You are on page 1of 9

UNIVERSITY OF GHANA DEPARTMENT OF NUTRITION

AND FOOD SCIENCE

Properties and gelatinization of starch

Nuamah Aboagye K M - 10358836

Lecturer: Prof G. S. Ayernor


9/12/2010
STARCH PROPERTIES

Next to cellulose, starch is the second largest biomass on earth. It is a predominant food
reserve for plants. Seeds, roots, tubers, stems, fruits and leaves are some important food
reserves for plants. Starch is primarily used in food industries to impart desirable functional
qualities to food. Starch is a polysaccharide produced by mostly higher order plants as a
means of storing energy. Their botanical source influences starch properties such as granule
shape, size, morphology, composition and molecular weight. It is stored intra cellularly in the
form of spherical granules ranging from 1 μm to more than 100 μm and shapes can be regular
(e.g., spherical, ovoid, or angular) or quite irregular. Starches have wide spectra of
application and depending on particular form of product required dictate the type of starch to
be used.

Chemically, starch is a polymeric carbohydrate consisting of anhydroglucose units linked


together primarily through α 1→4 glucosidic bonds. Although the detailed microstructures of
different starches are still being elucidated, it has generally been established that starch is a
heterogeneous material containing two microstructures, linear amylose, with some suggested
evidence of branches in its structure and branched amylopectin. Amylose is essentially a
linear structure of 1→4 linked glucose units, and amylopectin is a highly branched structure
of short 1→4 chains linked by 1→6 bonds. Again, colour formation upon reaction of starch
with iodine can be used to distinguish between amylose and amylopectin. Blue colouration
indicates the presence of amylose while red colouration tells amylopectin presence. The
moisture content of native starch granules is usually about 10%. Amylose and amylopectin
make up 98–99% of the dry weight of native granules, with the remainder comprising small
amounts of lipids, minerals, and phosphorus in the form of phosphates esterified to glucose
hydroxyls. Complexes between amylose and lipids, such as fatty acids, phospholipids and
monoacylglycerides, can significantly modify the properties and functionality of starch in
ways that are of interest to the food industry and for human nutrition. For example, a complex
of starch with lipids reduces the solubility of starch in water, alters the rheological properties
of pastes, decreases swelling capacity, increases gelatinisation temperature, reduces gel
rigidity, retards retrogradation and reduces the susceptibility to enzymic hydrolysis (Crowe,
Seligman, & Copeland, 2000; Kaur & Singh, 2000; Ozcan & Jackson, 2002; Tufvesson,
Wahlgren, & Eliasson, 2003a, 2003b Amylopectin is larger than amylose in most normal
starches and their chains are classified as small chains, with an average degree of
polymerization (DP) of about 15, and large chains, in which the DP is larger than 45. This
unique configuration ordered in the packing arrangement contributes to the crystalline nature
of the starch granule. This crystallinity reflects the organization of amylopectin molecules
within the starch granules, whereas amylose makes up most of the amorphous materials that
are randomly distributed among the amylopectin clusters (Blanshard, 1987). Amylose is
located in the low density layers of the growth rings, although amylose molecules are also
considered to be interspersed between amylopectin in the crystalline layers, disrupting the
crystal packing of amylopectin (Jane, 2006; Kozlov, Blennow, Krivandin, & Yuryev, 2007;)

2
Amylopectin chains with more than 10 glucose units are organized into double helices, which
are arranged into either A- or B-crystalline forms that may be identified by characteristic X-
ray diffraction spectral patterns. Different starch granules have also different x-ray diffraction
patterns: A, B and C (combination of A+B). X-ray diffraction studies show that the starch
granule is a mixture of crystalline and amorphous regions. The crystalline region consists of
starch in the form of a double helix which is very stable due to hydrophobic and Hydrogen
bonds and hence making it more insoluble in water at room temperature. This means that
more energy is required to break the bonds.

The double helical structures within the A- and B-type crystalline forms are essentially the
same (Imberty, Bule´ on, Tran, & Perez, 1991), but the packing of the helices in the A-type
crystalline structure is more compact than in B-type crystallites, which have a more open
structure with a hydrated core. The A-type crystal pattern has amylopectin molecules with
shorter chains (Jane, 2006). Cereal starches tend to have the A-type pattern, whereas tuber
starches and amylose-rich starches yield the B-type pattern, although both types may occur
together. Legume, root and some fruit and stem starches yield an intermediate C-type pattern,
but whether this is a mixture of the A- and B-type patterns or a distinct form is not clear
(Buleon et al., 1998). Consequently, the structure of a starch granule is influenced in part by
the fine structures, plant source and ratios of amylose and amylopectin molecules. Finally, the
granule structure in turn determines the accessibility of the starch structure to water and
chemical reagents, affecting molecular reaction patterns and properties of modified starches.

Granual % amylose Average x-ray Starch-P%


diameter content amylopectin
(um) chain length
Normal maize 5-20 23 24.4 A Trace
Wheat 22-36 26 22.7 a 0.001
Rice 3-8 21 22.7 A 0.013
Waxy maize 5-18 <1 23.5 A 0.002
Waxy rice 3-8 <1 18.8 A 0.003
Potato 15-75 18 29.4 B 0.089
Sweet potato 5-25 23 26.3 c 0.011
tapioca 5-25 18 27.6 A 0.008
Mung bean 10-27 31 24.8 A 0.011
High-amylose 6-15 40 30.7 B 0.013
maize-7

Table 1: Properties of some starch sources

Functional properties of starch are influenced by many factors. More hydrogen bonds require
higher energy to break bonds and hence a rise in gelatinization temperature. Again, starch
exhibits low gelatinization temperature, higher viscosity and clarity of starch paste when
more phosphates groups are present in starches. This can be attributed to the electrostatic
repulsion due to the charge phosphate group. Amylose and amylopectin ratio influences
starch properties in many diverse ways. For example, waxy starch of 0-8% amylose disperses
more easily, produces a clear viscous paste rather than gelling, normal starches 20-30%

3
amylose have high tendency to form gels, >50% amylose (high amylose starch) have very
high tendency to form gels.

STARCH GELATINIZATION

Starch has been used as an important raw material in biodegradable plastics, and its
gelatinization process has attracted much attention since it represents an important and unique
characteristic in the processing of starch-based materials (Svensson & Eliasson, 1995; Yu &
Christies, 2005). Starch gelatinization involves granule melting in an aqueous medium under
heating. In water, granule swelling increases with temperature and it leads to a transfer of
water in the suspension to water associated with starch components (amylose and
amylopectin) at 400c. When starch temperature reaches 60–70ºC, insoluble granules are
disrupted by the energy supplied breaking hydrogen bonds, leaking out amylose unit,
resulting in a loss of molecular organization and, consequently, loss of its crystallinity. This
process leads to increasing viscosity and starch solubilization, which is a result of irreversible
changes such as the disruption of granular and semicrystalline structure, also seen as a loss of
birefringence -maltese cross disappears (Douzals et al., 1996). However, optimum
gelatinization occurs at 900c and further heating decreases the viscosity. Differences in
swelling among native starches have been attributed to interplay of factors such as granule
size, crystallinity, amylose–lipid complex content, and interaction among starch chains in the
amorphous region. Many factors can affect gelatinization, for example if water is too low
then gelatinization will not occur, presence of sugar or salt competes for water hence
delaying gelatinization.

Fig1: Schematic representation of starch gelatinizing

4
On cooling, starch generally forms a viscoelastic, firm and rigid gel or precipitate. Starch
molecules begin to align and form double helixes making it less soluble, a process called
retrogradation in which the starch chains tend to re-associate in an ordered structure. The re-
association to form a gel is held together by intermolecular interactions involving amylose
and amylopectin molecules. Bread retrogrades and stales as it cools after baking. In the
process of baking, amylose leaches out and retrogrades almost fully on cooling by giving an
elastic tender crumb. Secondly, amylopectin slowly re-associate upon storage leading to
crumbing and staling of bread. Freezing inhibits retrogradation and occurs faster during
refrigeration. It is followed by another rise in viscosity when temperature is decreased,
usually referred to as setback. The peak time and peak viscosity are indicative of the water-
binding capacity of the starch and the ease with which the starch granules are disintegrated,
whereas higher setback values are usually correlated with the amylose content of the starch.
In gels that contain about 25% amylose, the starch molecules form a network resulting in a
firm gel, in contrast to waxy starch gels, which are soft and contain aggregates but no
network (Tang & Copeland, 2007b). Retrogradation is an ongoing process occurring over an
extended period. Amylose retrogrades over minutes to hours and amylopectin over hours to
days, depending on the ability of the branched chains to form associations. The retrogradation
of amylose in processed foods is considered to be important for properties relating to
stickiness, ability to absorb water, and digestibility, whereas retrogradation of amylopectin is
probably a more important determinant in the staling of bread and cakes. In rice, amylose
contributes texture and stickiness, whereas gelatinization temperature, cooking and pasting
properties are more closely related to amylopectin (Tran, Okadome, Murata, Homma, &
Ohtsubo, 2001). However, pasting profile of previously gelatinized starch does not show a
viscosity increase when exposed to changes in temperature (Rosa, Guedes, and Pedroso,
2004). The physicochemical properties of amylose and amylopectin are quite diverse and
they contribute in different ways to the pasting properties of starch. Size of the linear
molecule, amylose/amylopectin ratio, rate of cooling, temperature and interfering molecules
(fatty acids, emulsifiers and surfactants are some factors affecting retrogradation.
Understanding starch thermal transitions and starch rheological behaviour is essential for
studying starch structure, as well as for proposing new starch applications. Various
technologies have been developed and used to study gelatinization processes, such as
microscopy with hot-stage, viscometry, differential scanning calorimetry (DSC), X ray
diffraction, nuclear magnetic resonance and viscosity behaviour of starches. The most widely
used technique is DSC, which has proven to be an extremely valuable tool in quantifying the
gelatinization of starch, and has been widely used since the 1970s to study the thermal
behaviour of starches (Liu, Yu, Xie, & Chen, 2006; Shogren, 1992; Tananuwong & Reid,
2004; Tufvesson, Wahlgren, & Eliasson, 2003; Yu & Christie, 2001). Changes in starch
viscosity are evaluated using a Brabender amylograph . Starch slurry is heated (and cooled)
at a constant rate with constant stirring and viscosity measured. Information from graph is
useful in selecting the right starch type for an application. E.g high peak viscosity indicates
that the starch type requires more energy to stir and manipulate.

5
Fig. 2: Brabender amylograph of starch

SOURCES AND APPLICATION

Starches are the major storage polysaccharides in foods of plant origin. Worldwide dry starch
production is actually estimated at more than 64 million tons and almost 75 million tons are
expected by 2012 (Patil, 2009). Current annual production for primary starch sources is
estimated to be 46.1 million tons of corn, 9.1 million tons of cassava, 5.15 million tons of
wheat, and 2.45 million tons of potato (Roper and Elvers, 2008). Cereal starches have lower
moisture content, longer storage times and their starch extraction is easier and faster than in
roots and tubers. For example, in cassava, high water consumption is a critical factor in the
starch extraction process. In this process, water is used during the grinding, decantation, and
washing steps and these large quantities of water are converted to wastewater, which must be
treated before being released to the environment. Liquid waste has a high biochemical and
chemical oxygen demand; its treatment comprises several steps and requires a long retention
time. After extraction and separation, starch moisture content is from 35 to 40%, requiring a
great deal of energy in the drying process (Sriroth et al., 2000). Although cassava production
is growing and starch from cassava competes with the corn processors (Patil, 2009), corn still
remains the main starch source in the world, followed by cassava, potato, and wheat. More
than 70% of starch produced in the world is from corn (Roper and Elvers, 2008). However,
starch from roots and tubers shows some particular rheological and physical properties, such
as clear gel, high viscosity, and lower retrogradation, which are required in the formulation of
specific products. Demand for cassava starch has grown in the past few years and it is
actually the most widely traded form of native starch in the world, mainly in Thailand and
East Asia. Asia contributes around 90% of cassava starch produced for use in industry (LMC,
2008), with Thailand being the major producing country, followed by China and Indonesia
(FAO, 2008). In 2006, around 3.5 × 106 tons of cassava starches was produced in Thailand.
Of this amount, 2.3 × 106 tons was exported, with 1.67 × 106 tons as native starch and 638 ×
103 tons as modified starches (Roper and Elvers, 2008). In South America, which is

6
responsible for the other 10% of cassava starch production, Brazil is the main producer
(FAO, 2008). Considering starches from roots and tubers, potato starch is the second largest
starch source. In Europe, strong support of the grain sector has resulted in decreasing
production of potato starch (LMC, 2008). However, European countries are still responsible
for 80% of potato starch production in the world (LMC, 2008), with the Netherlands, France,
Belgium, Germany, and Switzerland as the main potato producers (FAO, 2008).

Due to its low cost, availability, and ability to impart a broad range of functional properties to
food and non food products (BeMiller, 2007), starch is utilized in several industrial
applications apart from its being used as a food ingredient. LMC, 2008 reported 57% of
produced starch consumed in the food industries and 43% in the non food sector. Starch
contributes greatly to the textural properties of many foods and has many industrial
applications as a thickener, colloidal stabilizer, gelling agent, bulking agent, water retention
agent and adhesive. Starch hydrolysis has been used to convert starches into Dextrins,
sweeteners, and syrups which are used in candy, sweet, chocolate, cake, dairy, dessert and
pastry products due to their anticrystallizing, sweetening, and water-holding properties. From
total worldwide starch production, excluding starch used to produce syrups and fermented
products, more than 30 million tons of natural and modified starches are used in the
production of dextrin, food ingredients, paper coatings, and adhesives (LMC, 2008). In
Europe, 57% of produced starch is used in sweeteners and hydrolyzed, 23% as native starch,
and 20% as modified starch. In European industries, the use of starch in sweet, drink, and
fruit processing is equivalent to 30% of produced starch. Convenience foods, bakery, food
ingredients, and dairy products account for 27%; the paper industry consumes 28%; and
chemical, fermentation, and other industrial products utilize 14%. Starch usage in feeds is 1%
(Roper and Elvers, 2008). However, in the United States most production is not destined for
native and modified starch products, but for sweeteners, particularly high fructose corn syrup
(LMC, 2008). More than 73% of corn starch produced in North America is used in U.S.
production of refinery products such as glycose syrups, high fructose syrups, syrup solids,
maltodextrins, and fructose (Roper and Elvers, 2008). In addition to the demand for starch to
produce syrups and dextrin, the demand for modified starches has grown in past years.
Modified starches are produced mainly in North America and Europe. Europe is still one of
the leading producers of modified starches and its paper and food industries tend to be at the
forefront in their use of innovative new products (LMC, 2008).

Conclusion
Plants sources of starch differ significantly in physicochemical, rheological, thermal,
gelatinization and retrogradation properties. Starches with specific functional properties are
in great demand in the food industry. Starches with desirable functional properties could play
a significant role in improving the quality of different food products and could replace
chemically modified starches that are currently being used in a number of products.

7
References

BeMiller, J.N. 2007, Carbohydrate Chemistry for Food Scientists St. Paul, MN: AACC
International, Inc.

Bule´ on, A., Colonna, P., Planchot, V., & Ball, S. (1998), Mini review: Starch granules
structure and biosynthesis, International Journal of Biological Macromolecules, 23, 85–112

Crowe, T. C., Seligman, S. A., & Copeland, L. (2000), Inhibition of enzymic digestion of
amylose by free fatty acids in vitro contributes to resistant starch formation, J. Nutrition, 130,
2006–2008.

Douzals, J.P., Marechal, P.A., Coquille, J.C., and Gervais, P. 1996. Microscopic study of
starch gelatinization under high hydrostatic pressure, J. Agric. Food Chem. 44:1403–1408

Imberty, A., Bule´ on, A., Tran, V., & Perez, S. (1991), Recent advances in knowledge of
starch structure. Sta¨rke. Starch properties, modifications, and applications, J. Macromol.
Sci. Pure Appl. Chem. A32:751–757.

Jane, J. L. (2006), Current understanding on starch granule structures, Journal of Applied


Glycoscience, 53, 205–213

Kaur, K., & Singh, N (2000), Amylose–lipid complex formation during cooking of rice flour,
Food Chemistry, 71, 511–517.

Kozlov, S. S., Blennow, A., Krivandin, A. V., & Yuryev, V. P. (2007), Structural and
thermodynamic properties of starches extracted from GBSS and GWD suppressed potato
lines. International Journal of Biological Macromolecules, 40, 449–460.

Liu, H., Yu, L., Xie, F. W., & Chen, L. (2006), Gelatinization of cornstarch with different
amylose/amylopectin content, Carbohydrate Polymers, 65, 357–363

LMC 2008, LMC International’s Global Markets for Starch and Fermentation Products
Database 2008 report, http://www.lmc.co.uk

Shogren, R. L. (1992), Effect of moisture content on the melting and subsequent physical
aging of cornstarch, Carbohydrate Polymers, 19

Ozcan, S., & Jackson, S. D. (2002), Impact of thermal events on amylose–fatty acid
complexes, Sta¨rke [Starch], 54, 593–602

Patil, S.K. 2009, Global Modified Starch Products Derivatives and Markets A Strategic
Review. February 1, 66 pp

Roper, H. and Elvers, B. 2008, Starch 3 Economic Aspects, In: Ullman’s Encyclopedia of
Industrial Chemistry (7th ed.). New York: John Wiley & Sons, pp. 21–22.

Rosa, D.S., Guedes, C.G.F., and Pedroso, A.G. 2004, Gelatinized and non-gelatinized corn
starch/poly blends: Characterization by rheological, mechanical and morphological
properties. Polímeros: Ciência e Tecnologia. 14:181–186.

8
Sriroth, K., Piyachomkwan, K., Wanlapatit, S., and Oates, C. (2000), Cassava starch
technology: The Thai experience. Starch/Stärke. 52:439–449.

Svensson, E., & Eliasson, A.-C (1995), Crystalline changes in native wheat and potato
starches at intermediate water levels during gelatinization. Carbohydrate Polymers, 26, 171–
176.

Tananuwong, K., & Reid, D. S. (2004), DSC and NMR relaxation studies of starch–water
interactions during gelatinization. Carbohydrate Polymers, 58, 345–358

Tang, M. C., & Copeland, L. (2007b), Investigation of starch retrogradation using atomic
force microscopy. Carbohydrate Polymers, 70, 1–7

Tran, U. T., Okadome, H., Murata, M., Homma, S., & Ohtsubo, K. (2001), Comparison of
Vietnamese and Japanese rice cultivars in terms of physicochemical properties, Food Science
and Technology Research, 7, 323–330

Tufvesson, F., Wahlgren, M., & Eliasson, A.-C (2003a), Formation of amylose–lipid
complexes and effects of temperature treatment, Part 1 Monoglycerides, Sta¨rke [Starch], 55,
61–71

Tufvesson, F., Wahlgren, M., & Eliasson, A.-C (2003b), Formation of amylose–lipid
complexes and effects of temperature treatment, Part 2 Fatty Acids, Sta¨rke [Starch], 55,
138–149

Yu, L., & Christie, G. (2001), Measurement of starch thermal transition using differential
scanning calorimetry, Carbohydrate Polymers, 46, 179–184.

Yu, L., & Christies, G. (2005), Effect of orientation on microstructure and mechanical
properties of starch-based materials. Journal of Materials Science, 40, 111–116.

You might also like