bonding, resulting in a fairly hydrophobic structure of low

solubility. The amylose content of starches is thus the major cause of resistant starch

formation (Chaplin, 2008).

Amylopectin, on the other hand, is formed by systematized α-16 branching of the

amylose-type α-(14)-D-glucose structure. This branching is determined by branching

enzymes that leave each chain with up to 30 glucose residues. Each amylopectin molecule

contains a million or so residues, about 5% of which form the branch points. There are

usually slightly more 'outer' unbranched chains (called A-chains) than 'inner' branched chains

(called B-chains). There is only one chain (called the C-chain) containing the single reducing

group (Chaplin, 2008; Parker and Ring, 2001).

1.8.5 Sago Starch

Sago is a powdery starch made from the processed pith found inside the trunks of

the Metroxylon sagu palm, a pinnate-leaved palm occurring in the hot humid tropics of

Southeast Asia and Oceania. The palm is soboliferous; it produces tillers or suckers. Once

planted, a regular succession of suckers is produced from the lowest part of the trunk,

forming a cluster in various stages of development. In contrast to other sources of starch,

sago yields are exceptionally high. Under good conditions the range is from 15 to possibly

25 MT of air-dried starch/ha (6.7-11.1 MT/acre) of M. sagu at the end of an 8-year growth

cycle (Flach, 1997).

The sago palm is a promising source of carbohydrate. In areas where the sago palm

grows, sago pith is used for human consumption and livestock feeding. The palms are

harvested at the age of 7 to 15 years just before they flower. They only flower and fruit

once before they die. When harvested the stems are full of the stored starch which would

otherwise be used for flowering and fruiting. The trunks are cut into sections and into

halves and the starch is beaten or otherwise extracted from the heartwood, and in some
17

traditional methods it is collected when it settles out of water. One palm yields 150 to 300

kg of starch (Flach, 1997).

In Indonesia and Malaysia, sago from Metroxylon is used as a starch in making

noodles, white bread, and sago pearls (similar to tapioca). In India pearl sago (a form of

sago) is called Sabudana, and is used in a variety of dishes including khichdi, wafers and

puddings (Wikipedia, 2008). In Malaysia, sago starch is being tested in making

biodegradable plastics and in producing fuel alcohol and ethanol (PCARRD, 2001).

Sago starch contains 27% amylose, the linear polymer, and 73% amylopectin, the

branched polymer (Ito et al., 1979). Kawabata et al., (1984), however, give the amylose

content of sago starch as 21.7%. A grain size distribution of 16.0-25.4 mm was found. Jong

(1995), however, found a mean length of almost twice as much, about 40 mm. It is probable

that the grain size increases with age of the trunk, until initiation of the inflorescence.

The maximum viscosity amounts to 960 Brabender units, and gelatinization

temperature is 70°C. Viscosity decreases with deterioration of the starch quality, owing to

microbial activity (Cecil et al., 1986). In the modern starch industry, starches can be

modified to quite an extent. Provided there is a regular supply of cheap, clean and non-

corroded starch, sago will be clearly competitive to all other starches, and for some

purposes, it may even be preferred (Flach, 1997).

1.8.6 Extrusion

Extrusion plays a significant part on the plastic industry. The process is performed

through an extruder, as shown in Figure 2. The molding compound is continuously

conveyed down an electrically heated barrel by a rotating screw. As it proceeds down the

barrel, the compound melts and is forced through a die, which imparts the said compound’s

shape. The formed item is generally cooled and solidified in a water bath (Wikipedia, 2008;

Rosen, 1982).

1.8.7 Extrusion of Starch


18

Natural starch is a polysaccharide which is highly inhomogeneous due to the helical

formation of its glucose units, and must be modified and destructured to be converted and

used for other applications. To do this, starch must be gelatinized in the presence of water

and heat to disrupt the granular organization of the starch molecules. Such process results to

the reduction of the melting temperature and glass transition temperature of the starch,

definitive characteristics of thermoplastic starch (Averous, 2004).

Except for the purpose as filler to produce reinforced plastics, native starch must be

modified to find applications such as destructured starch. The destructuring agent is usually

water. Starch gelatinization can be obtained with the combination of water (high content)

and heat. Gelatinization, essentially, is the disruption of the granule organization. The

starch swells forming a viscous paste with destruction of most of intermacromolecule

hydrogen links. A reduction of both, the melting temperature (Tm or Tf) and the glass

transition (Tg) is then obtained. According to the level of destructuration and the water

content, different products and applications can be obtained; for instance, expanded

structures can be produced with rather high water content. Such closed cells structures

(foams) have been developed to obtain shock absorbable and isothermal packaging. Most

starch applications require water dispersion and partial or complete gelatinization. By

decreasing the moisture content (20% w/w), the melting temperature tends to be close to the

degradation temperature. For instance, for pure dry starch Tm is 220-240°C (Russell, 1988),

as compared to 220°C, which is the temperature at the beginning of starch decomposition

(Shogren, 1992). To overcome this last issue, we add a non-volatile (at the process

temperature) plasticizer to decrease Tm, such as glycerol or other polyols (sorbitol, PEG,

etc.). A mixture of different polyols can also be held. Other compounds such as those

containing nitrogen (urea, ammonium derived, amines) can be used.

The disruption of starch can be accomplished by casting or by applying

thermomechanical energy in a continuous process (Averous, 2004). The combination of

thermal and mechanical outputs can be obtained by extrusion. The process can be one or
19

two stages. In a one-stage process, the extruder, usually a twin-screw extruder is fed with

native starch. Along the barrel, water and liquid plasticizer are successively introduced. In a

two-stage process, the first stage is a dry blend preparation (Averous et al., 2000). Into a

turbo-mixer, under high speed, the plasticizer is added slowly in to the native starch until a

homogenous dispersion is obtained. Then, the mixture is placed in a vented oven allowing

the diffusion of the plasticizer into the granule. The plasticizer swells the starch. After

cooling, the right amount of water is added to the mixture using a turbo-mixer. This dry

blend is then introduced into an extruder. The different stages of the extrusion process are

given in Figure 4. The starch granules are fragmented. Under temperature and shearing,

starch is destructured, plasticized, melted but also partially depolymerized. After the

processing, we obtain a homogenous molten phase (Averous, 2004).

Fig. 2. Schematic diagram of starch extrusion processing (Lifted from Averous, 2004).

The extrusion and viscous behavior of molten PLS is known to depend on

temperature, moisture content, and thermomechanical treatment (McMaster et al., 1987;

Parker et al., 1990; Pdmanabhan and Bhattacharya, 1991). Viscosity data and rheological

models that take into account these variables have been reported in literature. Martin et al.

(2003) have summarized the main rheological studies performed on starch, including the

measurement technique and models used. A thermoplastic like behavior of low hydrated

starch, with an Arrhenius dependence on temperature and similar for moisture content was

reported. Conversely, structural modifications of starch, which affect the viscosity of the

product, are reflected differently by specific mechanical energy (SME), by the screw speed,
20

or even by the extruder barrel pressure, which depend on the machine characteristics. This

discrepancy underlines the need to ascertain the dependence of starch melt viscosity upon

structural factors or variables directly involved in its transformation.

Lai and Kokini (1990) studied the rheological properties of high amylose (70%)

and high amylopectin (98%) cornstarches, and showed the strong influence of the

processing history undergone by products prior to viscosity measurements. Zheng and

Wang (1994) identified the contribution of shear and thermal energies in the conversion of

waxy cornstarch. Della Valle et al. (1994) took into account the starch transformation and

the resulting macromolecular degradation, evaluated by chromatography (SEC profiles) and

intrinsic viscosity, by confirming the importance of the term noted β. This latter is reported

in the following equation, which presents a pseudo-plastic model with the pseudo-plastic

index (m) and the consistency (K). The influence of SME on the degradation of starch

products was evaluated by intrinsic viscosity [η] measurements. The gradual decrease of [η]

with increasing SME confirmed that macromolecular degradation occurred (Martin et al.,

2000)

where MC is the moisture content; GC is the glycerol content; α, α’, and β, are

dimensionless coefficients.

A common difficulty of rheological studies is that a thermomechanical treatment is

needed to obtain a homogeneous molten starch phase prior to measurement. This material

shows a behavior not totally thermoplastic but thermo-mechano-plastic (Martin et al.,

2003). Both, mechanical energy and temperature are required to obtain a molten material to

cause it to flow. Consequently, traditional rheometry (rotational, capillary) is not

appropriate for the study of rheological behavior of low moisture starches. In that respect,

extruder-fed slit rheometry is well adapted (Senouci and Smith, 1988; Bastioli et al., 1994).

Martin et. al. (2000) have also shown great interest in combining capillary and slit dyes on
21

an in-line viscometer attached to the head of an extruder with thermomechanical conditions

(shear rate, SME, and temperature) comparable to those used during the processing.

Finally, starch molts are commonly considered to exhibit viscoelastic behavior. The

measurement of elastic components of PLS molten phases, associated with the first normal

stress difference (N1), is not trivial because conventional rheometers do not allow to

perform mechanical treatment and cannot prevent volatilional of plasticizers. In a recent

study using plane-plate geometry, PLS was shown to behave mainly as a solid-like material,

because subjected to insufficient mechanical treatment (della Valle et al., 1998). As an

alternative, Senouci and Smith (1988) related the entrance and exit pressure losses in a slit

viscometer dye to the elastic properties of potato starch-based materials. Entrance and exit

pressure effects have been also used to evaluate the elasticity of melt of plasticized wheat

starch.

1.8.8 Plasticizers

Plasticizers are additives used to increase the flexibility or plasticity of a polymer.

Occasionally they are used only to facilitate the processing of a polymer and are not meant

to permanently change the flexibility of the polymer. Subsequently, a volatile plasticizer is

used so that it will evaporate once the final product is produced (Daniels, 1989). The smell

associated with new plastic is often caused by volatile plasticizers (Rosen, 1982).

The mechanisms by which plasticizer work is to intersperse themselves around the

polymer and interfere with the chain-chain secondary bonding (Nolan-ITU Pty, 2002). This

interference reduces inter-chain bonding strength and allows the chain segments greater

mobility resulting in increased flexibility and decreased tensile strength. Usually, a

plasticizer is chosen for its solubility in the polymer, since a soluble plasticizer can

penetrate the polymer more easily than a non-soluble one. Plasticizers can also be of the

non-soluble type and merely mechanically dispersed throughout the polymer (Daniels,

1989).

Most liquid plasticizer are low molecular weight organic materials with a glass

transition temperature in the range of 125-225K. When this type of solvent as added to an
22

organic polymer the result is a weight averaging of the glass transition temperature between

that of polymer and that of plasticizer if they are miscible. The reduction of the glass

transition temperature of the polymer by the plasticizer under the above conditions is

usually about 5K per weight percent of plasticizer (Averous, 20002).

1.8.9 Plasticizer in Thermoplastic Starch Materials

To create thermoplastic materials from such renewable resources as starch, there is

a need to incorporate plasticizers. To plastify the starch, it is necessary to use a large

quantity of plasticizers or destructing compounds, such as glycerol and carbamide among

others, with hydrophilic properties. A plasticizer such as glycerol as added in a ratio

ranging from 20% to 40% to the starch weight. The plasticizer content is directly related to

the mechanical properties and glassy transition of the material (Vilpoux and Averous,

2004).

In a study by Ma and Yu (2004-a), a mixed plasticizer for the preparation of

thermoplastic starch was employed. According to Hulleman et al. (1998) as mentioned in

the study, during the thermoplastic process, water in starch and other plasticizers play a

crucial role in that the plasticizer could form the hydrogen bonds with starch, taking the

place of the strong action between hydroxyl groups of starch molecules, plasticizing the

starch. In their study, formamide was first used as plasticizer, which thereby suppressed the

retrogradation of the thermoplastic starch. However, it was weaker than conventional

glycerol-plasticized thermoplastic starch. When urea was introduced, the retrogradation and

mechanical properties ware ameliorated.

In the said authors’ another study regarding the role of plasticizers in the properties

of thermoplastic starch, they have emphasized the role of hydrogen bonding in the

plasticization of starch. Moreover, in another study, several plasticizers containing amide

groups were used. Such compounds were believed to form stronger hydrogen bonds with

the hydrogen in C-O-H bonds of starch, contrary to the belief that the oxygen
23

in such starch bonds would form hydrogen bonds with the plasticizers (Ma and Yu, 2004-

b).

1.8.10 Related Studies involving Sago Starch

In a study by Pranamuda and Tokiwa (1996), PCL-corn, cassava, and sago

thermoplastic starch blends have reported success. Results of the study have indicated that

both tensile strength and elongation of the blends decreased as the starch content increased.

Zhai et al. (2003) incorporated polyvinyl alcohol (PVA) in corn starch-based sheets under

ionizing radiation, improving the wet strength of flexibility of the plastic.

Nevertheless, thermoplastic starch with no other blended polymers have been

successful, and were in fact patented in the United States of America as sausage casings and

food package from thermoplastic corn starch, with infusion of certain cross-linking agents

and proteins (Hammer et al., 1999).

Studies on sago starch as components of thermoplastic starch are not reported so

far, but esterification processes to alter the flexibility of gelatinized sago starch have been

done. A study by Igura et al. (2007) have reported that esterified sago starch extraction

residue containing a sizeable starch portion have been converted to biodegradable plant pots

with improved thermal and decomposition characteristics.

Jeroen et al. (1995) have stated that thermoplastic starch materials have been

prepared by kneading, extrusion, compression molding, and injection molding of several

native starches with the addition of glycerol as plasticizer. Nonetheless, in a recent study,

Thunwall et al. (2007) have established the possible routes for film blowing thermoplastic

potato starch on a laboratory scale by a suitable choice of processing conditions, amount of

glycerol, and moisture content. However, difficulties such as sticky film surface formation,

weak properties, and foaming were encountered.

It has been observed that sago starch materials plasticized with glycerol pose a

significant vulnerability to moisture. To address this, Carvalho et al. (2005) have treated

films with isocyanates, epoxy functions, and stearoyl chloride to lessen their sensitivity to
24

moisture. The results have showed that, in all instances, the polar contribution in the films

was considerably reduced, indicating the increase in surface hydrophobicity of the films.

In a study by Sitohy and Ramadan (2001), thermoplastic films were prepared from

corn starch phosphormonoesters to address issues on degradability and water disintegration.

Using a method by Fishman et al. (1996), films were cut into squares from a thin layer

plate, and were submerged in water at 25C that was continually stirred. It was found out

that the phosphorylated starches were easily disintegrated than the non-phosphorylated

ones, thereby producing thermoplastic films with enhanced degradability.

Using a method by Fishman et al. (1996), films were cut into squares from a thin

layer plate, and were submerged in water at 25 degrees Celsius that was continually stirred.

It was found out that the phosphorylated starches were easily disintegrated than the non-

phosphorylated ones, thereby producing films with enhanced degradability.

S-ar putea să vă placă și