You are on page 1of 317

Title

Author(s)

Physical and chemical modification of some cereal, tuber and


root starches and the roles of{221}-cyclodextrin as a starch
modifyingagent
Anil Gunaratne, D. M.

Citation

Issued Date

URL

Rights

2006

http://hdl.handle.net/10722/50221

The author retains all proprietary rights, (such as patent rights)


and the right to use in future works.

PHYSICAL AND CHEMICAL MODIFICATION OF


SOME CEREAL, TUBER AND ROOT STARCHES AND
THE ROLE OF -CYCLODEXTRIN AS A STARCH
MODIFYING AGENT

D.M. ANIL GUNARATNE


Ph. D. THESIS

THE UNIVERSITY OF HONG KONG


2006

Abstract of thesis entitled


PHYSICAL AND CHEMICAL MODIFICATION OF SOME CEREAL,
TUBER AND ROOT STARCHES AND THE ROLE OF -CYCLODEXTRIN
AS A STARCH MODIFYING AGENT
Submitted by
D.M. Anil Gunaratne
for the degree of Doctor of Philosophy
at The University of Hong Kong
in May 2006
Functional and retrogradation properties, enzymatic digestibility and some molecular
characteristics of cereal, tuber and root starches were examined in a series of studies.
An analysis of the -cyclodextrin (-CD) : starch interaction showed that -CD
disrupted the amylose-lipid complex within the starch granules by complexing with
starch lipids resulting in increased amylose leaching, swelling factor and solubility.
Pasting, gelatinization, and gelling properties were also affected due to these
structural changes in the starch granule. -Cyclodextrin did not affect -amylase
hydrolysis or amylopectin retrogradation. Modified -CD (hydroxypropyl cylodextrin) proved more effective than -CD in disrupting the amylose-lipid
complex and inhibiting its formation.

Degree of acetylation was decreased by prior acid treatment. Acid treatment increased
the proportion of linear segments which are able to form double helical like structures
affecting gelatinization and retrogradation properties. Acid treatment increased the
gelatinization temperature and enthalpy but decreased viscosity. Gel hardness and
retrogradation were dependent on the concentration of acid used and on the type of
starch. Acetylation increased swelling, amylose leaching and viscosity but decreased

gel hardness and retrogradation. Acid treatment plus acetylation decreased the
viscosity. Introduction of acetyl groups to acid-thinned starch produced very soft gels.
The amylose-lipid complex showed a greater resistance to acid degradation.

Heat-moisture treatment before hydroxypropylation increased the degree of


modification and facilitated increased derivatization in the crystalline region,
producing very high final viscosity with greater pasting stability. Heat-moisture
treatment had no influence on transition parameters of the amylose-lipid complex.

Some mixtures of potato and amaranth starch showed very specific non-additive
pasting characteristics when swelling, and amylose leaching of individual
components in the mixture differed substantially. Each component gelatinized
independently in the mixture. Pasting properties of some mixtures showed high
resistance to shear thinning providing a potential alternative to chemical modification.

Some under-exploited tuber and root starches showed high resistance to shear
thinning and formed firm gels similar to cross-linked starches. Hydroxypropylation
increased swelling, amylose leaching, susceptibility to enzymatic hydrolysis, and
viscosity but decreased gelatinization parameters, gel hardness and amylopectin
retrogradation. Cross-linking decreased swelling ability and amylose leaching but
increased resistance to shear thinning and enzymatic hydrolysis. Cross-linking had no
influence on gelatinization and amylopectin retrogradation. Hydroxypropylation of
cross-linked starches produced a wider range of desirable functional properties.

Results of -cyclodextrin studies are of technological importance because


dissociation and formation of the amylose-lipid complex affects the quality of many

starch-based food products. Mixing of starches with different swelling and amylose
leaching characteristics and prior acid and heat-moisture treatment before acetylation
and hydroxypropylation respectively can be used to produce some novel starch
properties. Some excellent properties shown in some of the under-exploited starches
could permit direct utilization of these starches without modification.

PHYSICAL AND CHEMICAL MODIFICATION OF SOME CEREAL,


TUBER AND ROOT STARCHES AND THE ROLE OF -CYCLODEXTRIN
AS A STARCH MODIFYING AGENT

By

D.M. Anil Gunaratne


B.Sc. (Agric.) University of Peradeniya, Sri Lanka
M.Sc. (Food Science) Memorial University of Newfoundland, Canada

A thesis submitted in partial fulfillment of the requirements for


the Degree of Doctor of Philosophy
at The University of Hong Kong.
May 2006

DECLARATION
I declare that this thesis represents my own work, except where due
acknowledgement is made, and that it has not been previously included in a thesis,
dissertation or report submitted to this University or to any other institution for a
degree, diploma or other qualification.

Signed ______________________________
D.M. Anil Gunaratne

ACKNOWLEDGEMENTS
I wish to express my sincere appreciation to my thesis supervisor Dr. H. Corke for his
support, expert supervision, and advice and for the given freedom of thinking
throughout my studies.

Grateful appreciation is also extended to the University of Hong Kong Committee on


Research and Conference Grants for providing financial assistance.

I also thank Dr. Yizhong Cai and my fellow graduate students with whom I have been
privileged to associate these past years. As we have learned together, their
camaraderie has made life here interesting and most enjoyable.

Thanks are also extended to all members of Dept. of Botany for their friendly support
extended throughout my studies. Technical assistance provided by Mrs. Ming Wei
Fong and Mr. Joe Mock is also gratefully appreciated.

I would like to express my sincere thanks to the University of Sabaragamuwa, Sri


Lanka for granting me study leave to undertake my postgraduate studies at The
University of Hong Kong.

I greatly appreciate my parents, family members, wife (Chandrika) and daughter


(Imasha) for their enormous continuous support and encouragement extended in the
completion of this course of study.

II

CONTENTS
Abstract
Declaration
Acknowledgements
Table of contents

I
II
III

Chapter 1 Introduction and Review of Literature


1.1 Introduction

1.2 Review of literature

1.2.1. Starch-general introduction

1.2.2. Starch granules characteristics

1.2.3. Major components of starch and their structures

1.2.3.1. Amylose

1.2.3.1.1. Complex formation of amylose and amylose-lipid interaction 9


1.2.3.2. Amylopectin
1.2.4. Minor components of starch

11
13

1.2.4.1. Starch lipids

13

1.2.4.2. Phosphate

14

1.2.4.3. Protein

15

1.2.5. Starch structure

15

1.2.5.1. Amorphous region of starch granule

15

1.2.5.2. Semi-crystalline structure of starch granule and crystallinity

16

1.2.6. Functional properties of starch

17

1.2.6.1. Swelling and solubility

18

1.2.6.2. Gelatinization

19

1.2.6.2.1. Effect of water content on starch gelatinization

21

1.2.6.2.2. Effect of polyhydroxy compounds on starch gelatinization

22

1.2.6.3. Pasting

23

III

1.2.6.3.1. Effect of polyhydroxy compounds on starch pasting


1.2.6.4. Retrogradation

27
28

1.2.6.4.1. Effect of retrogradation on quality of starch-based food


products

30

1.2.6.4.2. Measurement of amylopectin retrogradation using DSC

30

1.2.6.5. Properties of starch gel

31

1.2.7. Starch hydrolysis by -amylase

32

1.2.8. Starch modification

34

1.2.8.1. Physical modification of starch

36

1.2.8.1.1. Heat-moisture treatment

36

1.2.8.2. Chemical modification of starch

38

1.2.8.2.1. Acetylation

38

1.2.8.2.2. Hydroxypropylation

39

1.2.8.2.3. Cross-linking

42

1.2.8.2.4. Acid modification

43

1.2.9. Cyclodextrin

46

1.2.10. References

50

Chapter 2 Effect of Cycloheptaamylose (-cylodextrin) on Physicochemical


Properties of Cereal, Tuber and Root Starches
Abstract

70

2.1. Introduction

71

2.2. Materials and Methods

72

2.2.1. Materials

72

2.2.2. Methods

72

2.3. Results and Discussion

75

2.4. Conclusions

83

2.5. References

83

IV

2.6. Tables

86

2.7. Figures

90

Chapter 3 Influence of Unmodified and Modified Cycloheptaamylose


(-Cyclodextrin) on Physical Properties and Enzymatic
Hydrolysis of Cereal, Tuber and Root Starches
Abstract

96

3.1. Introduction

98

3.2. Materials and Methods

99

3.2.1. Materials

99

3.2.2. Methods

99

3.3. Results and Discussion

103

3.4. Conclusions

111

3.5. References

111

3.6. Tables

114

3.7. Figures

118

Chapter 4 Thermal, Pasting and Gelling Properties of Wheat and Potato


Starches in The Presence of Sucrose, Glucose, Glycerol and
Hydroxypropyl -Cyclodextrin
Abstract

123

4.1. Introduction

124

4.2. Materials and Methods

125

4.2.1. Materials

125

4.2.2. Methods

126

4.3. Results and Discussion

127

4.4. Conclusions

138

4.5. References

138

4.6. Tables

141

4.7. Figures

145

Chapter 5 Functional Properties and Retrogradation of Heat-Moisture


Treated Wheat and Potato Starches in The Presence of
Hydroxypropyl -Cyclodextrin
Abstract

150

5.1. Introduction

151

5.2. Materials and Methods

152

5.2.1. Materials

152

5.2.2. Methods

152

5.3. Results and Discussion

154

5.4. Conclusions

161

5.5. References

162

5.6. Tables

164

5.7. Figures

169

Chapter 6 Physical Properties and Digestibility of Acetylated Wheat, Potato,


Waxy Maize and High-Amylose Maize Starches as Affected by
Hydroxypropyl -Cyclodextrin
Abstract

173

6.1. Introduction

175

6.2. Materials and Methods

176

6.2.1. Materials

176

6.2.2. Methods

176

6.3. Results and Discussion

180

6.4. Conclusions

187

6.5. References

187

6.6. Tables

190

6.7. Figures

195

Chapter 7 Influence of Prior Acid Treatment on Acetylation of Wheat,


Potato and Maize Starches
Abstract

200
VI

7.1. Introduction

201

7.2. Materials and Methods

202

7.2.1. Materials

202

7.2.2. Methods

203

7.3. Results and Discussion

205

7.4. Conclusions

213

7.5. References

213

7.6. Tables

216

7.7. Figures

220

Chapter 8 Effect of Hydroxypropylation on Some Structural and


Physicochemical Properties of Heat-Moisture Treated Wheat,
Potato and Waxy Maize Starches
Abstract

224

8.1. Introduction

225

8.2. Materials and Methods

227

8.2.1. Materials

227

8.2.2. Methods

227

8.3. Results and Discussion

229

8.4. Conclusions

235

8.5. References

236

8.6. Tables

239

8.7. Figures

242

Chapter 9 Gelatinizing, Pasting and Gelling Properties of Potato and


Amaranth Starch Mixtures
Abstract

247

9.1. Introduction

248

9.2. Methods and Materials

249
VII

9.2.1. Materials

249

9.2.2. Methods

249

9.3. Results and Discussion

252

9.4. Conclusions

261

9.5. References

261

9.6. Tables

264

9.7. Figures

271

Chapter 10 Functional Properties of Hydroxypropylated, Cross-Linked,


and Hydroxypropylated Cross-Linked Tuber and Root Starches
Abstract

274

10.1. Introduction

275

10.2. Materials and Methods

276

10.2.1. Materials

276

10.2.2. Methods

276

10.3. Results and Discussion

280

10.4. Conclusions

288

10.5. References

288

10.6. Tables

291

10.7. Figures

297

Summary and Direction for Future Studies

301

VIII

Chapter 1
Introduction and Review of Literature
1.1. Introduction
Functionality is the key to starch utilization in the wide range of food applications.
Properties such as specific viscosity, mouth-feel, freeze-thaw stability, clarity,
emulsion stability, color, film-forming properties and anti-caking are important.
When suitable characteristics do not appear in native starches, the alternative is some
form of value-added transformation. This may typically be by chemically or
physically modifying the native starch to alter its properties. Chemical modification
includes a large number of methods such as monosubstitution (etherification and
esterification),

cross-bonding,

oxidation

and

acid-thinning,

while

physical

modification is mainly performed by heat-moisture treatment, annealing and


pregelatinization. Although chemical modification is the most widely used technique
to increase the usefulness and value of starches, there is often difficulty in achieving
regulatory approval for new chemical reagents to be used and for higher levels of
treatment due to consumer safety concern (BeMiller, 1997). Therefore, it is worth
testing non-toxic chemicals and other possible alternatives for chemical modification
while aiming to achieve novel starch properties. Nowadays there is a great demand
for natural food components, avoiding as much as possible any chemical treatment.

Recently, there is a growing interest in -cyclodextrin (a non-toxic oligosaccharide


which has donut-shape molecular structure) in the pharmaceutical industry because of
its capability in making complexes with organic compounds such as fatty acids. It can
be speculated that -cyclodextrin can complex with native starch lipids altering the
structure and functionality of starch granules. However, very little information is
available on -cyclodextrin-starch interaction and there are no systematic studies
1

reported testing the influence of -cyclodextrin on starch structure and functionality.


Previous -cyclodextrin-starch studies were also limited mostly to a single source of
native starch and unmodified -cyclodextrin which is poorly soluble in water. Thus it
is worth studying -cyclodextrin-starch interaction using starches from diverse
botanical sources and modified -cyclodextrin (e.g., hydroxypropyl -cyclodextrin)
which is readily soluble in water. Furthermore, comparative study of -cyclodextrin
with other polyols such as sugars and with modifying the starch structure by physical
(heat-moisture) and chemical (acetylation) treatment before reacting with
cyclodextrin will provide more information on -cyclodextrin-starch interaction.

Mixing of different starches is another potential way to improve the properties of


native starches without using any chemicals. However, very little information is
available on the properties of starch mixtures and previous studies were confined only
to mixing native starches.

Some source of starches such as under-exploited tropical tuber and root starches have
not received much attention, but desirable functional properties may exist in those
starches. Investigation of properties of those starches may lead to their direct
utilization without any modification. Furthermore, employing dual modification on
those starches (hydroxypropylation and cross-linking) may result in novel starch
properties having a wider range of applications.

Control of reaction sites within the starch granule and the control of reaction sites on
starch molecules (amylose and amylopectin) are two higher level possibilities for
making novel starch products (BeMiller, 1997). In this respect, prior treatment like
acid-thinning and heat-moisture treatment before other chemical modification could

be employed to change the reaction site of the starch granules. For example, cleavage
of starch macromolecules by acid and reorientation and disruption of starch
crystallites by heat-moisture treatment may alter the reaction site of starch molecules
resulting in novel starch properties. Perhaps this may require lower concentration of
modifying chemicals to achieve the desirable functional properties. This area of study
is of significance, since there is a growing interest in investigating alternatives for
chemical modification and to minimize the use of undesirable chemicals in making
modified starch. Therefore, this research project was conceived with the following
objectives.
1. Investigate physicochemical properties of cereal, tuber and root starches in the
presence of -cyclodextrin.
2. Compare unmodified and modified -cyclodextrin (hydroxypropyl -cyclodextrin)
on functional and enzymatic digestibility of cereal, tuber and root starches.
3. Compare hydroxypropyl -cyclodextrin with other polyols (sucrose, glucose and
glycerol) on pasting, gelatinizing and gelling properties of starch.
4. Investigate how heat-moisture treatment can enhance the usefulness of
hydroxypropyl -cyclodextrin-starch interaction.
5. Use chemically modified (acetylated) starches to characterize the hydroxypropyl cyclodextrin-starch interaction.
6. Characterize some structural and functional properties of acid-thinned acetylated
starches.
7. Characterize some structural and functional properties of heat-moisture treated
starches before and after hydroxypropylation.
8. Investigate gelatinizing, pasting and gelling properties of native, heat-moisture
treated and cross-linked potato and amaranth starch mixtures.

9. Characterize properties of native, hydroxypropylated, cross-linked and


hydroxypropylated cross-linked starches of some underexploited tubers and roots.
1.2. Review of literature
1.2.1. Starch-general introduction
Starch can be found in all organs of higher plants (Shannon and Garwood, 1984) such
as pollen, leaves, stems, woody tissues, root, tubers, bulbs, rhizomes, fruits, flowers,
and the pericarp, cotyledons, embryo, and endosperm of seeds. Starch is the second
largest biomass product after cellulose produced on earth. The most important sources
of starches are cereal grains, pulses and tubers. The proportion of starch in the cereal
grain is generally between 60 to 75% by weight (Hoseney, 1986). For pulses and
tuber starches it ranges from 30% to 70% and 65 to 85% respectively. Starch is
predominantly processed in highly industrialized countries like the USA, EU
countries and Japan. Other starch-containing raw materials from which starch is
separated usually in small production units (mainly in Asia) are sago, palm, sweet
potato, arrowroot, amaranth, sorghum, lotus, taro, smooth pea, cassava, mung bean,
lentils, and wild rice. Tuber and root starches such as cassava, arrowroot, taro, sago,
sweet potato and yam serve as staple foods for people throughout the hot and humid
regions of the world. These crops are naturally suited to tropical agro-climatic
conditions and grow in great abundance with little or no artificial inputs. Although
agronomic properties of tropical tuber and root crops are well documented, starches
of these crops have not beenextensively studied for their potential industrial
applications.

1.2.2. Starch granules characteristics

The starch granule, the basic physical structural unit of starch, has a distinctive
microscopic appearance for each botanical source. Granule size and morphology have
received much attention recently, since size of granules are important in determining
the taste and mouth feel of some starch-based fat mimetics (Alexander, 1992). The
granules are partially crystalline, insoluble in cold water and their size, the shape, and
the composition are essentially genetically determined. Starch granules may be
spherical, oval, polygonal, disk and kidney shaped or elongated (Jane et al., 1994)
[Table 1]. In general, cereal starch granules are small and polyhedric, whereas, tuber
starch granules are large and spherical or ellipsoid. Most of the tuber and root
starches are simple granules, the exception being cassava and taro starches, which
appear to be a mixture of simple and compound granules (Hoover, 2001).

Starch granules are birefringent, indicating a high degree of internal order.


Birefringence is the ability to refract light in two directions that is evidenced by
distinctive patterns under a polarizing microscope.

The outer surface of the starch granules play important part in many applications of
starch, but there is a lack of definitive information on the nature of the starch surface
(Galliard and Bowler, 1987). When observed under a scanning electron microscope
the surfaces of all root and tuber starch granules appear smooth with no evidence of
any fissures (Hoover, 2001). However, Fannon et al. (1992) discovered pores on the
surface of corn, sorghum, and millet starch granules which are real anatomical
features of the native structure and not artifacts of drying, specimen preparation or
observation techniques. Surface pores on granules of corn, sorghum, and millet are
openings to channels that penetrate in a radial direction through the granule (Fannon
et al., 1993; Baldwin et al., 1994; Huber and BeMiller, 1997). Several researchers

(Gallant, 1973; Fuwa et al., 1997; Planchot et al., 1997) have postulated that pores on
the granule surface increase the accessibility of -amylase into the granule interior.
Planchot et al. (2000) have shown by dynamic light scattering, HPAEC-PAD and
polyethylene glycol molecular probes, that wet starch granules can be considered as a
porous substrate permeable to low molar mass solutes such as malto oligosaccharides
or small polyethylene glycols, whereas, molecules with a hydrodynamic radius
greater than 0.6 nm cannot penetrate.

1.2.3. Major components of starch and their structures

1.2.3.1. Amylose
The two major components of starch are amylose and amylopectin. Amylose, the
minor component, consists mainly of -(14) linked glucose units. The degree of
polymerization (DP) of this linear polymer is usually in the range of 500-600 units
(Jacobs and Delcour, 1998). However, it is now accepted that this polymer is slightly
branched, having occasional -(16) branch points. Hizukuri et al. (1981) postulated
that there are 9-20 branch points per molecule. The side chains range in chain length
from 4 to 100 (Hizukuri et al., 1981; Takeda et al., 1987). The extent of branching
depends on the origin of amylose (Takeda et al., 1987) and increases with its
molecular size (Banks and Greenwood, 1975; Greenwood and Thompson, 1959).
Evidence of the occurrence of branching points in amylose is its incomplete
conversion into maltose by -amylase; -amylolysis varies between 73% to 95%
depending on the extraction procedure and the botanical origin of amylose (Banks
and Greenwood, 1975; Morrison and Karkalas, 1990). The molecular weight of
amylose has been reported to vary between 105 and 106 Da (Morrison and Karkalas,
1990; Hizukuri et al. 1989). Despite its slight branching, amylose behaves essentially

like a linear polymer, forming films and complexes with ligands. Amylose isolated
from tuber and root starches, such as potato and tapioca have larger molecular sizes
than those isolated from cereal starches, such as maize, rice, and wheat (Takeda et al.,
1986). With linear chemical structure, amylose has the ability to change its
conformation. With many hydroxyl groups, there is a high hydrogen bonding
capability, with strong internal forces that permits these changes. An open channel in
the center of a helix permits complexing with other molecular species such as iodine
and fatty acids. However, the conformation of amylose has been the subject of
controversy and has been shown to vary from helical to an interrupted helix, to a
random coil. In alkaline solutions (KOH) and in dimethyl sulfoxide (DMSO) amylose
probably has an expanded coil conformation, while in water and neutral aqueous
potassium chloride solutions it is a random coil with short loose helical segments
(Banks and Greenwood, 1971). Jane and Robyt (1985) identified (using

13

CNMR)

expanded and compact helical conformations in aqueous amylose solutions in the


absence and presence of complexing agents, respectively. Comparison of amylose
content in starch of different maturities has suggested that amylose is more
concentrated at the periphery of the starch granule (Boyer et al., 1976). Blanshard
(1986) postulated that amylose is separated from amylopectin in the granules of
maize and wheat starches and is partly co-crystallized with amylopectin in potato
starch. Cross-linking of maize and potato starches and characterization of the
products by molecular sieve chromatography showed that amylose was cross-linked
with amylopectin (Jane et al., 1992; Kasemsuwan and Jane, 1994). These data
suggested that in granular starch, amylose molecules do not exist in the form of
bundles at the amorphous region but, rather, are interspersed among the amylopectin
molecules.

Table. 1. Some physicochemical properties of cereal, tuber and root starches


Starch source

Granular
size (m)

Granular shape

Amylose
(%)

Swelling factor Total lipids


at (80 C)
(%)

Potato

15-100

Oval, spherical

22

57.4

Potato (waxy)
Sweet potato

14-44
2-42

Cassava
Taro

5-35
3-35

Round, oval
Round,
oval
polygonal
Round to variable
Round to variable

Lotus
Buffalo gourd
Elephant yam
New cocoyam

15-40
2-24
3-30
10-50

Rod like, round


Oval to eliptical
Round, polygonal
Polygonal to variable

26

True yam (Dioscorea alata)


True yam (Dioscorea esculenta)
Oat

6-100
10-70
6-10

Oblong to oval
Round, oval
Polygonal, compound

28.5
30
19.4

Wheat

2-35

Round,
lenticular

Maize
Regular

5-25

Round, polygonal

Waxy

5-25

High amylose
Rice

and 19.1
22

0.19

Total
phosphorous
(%)
0.10

0.06-0.6
43
36

15.9
23.2

0.12
0.15

0.01

0.92-1.14

48 (ppm)
0.01-0.06

18

0.40

0.02

26

0.03

0.03
0.03

20.2

1.13

9.7

0.80

26

12

0.70

Round, oval identations

~1

70

0.05

4-16

Round

Up to 80

4.3

1.2

3-8

Polygonal, clusters

17

9.8

1.01

elliptical, 25

Reference
Moorthy, (1994); Gunaratne
and Hoover (2002)
McPherson and Jane, (1999)
Lim et al., (1994); Collado et
al., (1999)
Gunaratne and Hoover, (2002)
Lim et al., (1994); Gunaratne
and Hoover, (2002)
Suzuki et al., (1992)
Dreher and Berry, (1983)
Moorthy, (1994)
Gunaratne and Hoover, (2002);
Moorthy, (1994)
Gunaratne and Hoover, (2002)
Gallant et al., (1982)
Hoover and Vasanthan, (1992,
1994)
Hoover and Vasanthan, (1994);
Tester and Morrison, (1990a);
Collado and Corke, (2004)
Jayakody and Hoover, (2001);
Collado and Corke, (20040
Jayakody and Hoover, (2001);
Collado and Corke, (2004)
Jayakody and Hoover, (2001);
Collado and Corke, (2004)
Jayakody and Hoover, (2001);
Collado and Corke, (2004)

1.2.3.1.1. Complex formation of amylose and amylose-lipid interaction


Amylose can form films and complexes with ligands such as fatty acids. When
amylose forms complexes with various ligands the resulting ordered crystalline
structure is called V-polymorph (Biliaderis, 1998). Rappenecker and Zugenmaier
(1981) and Hinrichs et al. (1987) have shown that the chain conformation in Vamylose is a left handed single helix with six residues per turn and for complexes
with aliphatic alcohols and monoaceyl lipids the rise per monomer residue is
approximately 1.32-1.36 A0. However, when the ligand is bulkier than a hydrocarbon
chain, helices of seven or eight glucose residues per turn are also feasible (French and
Murphy, 1977). X-ray diffraction diagrams of granular starches do not usually show
the presence of V-structures, with the exception of wrinkled pea starch, amylo maize,
and some other maize genotypes (dull, su) all with amylose greater than 30% (Zobel,
1988a; Gernat, et al., 1993; Zobel, 1992). The lack of V-type characteristic peaks
upon X-ray analysis does not necessarily prove the absence of amylose-lipid
complexes. It merely indicates the absence of organized helices into well-defined
three-dimensional structures (Biliaderis, 1998). Recently,

13

CCP/MAS-NMR studies

provided the proof for the presence of V-conformation in granules of maize, oat,
barley, and wheat starches (Morgan et al., 1995; Morrison et al., 1993a). Gernat et al.
(1993) also showed the existence of V- structure related to amylose-lipid complex in
native starch granules by X-ray scattering studies on enzymatically degraded wheat
starch. Calorimetry and X-ray diffraction have been widely used to study the
amylose-lipid complex (Hoover and Hadziyev, 1981; Biliaderis et al., 1993; Karkalas
et al., 1995; Biliaderis et al., 1986b; Biliaderis and Senaviratne, 1990; Galloway et al.,
1989).

Monoacyl lipids that are found in the starch granule as non-starch lipids, granule
surface lipids, and internal lipids can complex with amylose (Buleon et al., 1998). A
typical amylose-lipid complex is composed of a left-handed amylose helix with six
residues per turn, in which the aliphatic part of the lipid is included, while polar group
lies outside (Karkalas et al., 1995) [Fig. 1]. Amylose can be complexed with lipids
when heat energy is supplied. Formation of amylose-lipid complex occurs during
heat-moisture treatment for naturally containing lipids (Kugimiya et al., 1980;
Kugimiya and Donovan, 1981) or when lipids are added to pure amylose which are
free of lipids (Biliaderis et al., 1985) or defatted starches (Biliaderis et al., 1986a) has
been demonstrated. Combining studies of X-ray diffraction and DSC has revealed
that amylose-lipid complexes can be crystalline (form II) which has V-pattern in Xray diffraction or amorphous (form I) that melts at a lower temperature in the DSC
(Biliaderis and Senaviratne, 1990). The formation of form II or form I depends on
the interactions between heating temperature, water content in the starch-lipid system,
type of ligand and extent of the amylose leached from the starch granule (Biliaderis,
1992; Tufvesson et al., 2003; Hoover, 1998; Le Bail et al., 1999).

The thermal properties of the amylose-lipid complex have been studied using DSC by
several researchers (Eliasson and Krog, 1985; Biliaderis et al., 1986a; Tufvesson et al.,
2003). Endothermic transition of wheat starch can be detected at temperature above
the gelatinization temperature range (Fig. 3) around 90-100 C at excess water
conditions (Eliasson, 1980). This transition is due to the transition of the amyloselipid complex (Kugimiya et al., 1980) and will be present in lipid-containing starches
(Fredricsson et al., 1998). The transition parameters of amylose-lipid complex depend
on the water content (Eliasson, 1980) and the nature of the polar lipid in the complex
(Eliasson and Krog, 1985; Kowblansky, 1985). At intermediate water content (~50%),

10

the melting temperature of these crystals is around 110 C. DSC characteristics of the
transition related to the amylose-lipid complex for some cereal starches are shown
(Table 2). In food systems complex formation of amylose with various types of lipids
are of interest because they can affect the functional properties of food. For example
decreased swelling, solubilization and thickening power of starch (Galliard and
Bowler, 1987), retardation of starch retrogradation and bread firming (Krog, 1971;
Biliaderis and Tonogai, 1991), prevention of stickiness of dried potato (Hoover and
Hadziyev, 1981), and improvement of structural integrity of cereal kernels during
cooking (e.g., parboiled rice) [Biliaderis et al., 1993].

1.2.3. 2. Amylopectin
Amylopectin is a branched molecule with -(14) linked glucose units in linear
chains and -(16) linked branched points (Greenwood, 1964; Lineback, 1984).
Average molecular weight polymer is 107-109 (Aberle et al., 1994). While the average
size of unit chains of amylopectin is 20-25 (Hizukuri, 1985), the amylopectin
molecule contains several distributions of chains differing in their chain length as A,
B, and C (Hizukuri, 1986). The unbranched A- chains are linked to B-chains and do
not carry any other chains, the B-chains (B1-B4), carry one or more A-chains and /or
B-chains, while the C-chain contains the reducing end group of the molecule.
According to Hizukuri (1985) the molar ratio of short to long chain varies between
3:1 and 12:1, depending on the botanical origin of starch. Generally cereal starches
contain shorter chain length in both long and short chain fraction compared to those
of tuber and root starches (Hizukuri, 1985; Hizukuri, 1986). Since amylopectin is a
branched molecule it is less prone to retrogradation, gelation and syneresis. As
Manners (1985) reported, in his proposed amylopectin model, branched molecules

11

Fig. 1. Schematic illustration of amylose-lipid inclusion


Source: Carlson et al. (1979).

12

can be involved in the formation of alternating crystalline and amorphous regions of


the granule. The packing of clustered branches would provide crystallinity, while the
branch points are considered to be in the amorphous region. Highly ordered
crystalline regions are more resistant to enzymatic and chemical action and water
penetration than are less organized bulky amorphous regions. Amylopectin is
extensively branched thus restricting the degree of hydrolysis by -amylase, so the amylolysis limit is around 55-60%. This is significantly lower compared with the
linear chain amylose polymer. Because of the short length of unit chains, only small
amounts of iodine can bind with amylopectin (<0.6%), forming a red-brown color.

1.2.4. Minor components of starch


The most abundant components of starch are amylose and amylopectin, which
constitute nearly 100% starch of dry matter. Apart from these main components,
smaller amounts of other components such as proteins, free fatty acids, other lipids
and phosphate groups may also be present in amounts depending on the botanical
source and starch isolation procedure (Morrison and Laignelet, 1983; Morrison and
Karkalas, 1990) and they impart dramatic effects on the physicochemical properties
of starch.

1.2.4.1. Starch lipids


Cereal starches are classified as higher lipid content starches while tuber and root
starches are virtually lipid free (Table 1). In general, cereal starches contain 1-2%
lipids (Tester, 1997b) but the content is lower in waxy varieties and high amylose
starches (Morrison, 1995). Cereal starch lipids have been categorized into three
groups by Morrison (1981) as non-starch lipids, granule surface lipids, and internal
lipids. The surface lipids are mainly triglycerides, followed by free fatty acids,

13

glycolipids and phospholipids. The internal lipids of cereal starches are mainly
monoacyl lipids, with the major components being lysophospholipids and free fatty
acids (Hargin and Morrison, 1980; Morrison, 1981). At least 90% of the lipids from
wheat starch are lysophospholipids (LPL) (Meredith et al., 1978) which may complex
with helical segment of amylose. Lysophosphatidylcholine (LPC) is the main
component (70-90%) of total lipids (Morrison et al., 1993b). Although many solvents
may be used to extract starch lipids it is difficult to extract all the lipids until starch
granules are denatured (Morrison, 1981). It is known that lipid complexed with
amylose is difficult to leach out (Hoover and Hadziyev, 1981). Starch lipids have
many implications on properties of starch e.g., forming complexes with amylose
prevents part of the amylose from contributing thickening power to gelatinized starch,
and can form undesirable flavors by oxidation of unsaturated lipids and reduce
granular swelling (Swinkles, 1985).

1.2.4.2. Phosphate
Several investigations have shown that phosphorus in tuber and root starches is
primarily in the form of starch phosphate monoesters (Lim et al., 1994), whereas in
normal cereal and legume starches it is in the form of phospholipids (Lim et al., 1994;
Maningat and Juliano, 1980; Hizukuri et al., 1983). Jane et al. (1996) reported that
native potato starch contains phosphate monoesters that are mainly located in the
amylopectin and that many of the desirable qualities of potato starch such as
enhanced paste clarity, high peak consistency, and significant shear thinning and slow
rate extent of retrogradation are related to its phosphate content. According to Takeda
and Hizukuri (1982) potato amylopectin contains one phosphate monoester group per
317 glucose residues. The distribution of phosphate monoesters on the C2, C3 and C6
positions of the glucose units of potato amylopectin has been reported to be 1, 38, and

14

61%, respectively (Hizukuri et al., 1970; Tabata and Hizukuri, 1971). The phosphate
groups in potato starch are mainly located on the longer unit chains (30-100 glucose
units) (Takeda and Hizukuri, 1982). Generally, root and tuber starch contain
significant amounts of covalently bound mono phosphate esters (Lim et al., 1994;
Kasemsuwan and Jane, 1994). Phosphate content of various starches is shown (Table
1).

1.2.4.3. Protein
Nitrogen present in the starch is generally considered as protein, but it may also be
part of the starch lipids (Lineback and Rasper, 1988). The protein content of purified
starch is a good indicator of its purity. Alkali extraction is very effective in
solubilizing protein, therefore, careful washing of crude starch with diluted alkali can
reduce protein level in purified starch. In wheat starch, the protein content has been
estimated to be 0.1-0.25% (Eliasson and Larson, 1993). Approximately 10% of starch
protein appears to be associated with the granule surface (Galliard and Bowler, 1987).

1.2.5. Starch structure

1.2.5.1. Amorphous region of starch granule


The amorphous region accounts for 70% of the starch granule (Oostergetel and Van
Bruggen, 1993), and consists of free amylose, lipid-complexed amylose, and some
branched points of amylopectin (Hizukuri, 1996). The conformation of chains in the
amorphous domain appears to be mainly a single helix or random coil (Gidley and
Bociek, 1985, 1988). The amorphous region has shown to be very susceptible to
chemical and enzymatic modification (Hood and Mercier, 1978; Robyt, 1984).
Diffusion of small water soluble molecules (< 1000 Dalton) in the granule also occurs

15

through the amorphous phase. Biliaderis (1998) has postulated that there is no sharp
demarcation between crystalline and amorphous domains in granular starch. Instead,
a range of structures is expected between well-developed crystallites and fully
disordered regions. In this type of super molecular organization, the amorphous and
crystalline phases are interdependent.

1.2.5.2. Semi crystalline structure of starch granule and crystallinity


The molecular order of the starch granule (the arrangement of amylose and
amylopectin) is directly related to the physicochemical properties of native starches.
However it is still under investigation. Hizukuri (1996) reported that native starches
are semi-crystalline in nature with a degree of crystallinity of 20-40%. Crystalline
regions are made up of amylopectin double helices, which are packed in parallel
fashion. The crystalline lamellae exist in the granule alternately with amorphous
lamellae (Fig. 2). The combined thickness of crystalline lamellae plus amorphous
lamellae is 9 nm and 9.2 nm for A-type starches and B-type starches respectively
(Jenkins et al., 1993; Jane, 1997). Yamaguchi et al. (1979) showed that clusters of
amylopectin short chains occur within the crystalline domains of the granule. X-ray
diffractometry has been used to characterize the crystalline structure of starch
granules (Zobel, 1988a; Hizukuri et al., 1983; Cheetham and Tao, 1998). Jenkins et
al. (1993) ascribed the crystallinity of the granule mainly to double helices formed by
amylopectin branches rather than to amylose. Several authors reported that
crystallinity is due to the amylopectin component (Hizukuri, 1996; Blanshard, 1987;
Biliaderis, 1998). Mainly A chains and outer B chains of amylopectin are involved
in the construction of the crystalline region (Hizukuri, 1996).

16

Based on the packing arrangement of the starch crystallites, starch can be classified
into three groups as A-, B-, and C- type. Cereal starches show mainly A type
packing arrangement while most tuber and root starches as well as high amylose
starches exhibit the typical B type X-ray pattern (Zobel, 1998a) [Table 3]. The C
pattern is believed the superposition of the A and B patterns. It has been reported
that starches with shorter amylopectin chain length (< 20 residues) exhibit A-type
crystallinity while amylopectins with longer average chain length show the B
pattern (Hizukuri, 1986; Hizukuri et al., 1983). The amylopectin of A type starches
has a closer packing arrangement compared with that of B type starches. The unit
cell of amylopectin is estimated to hold 8 water molecules for A-type and 36 water
molecules for B-type (Imberty et al., 1991; Zobel, 1988a). The C-polymorph is a
mixture of A and B unit cells, and is thus intermediate between A and B types in
packing density. Chain length (CL) of amylopectin is the main factor that determines
the crystalline pattern of the starch granule (A-type CL <19.7; B-type CL 21.6),
and starches having CL between 20.3 and 21.3 exhibit A, B or C-type patterns
(Hizukuri et al., 1983). Gunaratne and Hoover (2002) suggested that B-type starches
are more responsive to heat-moisture treatment than A-type starches due to the
differences in characteristics of the unit cell structure. Less compactly packed helices
in B-type starches would be more mobile and hence more prone to disruption during
heat-moisture treatment. Jane et al. (1997) have shown that in A-type starches, the
branch -(16) linkages are located within the crystalline region and amorphous area,
whereas in B-type starches, the branches are located mainly within the amorphous
area. Thus B-type starches are more susceptible to acid hydrolysis.

1.2.6. Functional properties of starch

17

Functionality is the key for the industrial utilization of starch. Functional properties
can be defined as the characteristics that govern the behavior of a food component
during processing, storage and preparation. Factors such as botanical source, granule
architecture and structural characteristics of the granule and pretreatments of starch
directly affect functional properties of starch.

1.2.6.1. Swelling and solubility


Native starch granules are insoluble in water. Although small amounts of water can
be absorbed at room temperature, granular swelling is limited in intact granules.
French (1984) reported that when dry starch granules are immersed in water the
resulting considerable but limited swelling is presumably related to the swelling of
the amorphous or gel phase. If the temperature is increased the amount of absorbed
water increases and the swelling process is reversible up to the gelatinization onset
temperature. Swelling power and solubility provide evidence of the magnitude of
starch chains interaction within the amorphous and crystalline domain of the starch
granule. The extent of this interaction is influenced by the amylose/amylopectin ratio
and by the characteristics of amylose and amylopectin. Swelling begins in the least
organized, amorphous, intercrystallite regions of the granule. French (1984) reported
that swelling is due to the hydration and lateral expansion of amylopectin crystallite.
By comparative studies on non-waxy and waxy maize starches, Tester and Morrison
(1990a) have shown swelling is primarily a property of amylopectin.

Many factors have been shown to influence the swelling properties of starch granules.
It is widely believed that lipid complexed with amylose restricts granule swelling
(Tester and Qi, 2004; Galliard and Bowler, 1987). Among the other factors that have
been shown to influence granular swelling are granular size (Vasanthan and Bhatty,

18

1996), crystallinity (Robin et al., 1975), temperature (Colonna and Mercier, 1985),
amount of phosphate linked to amylopectin (Jane et al., 1996 ), amylose content
(Tester and Morrison, 1990a), protein (Han and Hamaker, 2002; Han et al., 2002).
Several methods are currently available to measure the extent of granular swelling
when starch is heated in excess water such as swelling volume, swelling power, and
swelling factor. Swelling factor measures only the intergranular water content and
hence reflects the true swelling of starch granule (Tester and Morrison, 1990a).
Swelling volume is a modified version in which programmed shaking of starch
suspension is used instead of stirring, and measurement is of the volume expanded
due to granular swelling (Crosbie, 1991). Swelling factor of some cereal, tuber and
root starches is shown (Table 1).

1.2.6.2. Gelatinization
Gelatinization is one of the most important processes affecting starch. In the majority
of food applications, starch functions (e.g., imparts texture for consumption) in a
gelatinized form. Melting and gelatinization are phase transitions that are associated
with the transformation of the crystalline regions of partially crystalline starch to
amorphous liquid. The above phase transition is associated with the diffusion of water
into the granule, water uptake by the amorphous background region, hydration and
radial swelling of the starch granules, loss of optical birefringence, uptake of heat,
loss of crystalline order, uncoiling and dissociation of double helices in crystalline
regions and amylose leaching (Donovan, 1979; Biliaderis, et al., 1980; Evans and
Haismann, 1982; Hoover and Hadziyev, 1981). According to the well-known
Donovan (1979) theory, gelatinization in high water level is a swelling driven process
where the stress build up on crystalline region due to the expansion of amorphous
region disrupts the crystalline region. Recently, Waigh et al. (2000a) have proposed a

19

Fig. 2. Model for semi crystalline structure of the starch granule


Source: Donald (1997).
A. Stacks of amylopectin lamellae are separated by amorphous growth
ring.
B. A magnified view of one stack, showing that it is made up of
alternating crystalline and amorphous lamellae.
C. Double helices formed from amylopectin branches in crystalline
lamellae and amylopectin branch points are located in the amorphous
lamellae.

20

model for gelatinization based on the side-chain liquid crystalline model for starch.
Presently various analytical methods are available for the determination of starch
gelatinization such as light microscopy (LM), electron microscopy (EM), viscometry,
enzymatic analysis (EA), X-ray crystallography, nuclear magnetic resonance (NMR),
and differential scanning calorimetry (DSC) [Zobel, 1984; Gunaratne and Corke,
2004]. Differential scanning calorimetry has become perhaps the most widely used
method for studying starch gelatinization. The amylopectin melting process can be
observed as an endothermic peak in the DSC analysis. DSC measures the dissociation
parameters To (onset), Tp (peak), Tc (conclusion) and H (endothermic heat
absorption). Noda et al. (1998) reported that gelatinization temperatures are
influenced more by the granular architecture than the amylose-amylopectin ratio. The
gelatinization enthalpy is a measure of the overall crystallinity of the amylopectin
(quality and quantity of starch crystals) [Tester and Morrison 1990b] and crystalline
perfection is reflected in the gelatinization temperatures (Tester 1997b). Cooke and
Gidley (1992) suggested that gelatinization enthalpy primarily reflects the loss of
double helical order. According to Gernat et al., (1993) the amount of double helical
order in the native starches is strongly correlated to the amylopectin content and the
crystallinity in the granules increases with the amylopectin content. Complex
formation of amylose with some of the free lipids in cereal starch granules during
gelatinization could decrease the endothermic gelatinization enthalpy because
amylose-lipid complex formation is an exothermic process (Eliasson, 1986a).
Gelatinization has been shown to be influenced by many factors such as botanical
source of starch (Table 2), water content, added solutes, growth conditions, heating
rate, and thermal history.

1.2.6.2.1. Effect of water content on gelatinization

21

Depending on the water content available during gelatinization, changes in


endothermic transition have been detected in terms of number and the position of the
differential scanning calorimetric endotherms. A single peak occurs when higher
moisture content is available while two transition endotherms that are shifted to
higher temperature appear at low moisture content (Fig. 3). Donovan (1979)
postulated that if sufficient water were present, all the crystallites might be pulled
apart by the swelling leaving non to be melted at higher temperature producing a
single endotherm. If there is insufficient water left, the remaining crystallites melt
later at higher temperature causing second endotherm. According to Kaletunc and
Breslauer (2003) the observation of more than one transition peak in limited moisture
content (~25%-60%) could be attributed to non-uniform moisture distribution or the
presence of microcystallites with different thermal stabilities. The double endotherm
has also been described as resulting from a redistribution of water between crystalline
domain and gelatinized starch (Evans and Haisman, 1982).

1.2.6.2.2. Effect of polyhydroxy compound on starch gelatinization


Polyhydroxy compounds such as sugars are widely used in many food products.
Although sugars provide sweet flavor as the main contribution, they also significantly
change the structural and textural properties of food products by affecting
thermomechanical behavior. The effect of sugars on starch gelatinization has been
studied by several researchers using a wide range of techniques such as light
microscopy (Bean and Yamazaki, 1978), rheological measurements (DAppolonia,
1972), Differential scanning calorimetry (Wootton and Bamunuarachchi, 1980;
Eliasson, 1992; Evans and Haisman, 1982), electron spin resonance (Johnson et al.,
1990), and nuclear magnetic resonance (Lim et al., 1992), and combining differential

22

scanning calorimetry (DSC), electron spin resonance (ESR), and dynamic mechanical
thermal analysis (DMTA) (Chiotelli et al., 2000).

The elevation of starch gelatinization temperature with the inclusion of sugars and
other polyols is well documented. Suggestions proposed for the mechanism of this
elevation are (1) that sugars reduce the water activity in the system due to competition
for water with starch molecules which always favors sugars (DAppolonia, 1972) (2)
that there is a sugar-starch interaction (Spies and Hoseney, 1982; Baek et al., 2004;
Hoover and Senanayaka, 1996; Chiotelli et al., 2000) (3) that there is less plasticizing
effect of polyol-water solvent (Perry and Donald, 2002; Levine and Slade, 1988).
Although several such theories have been proposed, the exact mechanism still
remains unclear. Specially, conflicting results have been reported for gelatinization
enthalpy. Gelatinization enthalpy has been shown to increase (Baek et al., 2004;
Chiotelli et al., 2000; Ahmad and William, 1999; Sopade et al., 2004), decrease
(Wootton and Bamunuarachchi, 1980; Chungcharoen and Lund, 1987), or remained
unchanged (Eliasson, 1992; Maaruf et al., 2001; Evans and Haisman, 1982) in the
presence of sugars. A literature survey reveals that many studies were carried out to
characterize sugar-starch interaction and its influence on starch gelatinization, but
very little information is available for the other poly hydroxyl compound-starch
interactions and their effect on starch gelatinization.

1.2.6.3. Pasting
Starch heated in excess water undergoes various changes as a result of heat and
moisture transfer. Gelatinization and pasting occur in the same system can be used to
describe many of the changes that occur. Gelatinization may be used to refer to early
changes whereas pasting includes later changes. Pasting is defined as the phenomena

23

Table. 2. DSC characteristics of the transition related to the amylose-lipid complex of some cereal starches
Starch
Conditions
Tcx
Hcx
Wheat
50% water
110.1
1.4
Rye
50% water
107.8
0.8
Barley
50% water
110.3
1.8
High-amylose barley
50%water
110.8
2.8
Waxy barley
ND
ND
ND
Source: Eliasson (2003).
Tcx = melting temperature of amylose-lipid complex.
cx = melting enthalpy of amylose-lipid complex.
ND = not detected.
Table. 3. Gelatinization parameters (as measured by DSC), x-ray pattern and crystallinity of some cereal, tuber and root starches
Starch source
Starch;water Tp
H
Reference
X-ray
Crystallinity
Reference
ratio
(%)
( C) (J/g)
pattern
Potato
1:3
66.3
16.3
Gunaratne and Hoover, (2002)
B
30.0
Gunaratne and Hoover, (2002)
Sweet potato
1:3
75.2
11.2
Collado and Corke, (1999)
A,C,Ca
Zobel, (1998a), Morthy,(1994)
Cassava
1:3
71.5
12.3
Gunaratne and Hoover, (2002)
A
37.0
Gunaratne and Hoover, (2002)
True yam (Dioscorea 1:3
80.0
17.8
Gunaratne and Hoover, (2002)
B
32.0
Gunaratne and Hoover, (2002)
alata)
New coco yam
1:3
77.2
13.1
Gunaratne and Hoover, (2002)
A
45.0
Gunaratne and Hoover, (20020
Taro
1:3
83.0
14.5
Gunaratne and Hoover, (2002)
A
31.0
Gunaratne and Hoover, (2002)
Wheat
1:3
59.9
10.7
Tester et al., (1998)
A
36.0
Zobel, (1998b)
Regular maize
1:3
71.3
11.0
Liu et al., (1999)
A
40
Zobel, (1998b)
Waxy maize
1:3
72.8
13.6
Liu et al., (1999)
High-amylose maize
1:3
92.3
13.7
Liu et al., (1999)
B
15-22
Zobel, (1998b)
Oat
1:3
62.9
10.2
Jayakody and Hoover, (2001)
A
33.0
Zobel, (1998b)
Rice
1:3
78.0
13.0
Jayakody and Hoover, (2002)
Barley
1:3
Tp = gelatinization temperature; H = gelatinization enthalpy.

24

Fig. 3. DSC curve for wheat starch gelatinization at low water content
Source: Eliasson (2003).
G = low temperature gelatinization endotherm.
M1 = higher temperature endotherm.
M2 = amylose-lipid transition endotherm.
Tm = gelatinization endotherm.
Tcx = amylose-lipid complex melting temperature.

25

following gelatinization in the dissolution of starch, involving granular swelling,


exudation of molecular components from the granule, and, eventually total disruption
of the granules (Atwell et al., 1988).

Properties of starch paste are highly correlated with the textural characteristics and
storage ability of many food products. These properties can be measured using
various instruments such as Brookfield viscometer, Brabender Viscoamylogram (BV)
and more recently Rapid Visco-Analyzer (RVA). Because of some technical
shortcomings with the BV, such as large sample size requirement and inability to
program the temperature profiles, the more recent equipment, the RVA, became more
popular for analyzing pasting properties. The RVA differs from the BV due to rapid
heating rate and stronger mixing action. However, controlled heating rate at 1.5 C
min-1 in RVA provides similar results to those observed in the BV (Deffenbaugh and
Walker, 1989). RVA measures changes in viscosity of a system over time as the
system is stirred during programmed heating and cooling, according to a pre-set
temperature and shear profile. In a typical pasting curve the temperature at which
viscosity begins to increase is termed pasting temperature. With further swelling rapid
increase in viscosity is due to granule swelling and starch leaching, with a peak in
viscosity occurring above the starch gelatinization temperature (Whistler and
BeMiller, 1997). Further heating at elevated temperature under shear force tends to
reduce the viscosity due to disintegration of swollen granules and orientation of
soluble starch molecules in the direction that the system is being stirred (Whistler and
BeMiller, 1997; Hoseney, 1994). In the final step (cooling step) decrease of energy in
the system and subsequent hydrogen bond formation between starch chains mainly in
amylose increases the viscosity (setback) (Hoseney, 1994). Therefore a typical
pasting profile exhibits three distinct viscosity developments, peak viscosity, hot

26

paste viscosity, and cold paste viscosity. Pasting curves vary according to botanical
source of starch, starch concentration, and programmed heating-cooling cycle chosen.
Based on the pattern of the pasting curve starch can be classified into four groups
(Schoch and Maywald, 1968). According to this classification, high swelling waxy
cereals, tuber and root starches exhibit a high peak followed by rapid and major
thinning during cooking, whereas moderate swelling normal cereal starches exhibit
low peak and much less thinning. Cross-bonded restricted swelling starches show no
clear peak but remain at high viscosity during cooking. Starches with amylose content
> 55% or highly restricted swelling starch have little swelling and do not give a
viscous paste at normal concentrations.

RVA and BV relate to the investigation of the ultimate or large deformation


properties of the starch paste. Equally important small deformation technique,
particularly the dynamic oscillatory rheometry can also be used to study the
rheological properties of starch paste (Miles et al., 1985; Ring et al., 1987). Dynamic
oscillatory rheometry allows continuous assessment of dynamic moduli without
breaking structural elements in the starch gel. This technique offers complementary
evidence of molecular processes occurring in the intact starch network following
gelatinization.

1.2.6.3.1. Effect of polyhydroxy compounds on starch pasting properties


Several studies have shown that sugars and sugar alcohols delay the viscosity
development and increase the maximum peak viscosity depending on the type of
sugars and sugar alcohols (Savage and Osman, 1978, DAppolonia, 1972; Bean and
Yamzaki 1978; Bean and Osman, 1958). The inhibition of starch granule hydration
by sugar molecules is indicated by the higher pasting temperature (DAppolonia,

27

1972). However, in a study of wheat starch pasting in the presence of sucrose,


Richardson et al. (2003) found early viscosity increase. The greater the sucrose
concentration, the more the early viscosity increased. According to Richardson et al.
(2003) early viscosity increase in the presence of sucrose is due to its influence on
close packing of the granules. Further they reported that greater peak viscosity
increase by sucrose is related to its influence on the rigidity of the starch granules
since the granules were seen to be less fragmented in the presence of sugar. Pasting
characteristics are generally influenced by the extent of swelling, leaching of soluble
carbohydrate and rupture of starch granules. In dilute starch solution about less than
70% (v/v) starch viscosity is primarily governed by the volume fraction of starch
(Steeneken, 1989; Doublier et al., 1987). As the paste becomes more concentrated
viscosity has shown to be governed by the rigidity of the granules (Steeneken, 1989;
Doublier et al., 1987). Doublier et al. (1987) have reported the viscosity remains very
low until close packing of the granules is attained. For more precise information on
pasting characteristics in the presence of polyhydroxy compounds, studies should
focus on different types of polyhydroxy compound starch interaction such as
cyclodextrin-starch interactions.

1.2.6.4. Retrogradation
Reassociation of starch polymers via hydrogen bonding in gelatinized starch on
cooling is generally termed retrogradation, and is time and temperature dependent.
This association of starch molecules in starch gel results in precipitation, gelation, and
changes in consistency followed by the gradual formation of crystallites resulting in
phase separation between polymer and solvent. Retrogradation is of great interest for
food scientists since it has a profound effect on quality, shelf-life, acceptability and
nutritional value of starch-based food products (Biliaderis, 1991). Starch gels tend to

28

undergo structural changes during storage as they are metastable and non-equilibrium
systems (Ferrero et al., 1994).

Both amylose and amylopectin are involved in retrogradation, where rapid amylose
aggregation causes a short term development of starch gel providing initial firmness.
Branched amylopectin recrystallization, particularly outer branches of the
amylopectin molecule, is correlated with long-term development of starch gel (Miles
et al., 1985). During retrogradation, amylose forms double-helical associations (Fig. 4)
of 40-70 glucose units (Jane and Robyt, 1984; Leloup et al., 1992), whereas
amylopectin crystallization occurs by association of the outermost short branches
(DP=15) [Ring et al., 1987]. The ordered retrograded starch crystallites show B-type
X-ray diffraction pattern (Zobel, 1988a) containing both crystalline and amorphous
region.

Many factors have shown to influence starch retrogradation: 1) starch concentration


(Orford et al., 1987; Gudmundsson and Eliasson, 1990; Biliaderis and Tonogai, 1991;
Liu and Thompson, 1998), 2) storage temperature (Slade and Levine, 1987), 3) initial
heating temperature (Liu and Thompson, 1998), chain length distribution of
amylopectin (Yuan et al., 1993; Shi and Seib, 1995; Liu and Thompson, 1998), 4)
molecular size of amylose (Lu et al., 1997), 5) salts (Ward et al., 1994), 6) lipids
(Biliaderis and Tonogai, 1991; Ward et al., 1994), 7) sugars (Biliaderis and
Prokopowich, 1994), 8) physical modification (Hoover et al., 1994; Gunaratne and
Hoover, 2002), 9) chemical modification (Hoover et al., 1988; Wu and Seib, 1990;
Perera and Hoover, 1998), 10) starch source (Orford et al., 1987; Hoover, 1995).
Several methods have been developed to determine starch retrogradation, because of

29

its great influence on food and non-food products, such as DSC, X-ray analysis,
rheological methods, and spectroscopic methods (Gunaratne and Corke, 2004).

1.2.6.4.1. Effect of retrogradation on quality of starch-based food products


The influence of retrogradation can be desirable or undesirable in starch-related food
products. Mostly, it is undesirable such as bread staling or undesirable firming of
bread and other starch-based products (Kulp and Ponte, 1981; Knightly, 1977; Seow
and Thevamalar, 1988) and greater susceptibility of legume starch gel to syneresis or
phase separation which makes these types of starches unsuitable for products
requiring low temperature storage. Although several factors contribute to bread
staling, it has been found that retrogradation is the key physical change associated
with bread staling (Gunaratne and Corke, 2004). However, retrogradation sometimes
aids processing in some food products such as hardening of parboiled rice and to
improve textural characteristics of certain types of noodles and cooked potato
products (Colonna et al., 1992; Ooraikul, et al., 1974; Watanabe, 1981).
1.2.6.4.2. Measurement of amylopectin retrogradation using differential
scanning calorimetry (DSC)
Several studies on starch retrogradation have been examined using DSC techniques
(Silverio et al., 2000; Eliasson, 1985b; Zeleznak and Hoseney, 1987; Obanni and
BeMiller, 1997; Ortega-Ojeda and Eliasson, 2001; Gunaratne and Hoover, 2002).
DSC is probably the best thermal analysis method to study the starch-aging process
for different systems such as bread staling (Eliasson, 2003). In contrast to faster
crystallization of amylose, the crystallization of amylopectin is much slower and
continues to occur over a period of days or weeks. Highly branched amylopectin has
limited and weaker chain interaction and thus produces less stable crystals compared
with amylose crystals. While recrystallized amylopectin melts in the temperature
30

range 40-100 C, amylose crystallites do so only at much higher temperatures


(Eerlingen et al., 1994). Immediate cooling and reheating of gelatinized wheat starch
sample produces only the transition endotherm of amylose-lipid complex (Eliasson et
al., 1988), but if the sample is stored for at least a day another transition endotherm
can be seen. This is due to the melting of recrystallized amylopectin which shows
about the same temperature interval as the gelatinization endotherm (Eliasson, 1985b).
The extent of amylopectin retrogradation can be detected by measuring the melting
enthalpy (Eliasson, 1985b) [Table 4]. All retrograded amylopectin crystals of most of
starches melt when heated above 80 C.

Many factors have been shown to influence amylopectin retrogradation such as


storing condition (temperature and time), starch:water ratio, and other ingredients
(enzymes, emulsifiers). Maximum amylopectin retrogradation has been reported to
occur at intermediate water content (40-50%) [Zeleznak and Hoseney, 1987]. By
cycling the storage temperature between low (6 C), to facilitate the nucleation and a
higher temperature (30 or 40 C) to facilitate the growth of crystals, the quality of
retrograded amylopectin crystals can be improved (Eliasson, 2003). This kind of
crystals has greater homogeneity and thus melts over a narrow range of temperature.

1.2.6.5. Properties of starch gel


Starch gel can readily be formed when gelatinized starch dispersions are subject to
cooling. Starch gel is a solid-liquid system having a solid, continuous network in
which the liquid phase is entrapped. Amylose chains that are leached out from the
swollen granules during the gelatinization can form three dimensional gel matrix
(Eliasson, 1985a) surrounding the swollen granules on cooling. Ring (1985) proposed
that starch gel can be regarded as the composite in which swollen gelatinized starch

31

granules reinforce an interpenetrating amylose gel matrix (Fig. 4). Morris (1990)
explained that gel properties relate to the properties of the gel matrix (amylose), the
deformable filler (swollen granule), the volume fraction of filler, and the filler-matrix
interaction. Doublier et al. (1987) suggested that the main structural parameters
involved in the starch gel are the deformability of the swollen particles and the
amylose concentration of the continuous network. The mechanical properties of
starch gel depend on the rheological characteristics of the amylose matrix, the volume
fraction and the rigidity (deformability) of the amylopectin-rich gelatinized granules,
and the interactions between the dispersed and the continuous phase (Eliasson,
1986b).

Many factors have been shown to influence the structure and properties of starch gel
(rheological properties) such as chain length and concentration of amylose, presence
of other compounds (sugars, salts, lipids), starch concentration, pasting procedures,
morphology of starch granules and granule size distribution, amylose-amylopectin
ratio and addition of amylose (Ortega-Ojeda and Eliasson, 2004; Yoo and Yoo, 2005;
Evageliou et al., 2000; Doublier, et al., 1987; Kokini et al., 1992; Gidley, 1990;
Doublier, 1990; Biliaderis, 1992; Launay et al., 1986).

1.2.7. Hydrolysis by -amylase


Alpha-amylases are endoenzymes (endo attack), which catalyze the hydrolysis of (14) linkages in amylose and amylopectin randomly releasing oligosaccharides.
Two mechanisms were proposed for the action of -amylase on amylose solution as
multiple attack and multichain attack (Banks and Greenwood, 1977; Mazur, 1984).
The multiple attacks lead to the formation of small saccharides during early stages of
amylose hydrolysis whereas multichain attack does not. In multiple attack mechanism,

32

Table. 4. The amylopectin retrogradation enthalpy (H on amylopectin basis)


obtained for different starches after two days of storage at 6 C followed by 30 C
Starch
Starch concentration
H (J/g AP)
Wheat
50%
7.7
Rye
50%
6.4
Barley
50%
8.2
High-amylose barley
50%
11.1
Waxy barley
50%
10.0
Waxy maize
50%
12.2
Potato (Desiree)
50%
11.7
Potato (Prevalant)
50%
13.2
High-amylopectin potato
50%
13.5
Pea
50%
11.8
Source: Silverio et al. (2000).
Table. 5. In vitro amylolysis of some cereal, tuber and root starches
Starch source
-amylase source
Reaction Hydrolysis Reference
time
(%)
Potato
Porcine pancreatic
72
5.4
Hoover and Vasanthan, (1994)
Pancreatic
8.5
Fuwa et al., (1997)
5.0
Fuwa et al., (1997)
Bacillus subtilis
Sweet potato
55
14.9
Fuwa et al., (1997)
Bacillus subtilis
Pancreatic
43.3
Fuwa et al., (1997)
Porcine pancreatic
Cassava
24
44
Valetudie et al., (1993)
Bacillus subtilis
Porcine pancreatic
52.9
Valetudie et al., (1993)
True yam
24
3.5
Valetudie et al., (1993)
Bacillus subtilis
Porcine pancreatic
24
3.1
Gunaratne and Hoover, (2002)
Porcine pancreatic
72
4.9
Gunaratne and Hoover, (2002)
Taro
Porcine pancreatic
24
22.2
Gunaratne and Hoover, (2002)
Oat
Porcine pancreatic
72
32
Hoover and Vasanthan, (1994)
Wheat
Porcine pancreatic
72
63
Hoover and Vasanthan, (1994)
Waxy maize
Bacterial -amylase
14
98.9
Liu et al., (1999)
High amylose- Bacterial -amylase
14
40.1
Liu et al., (1999)
maize

33

once the enzyme forms a complex with the substrate and forms the first cleavage, the
enzyme remains with one of the fragments of the original substrate and catalyzes the
hydrolysis of several bonds before it dissociates and forms a new active complex with
another molecular substrate (Robyt, 1984). The direction of the multiple attack by
porcine pancreatic -amylase is from the reducing towards the non-reducing end
(Robyt, 1984). In the multichain attack mechanism, the encounter between enzyme
and the substrate leads to a single hydrolysis, both molecules being released after the
catalytic event. Differences in the in vitro -amylase digestibility of native starches
among and within species have been attributed to the interplay of many factors such
as starch source (Ring et al., 1988), granule size (Vasanthan and Bhatty, 1996),
amylose/amylopectin ratio, (Hoover and Sosulski, 1985a), type of crytallinty
(Planchot et al., 1997), location of amylopectin branch points and the nature of
crystalline structure (Jane et al., 1997) and nature of the surface of the granule (Fuwa
et al., 1997). It has been reported that the hydrolysis by -amylase predominantly
occurs in the amorphous regions of the granules (Franco et al., 1987). The nature and
the extent of enzyme hydrolysis were mostly studied by the conventional in vitro
method providing standard temperature and pH. However, in the majority of starch
applications, starch is subject to higher temperature, shear force, heating/cooling
process, and different pH conditions. Thus it is worth employing a method that can
detect the extent of enzyme action on starch under those conditions. In this respect,
viscoamylography can successfully be employed to detect the action of enzyme.
Alpha-amylase hydrolysis of various starch sources is shown (Table 5).

1.2.8. Starch modification


Modification of starch is required when functional properties are not desirable or
controllable in native starch. Modified starches can be defined as products, whose

34

Fig. 4. (A) Schematic representation of starch gel. Swollen granules fill the
amylose gel matrix. Source: Morris (1990).
(B) Conformational changes occurring during amylose gelation
Source: Colonna et al. (1992).

35

properties have been altered by physical or chemical means or by the introduction of


substituents and whose granular and molecular structure respectively are more or less
retained (Tegge, 1979).

1.2.8.1. Physical modification


Heat-moisture treatment, annealing and pregelatinization are common physical
modification techniques. Since physical modification involves only heat and moisture
treatments, there is a growing interest in these techniques, especially for food
applications due to its higher safety compared to chemical modification.

1.2.8.1.1. Heat-moisture treatment


Heat-moisture treatment of starch is defined as a physical modification that involves
incubation of starch granules at low moisture level (< 35% water [w/w]) during a
certain period of time, at a temperature above the glass transition temperature but
below the gelatinization temperature. Influence of heat-moisture treatment on the
starch structure and functional properties has been extensively studied. Granule
morphology of maize, wheat, potato, yam and lentil starches has been shown to
remain unchanged after HMT (Kulp and Lorenz, 1981; Stute, 1992; Hoover and
Vasanthan, 1994; Franco et al., 1995). Heat-moisture treatment has shown to change
the wide angle X-ray pattern from B to A or A+B type (Donovan et al., 1983; Kuge
and Kitamura, 1985; Hoover and Vasanthan, 1994; Gunaratne and Hoover, 2002;
Kawabata et al., 1994). Several theories have been suggested to explain changes in Xray pattern during heat-moisture treatment such as destruction of crystallites (Hoover
and Vasanthan, 1994), growth of new crystallites (Hoover and Vasanthan, 1994),
reorientation of the already existed crystallites and changes in the packing
arrangement (Lorenz and Kulp, 1982; Hoover and Vasanthan, 1994). Hoover and

36

Vasanthan (1994), and Lorenz and Kulp (1982) observed a decreased apparent
amylose content in heat-moisture treated wheat and potato starch, indicating
additional interaction between native starch lipids and amylose chains. However,
Donovan (1983) found unchanged DSC amylose-lipid dissociation endotherm for
heat-moisture treated wheat starch.

Gelatinization transition temperatures and breadth of the transition endotherm have


been shown to increase after heat-moisture treatment (Kulp and Lorenz, 1981;
Donovan et al., 1983; Radosta et al., 1992; Hoover and Vasanthan, 1994; Gunaratne
and Hoover, 2002). The gelatinization enthalpy was decreased (Donovan et al., 1983;
Kuge and Kitamura, 1985; Radosta et al., 1992; Stute, 1992; Hoover and Vasanthan,
1994; Hoover et al., 1994; Eerlingen et al., 1996) or unchanged after heat-moisture
treatment. Hoover and Vasanthan (1994) suggested that starch chain interaction
within the amorphous region during heat-moisture treatment decreased its
destabilization effect on crystalline region during gelatinization and thereby increases
the gelatinization temperature.

Swelling was generally found to decrease after heat-moisture treatment due to the
interplay of following reasons; changes in the packing arrangement of the starch
crystallites, interaction between or among starch chains in the amorphous regions of
the granules and amylose-lipid interaction. The later two factors could decrease the
amylose leaching (Kulp and Lorenz, 1981; Hoover and Vasanthan, 1994; Hoover and
Manuel, 1996).

Generally, work on pasting properties of heat-moisture treated potato starch have


shown higher onset temperature of viscosity development, a lower peak viscosity, and

37

a higher or lower end viscosity (Kulp and Lorenz, 1981; Hoover and Vasanthan 1994;
Stute, 1992). Similar observations were reported for maize (Hoover and Manuel,
1996), wheat, lentil, oat, and yam (Hoover and Vasanthan, 1994).

Perera et al. (1997) studied the influence of heat-moisture treatment on


hydroxypropylation of potato starch. The results indicated that heat-moisture
treatment increased the degree of substitution. Although extensive studies have been
performed on impact of heat-moisture treatment on the starch structure and functional
properties, little effort has been made to find its usefulness as a prior treatment for
chemical modification aiming to produce novel starch properties.

1.2.8.2. Chemical modification


Chemical modification of starch is made possible because of the many hydroxyl
groups available for reaction with the chemical used and is the most widely used
modification technique in both food and non-food applications.

1.2.8.2.1. Acetylation
Starch acetates are derivatives of starch obtained through esterification. In acetylation,
hydrophilic hydroxyl groups are substituted with hydrophobic acetyl groups. Usually,
acetic anhydride in aqueous medium is used in the presence of alkaline medium such
as sodium hydroxide to make starch with low degree substituent. This is an example
of nucleophilic substitution (Fig. 5A) at an unsaturated carbon atom of acetic
anhydride (Xu et al., 2004). In the acetylation reaction, introduction of bulky acetyl
groups to starch molecules (amylose and amylopectin) reduces the bond strength
between the starch molecules interrupting the ordered structure of starch and
interfering with the association of amylose and amylopectin in gelatinized starch

38

leading to decreased gelatinization temperature, increased swelling, solubility, clarity


and improved storage stability (Jarowenk, 1986; Rutenberg and Solarek, 1984).
Biliaderis (1982) reported that acetylation occurred exclusively in certain parts of the
starch granule. After analyzing X-ray patterns of native and acetylated legume starch,
Hoover and Sosulski (1985b) suggested that acetyl groups mainly enter the
amorphous region. Acetylation can prevent the association of amylopectin branches
to a greater extent (Bentacur et al., 1997). A substantial amount of work carried out
on different botanical sources of starches have shown that acetylation increased
swelling, solubility, peak viscosity, cold storage stability, clarity of starch paste, and
enzymatic digestibility, but decreased gelatinization temperature and enthalpy (Liu et
al., 1997 and 1999; Hoover and Sosulski, 1985b; Lawal and Adebowale, 2005; Perez
et al., 2000; Hung and Morita, 2005; Craig et al., 1989; Reddy and Seib, 2000).

Although much research on acetylation has been done, there is very little literature on
the structural and functional properties of dually modified starches where one is
acetylation. In a study of the effect of autoclaving on the acetylation of mutant maize
starch Liu et al. (1997, 1999) reported that prior autoclaving decreased the degree of
acetylation. Vasanthan et al. (1995) found that autoclaving of starches before
acetylation could increase the reactivity so that it needed less concentration of the
modifying reagent to achieve a desired degree of chemical bonding. These findings
suggest that some prior treatments before acetylation assist to minimize the
concentration of modifying reagent while achieving similar desirable outcome.

1.2.8.2.2. Hydroxypropylation
Low substituted hydroxypropyl starches are widely used in the food industry. In this
substitution reaction, the etherifying reagent, propylene oxide reacts with starch under

39

alkaline conditions resulting in the introduction of hydroxpropyl group onto the


polymeric chains of starch (Fig. 5B). Alkaline regents (NaOH, KOH) are used in
these substitution reactions to activate the starch molecule to make O-H bond
nucleophilic and to facilitate the formation of O -. Gray and BeMiller (2005)
reported that, in the hydroxypropylation reaction, reaction reagents enter the less
ordered (more amorphous) areas of starch granules most readily, swelling and
reacting there first, and that the derivatization in these regions causes them to swell
opening the highly ordered crystalline region and thereby facilitates the access of
reaction reagent to the crystalline region. This implies that granule derivatization is a
cooperative and concerted process that starts in the most accessible amorphous
regions and proceeds through various regions until it reaches the highly ordered
crystalline region (Han and BeMiller, 2005).

Several studies have investigated the location of reaction site of the starch granule in
the derivatization process. Biliaderis (1982) showed that hydroxypropylation occurs
more uniformly throughout the granule. Hood and Mercier (1978) have shown that
hydroxypropyl groups in amylose were distributed at/or near the reducing end or
along the entire amylose molecule. For amylopectin, the substituents were largely
distributed in the amorphous region, rich in -1,6 linkages. It has been reported that
amylose is derivatized to a greater extent than amylopectin (Shi and BeMiller, 2002;
Kavitha and BeMiller, 1998) and that derivatization increased the amylose leaching.
However, Perera et al. (1997) have observed decreased amylose leaching for
hydroxypropylated potato starch. By light microscopy, Kim et al. (1992) have shown
that the central region of the potato starch granule is where hydroxypropylation
mainly takes place or where hydroxypropyl groups are mainly distributed.
Controlling the location of reaction site within the granule is one possible avenue for

40

making novel starch properties (BeMiller, 1997). In this respect, prior treatments such
as heat-moisture treatment may be applied to change the location of reaction site
because structural changes of the starch granule taking place during heat-moisture
treatment could alter the reaction site by affecting the accessibility to reaction
reagents. Hydroxypropylation increases the hydrophilic nature in starch granule.

Introduction of hydroxypropyl groups onto the starch polymers alters the starch
structure by weakening the associative bonding forces between adjacent starch chains
leading to formation of a loosened starch structure (Liu et al., 1999). The extent of
these changes depends on the level of molar substitution (Perera et al., 1997).
Loosened starch structure due to the inhibition of interchain association facilitates the
hydration of the starch granule resulting in increased swelling, solubility, peak
viscosity, clarity of starch paste, and susceptibility to enzymatic hydrolysis, but
decreased gelatinization temperature and retrogradation (Liu et al., 1999; Perera et al.,
1997). The penetration of water into the starch granule could increase the tendency of
gelatinization by increasing the initial rate of plasticization of the amorphous region
(Seow and Thevamalar, 1993). Dias et al. (1997) reported that main effect of
etherification is to inhibit retrogradation. Various physicochemical properties of
hydroxypropylated starches from different sources have been studied (Liu et al., 1999;
Perera et al., 1997; Hoover et al., 1988; Butler et al., 1986; Hung and Morita, 2005;
Pal et al., 2002; Wang and Wang, 2000; Choi and Kerr, 2004; Islam and Azemi,
1997). However, little work has been reported on the effect of dual modification such
as cross-linking of hydroxypropylated starches for different starch sources. In dual
modification, usually, cross-linking has to be done before hydroxypropylation. Wang
and Wang (2000) observed the effect of modification sequence on chemical structure
and physicochemical properties of hydroxypropylated and cross-linked waxy maize

41

starch. The results showed that the modification sequence alters the location of
substitution and susceptibility to enzymes. Primary use of alkaline treatment (NaOH
and Na2SO4) is to catalyze the derivatization reaction while providing initial swelling
for the starch granule (Gray and BeMiller, 2005). Hauber et al. (1992) concluded that,
in order for reactions with granular starch to take place, granule swelling must first
occur. Thus the importance of granule swelling to the reactions can be explained
using based-catalyzed derivatization. By using native, heat-moisture treated and
defatted potato starch, Perera et al. (1997) reported that the reaction conditions used
in hydroxypropylation disrupted double helices within the amorphous regions and
altered crystallite orientation. However, no systematic study has been carried out to
investigate the influence of reaction conditions in the derivatization reaction on the
starch structure and properties using different starch sources.

1.2.8.2.3. Cross-linking
Cross-linking introduces covalent bonds which are much stronger than hydrogen
bonds (Fig. 5C). This makes the granules more difficult to disrupt by chemical or
mechanical means. Cross-linked starches are thus less susceptible to acid attack or
shear, and are widely used in the food industry as thickeners in food, particularly
where high and stable viscosity is needed. The following agents can be used to crosslink food grade starches: monosodium phosphate (SOP), sodium trimetaphosphate
(STMP), sodium triphosphate, epichlorohydrin, phosphoryl chloride (POCl3), a
mixture of adipic and acetic anhydrides, and a mixture of succinic anhydride and
vinyl acetate (Woo and Seib, 1997). Phosphoryl chloride is an efficient cross-linking
agent in aqueous slurry at pH > 11 in the presence of a neutral salt (Wu and Seib,
1990).

42

By X-ray diffraction analysis, Hoover and Sosulski (1986) reported cross-linking


occurs in the amorphous regions of the starch granule. Jane et al. (1992) reported that
amylopectin molecules seem to be more cross-linked than amylose. Hood and
Mercier (1978) speculated that cross-linked sites are located at the amorphous region
of the amylopectin. Much research has been reported on cross-linked starch and its
influence on starch properties. Decreased water binding capacity, swelling power, amylase digestibility and 95 C viscosity but increased setback was reported for
cross-linked legume starches (Hoover and Sosulski, 1986). Liu et al. (1999) found
increased shear stability, gel hardness and decreased swelling power and solubility in
cross-linked normal and waxy rice starches but increased peak viscosity, pasting
temperature and heat of gelatinization was found only for cross-linked waxy rice
starches. Decreased trend of swelling power for cross-linked waxy barley starch has
been reported when the level of cross-linking increased (Wu and Seib, 1990). Greater
increase of gel hardness (stiffness) was observed for cross-linked potato starch
(Bohlin and Eliasson, 1986). The properties of cross-linked starches vary widely,
depending on the level of cross-linking (Yoneya, et al., 2003; Wurzburg and
Szymanski, 1970). At higher level of treatment, the granules are unable to swell fully.
If the degree of cross-linking is high enough, the granules will eventually reach a
point where they will not cook under normal conditions. It has been reported that
cross-linking has no influence on starch retrogradation and thus provides no
advantages in systems where storage stability is essential (Wurzburg and Szymanski,
1970). In a study of cross-linking of hydroxyl propylated potato starch, Kim and
Eliasson (1993) reported that cross-linking can reduce the freeze-thaw stability of
hydroxypropylated starch.

1.2.8.2.4. Acid modification

43

In acid modification, acid molecule (hydronium ion) attacks the glycosidic oxygen
atom and hydrolyzes the glycosidic linkage. In the typical procedure of
manufacturing acid-thinned starch, starch slurry (30-40% solids) is treated with
mineral acids at temperatures below gelatinization (40-60 C) for a certain time
period depending on the desired viscosity or degree of conversion (Wurzburg, 1986).
Acid modification changes the physicochemical properties of starch without
destroying its granule structure, and the properties of acid-thinned starches differ
according to their origin. Acid hydrolysis follows two distinctive steps, where initial
attack is to the less compact amorphous region at a fast rate and in the second step,
acid molecules attack the more ordered crystalline region at a slower rate (Hoover
and Vasanthan, 1994; Inouchi et al., 1987). Glycosidic oxygen atom of starch double
helices is buried in the interior, which prevents the accessibility of hydronium ion to
the glycosidic oxygen in the starch crystallites (French, 1984). Wang and Wang (2001)
reported that acid primarily attacks the amorphous regions within the starch granules
and both amylose and amylopectin are hydrolyzed simultaneously. The same authors
reported that acid modification decreased the longer chain fraction and increased the
shorter chain fraction of maize and rice starch but the opposite was found for potato
starch. Acid molecules preferentially attack the branched molecules rather than the
linear molecules (BeMiller, 1965). Increased shorter chain fraction and decreased
degree of branching could be the reason for relatively greater gelling power or
retrogradation of the acid-modified starches (Rohwer and Klem, 1984). After
studying lintnerized non-waxy and waxy barley starch by

13

C CP/MAS-NMR,

Morrison et al. (1993c) reported that lintnerization increased the double helical
content relative to no-ordered conformation due to the retrogradation of free amylose
on lintnerization. Atichokudomchai et al. (2004) reported that the initial increase in
relative double helix content and crystallinity in acid-modified tapioca starch was due

44

to a hydrolytic destruction in the amorphous domain, retrogradation of partially


hydrolyzed amylose and crystallinity of free amylopectin double helices.

Amylose-lipid complexes in native cereal starches have been shown to resist the acid
degradation and thus persist through extensive acid treatment (French, 1984). Gelders
et al. (2005) studied the enzyme and acid resistance of amylose-lipid complexes. The
results showed that enzyme and acid resistance increased with increasing amylose DP,
lipid chain length and complexation temperature.

Influence of acid modification on physicochemical properties of starch has been


extensively studied. Gelatinization temperature and the breadth of gelatinization
endotherm have been reported to increase by acid hydrolysis (Shi and Seib, 1992;
Jacobs et al., 1998; Tang et al., 2001; Thirathumthavorn and Charoenrein, 2005;
Atichokudomchai et al., 2002), However, gelatinization enthalpy has been shown to
vary depending on the types of the starches and the hydrolysis reaction conditions
(Hoover and Vasanthan, 1994; Hoover, 2000). The degradation of amorphous region
and cleavage of macromolecules by acid molecules decreased the viscosity of
starches. Decreased viscosity and increased solubility were studied for acid-modified
starches (Kim and Ahn, 1996; Thirathumthavorn and Charoenrein, 2005). More linear
fractions produced by the action of acid increase the gelling power or retrogradation
(Rohwer and Klem, 1984). However, amylopectin retrogradation of gelatinized acidmodified rice starch which was stored at 4 C for 7 days showed no significant
difference compared with the corresponding native starches (Thirathumthavorn and
Charoenrein, 2005). But, in a freeze-thaw study of acid-modified tapioca starch gel,
Atichokudomchai et al. (2002) observed that retrogradation melting enthalpy of
tapioca starch increased with the increase of hydrolyzing time. As they speculated,

45

increasing hydrolysis time may increase the proportion of short chain amylose and
amylopectin molecules, which are able to form double helices, resulting in an
increase in enthalpy. These results show that amylopectin retrogradation studied by
DSC for acid-thinned starches could vary depending on the extent of acid hydrolysis
and the type of starches. Rheological properties of starches have also shown to be
affected by acid modification. More elastic but less rigid gel resulted from acidmodified oat starch (Virtanen et al., 1993). However, increased gel strength for acidmodified starches has also been reported (Kim and Ahn, 1996).

Cleavage of macromolecules during acid treatment could permit much easier


realignment and self-association of these molecules within the starch granule (French,
1984). This may be applied to alter the reaction site for other chemical modification
including acetylation and hydroxypropylation. Thus, prior treatment like acid
modification before other chemical modification may result in novel starch properties
while reducing the concentration of reaction reagents. Acid modification has not been
adequately tested as a prior treatment before other chemical modification techniques.

1.2.9. Cyclodextrin
Cyclodextins are cyclic oligomers composed of six, seven, or eight anhydrous
glucopyranosyl units (, , and respectively) linked together by -1,4-bonds.
Among cyclodextin compounds, -cyclodextrin (cycloheptaamylose) [Fig. 6] is the
most commonly studied. Cycloheptaamylose is derived from starch and its seven
glucose units are linked by -1,4-bonds oriented in a manner to form a hydrophobic
cavity in the center of the molecule. This apolar cavity in the center of the molecule
has the capability of forming an inclusion complex with various organic compounds
such as fatty acids, cholesterol, and volatile flavor compounds. It may thus be used in

46

Fig. 5. A. The reaction of starch with acetic anhydride (acetylation).


B. The reaction of starch with propylene oxide (hydroxypropylation).
C. The reaction of starch with phosphorus oxychloride (cross-linking).

47

many food and pharmaceutical applications such as for cholesterol removal from lard
and egg yolk, and flavor retention and modification of the flavor delivery in food
systems (Chiu et al., 2004; Robert et al., 2004; Yen and Tsui, 1995; Smith et al.,
1995). The various compounds inside the cavity of a cyclodextrin molecule produce a
change in the electronic environment of the atoms of the cyclodextrin (Chiu et al.,
2004). Complex formation primarily depends on the hydrophobicity, size and
geometry of the guest molecule. Ideal guest molecules should have a hydrophobic
body and hydrophilic ends which can form hydrogen bonds to hydroxyl groups on
both sides of the host molecule (Anibarro et al., 2001). Due to the limited physical
space available for guest inclusion, usually -CD makes complexes with guest
molecules on a 1:1 stoichiometic basis (Qi and Hedges, 1995). -cyclodextrin is a
non-toxic (Rao et al., 2000), edible, non-hygroscopic and chemically stable
compound and is a permitted food additive for different purposes such as a processing
aid and flavor carrier for several food products. However, the low solubility of
unmodified -CD limits its range of uses. Therefore, -CD has been modified by
substituting its hydroxyl groups with different chemical groups such as
hydroxypropyl, hydroxymethyl, methyl, and hydroxyethyl (Reguera et al., 2004).
These substituted groups can substantially increase the solubility in water, also can
have an effect on the binding of guest to the -CD molecule.

Although -CD has been well studied for wide range of applications, its potential
application in modifying the starch granule has not been adequately tested. Native
starch granules contain lipids in either free or complexed form which has capability
of complexing with -CD. This could modify the starch structure affecting functional
properties. Work on -CD-wheat starch interaction has revealed its capability in
complexing with starch lipids resulting in the disruption of amylose-lipid complex

48

Glucose atom
Hydrophobic core

Fig. 6. The molecule of -cyclodextrin


Source: Malpezzi et al. (2004).
Hydroxpropyl -cyclodextrin is produced by substituting
hydroxypropyl groups (-OCH2CHOHCH3) to hydroxyl
groups (OH) of glucose atoms in -cyclodextrin.

49

(Kim and Hill, 1984b). The effect of -CD on wheat flour dough properties (Kim
and Hill, 1984a), inhibition of -amylase hydrolysis (Weselake and Hill, 1983), and
pasting properties of wheat starch (Li et al., 2000) have been studied. However, a
systematic study, using different physical and chemical probes to characterize the CD-starch interaction and its influence on starch functional properties still remains to
be performed.

1.2.10. References
Aberle, T.H., Buchard, W., Vorweg, W., and Radosta, S. 1994. Conformation,
contributions of amylose and amylopectin to the properties of starches from various
sources. Starke 46: 329-335.
Ahmad, F.B., and Williams, P.A. 1999. Effect of sugars on the thermal and
rheological properties of sago starch. Biopolymers 50: 401-412.
Alexander, R.J. 1992. Carbohydrates used as fat replacers. In: Developments in
Carbohydrate Chemistry. R.J. Alexander, and H.F. Zobel (Eds.), pp. 343-370.
American Association of Cereal Chemists, St. Paul, MN.
Anibarro, M., Gebler, K., Uson, I., Sheldrick, G.M. and Saenger, W. 2001. cyclodextrin -2,7- dihydroxy-naphthalene 4.6 H20: an unusually distorted macrocycle.
Carbohydr. Res. 333: 251-256.
Atichokudomchai, N., Varavinit, S., and Chinachoti, P. 2002. A study of annealing
and freeze-thaw stability of acid-modified tapioca starches by differential scanning
calorimetry (DCS). Starch/Starke 54: 343-349.
Atichokudomchai, N., Varavinit, S., and Chinachoti, P. 2004. A study of ordered
structure in acid-modified tapioca starch by 13C CP/MAS solid-state NMR.
Carbohydr. Polym. 58: 383-389.
Atwell, W.A., Hood, L.F., Lineback, D.R., Varriano-Marston, E., and Zobel, H.F.
1988. The terminology and methodology associated with basic starch phenomena.
Cereal Foods World 33: 306-311.
Baek, M.H., Yoo, B., and Lim, S-T. 2004. Effects of sugars and sugar alcohols on the
thermal transition and cold stability of corn starch gel. Food Hydrocolloids 18: 133142.
Baldwin, P.M., Alder, J., Davies, M.C., and Melia, C.D. 1994. Holes in starch
granules: confocal, SEM and light microscopy studies of starch granule structure.
Starch 46: 341-346.

50

Banks, W., and Greenwood, C.T. 1971. Amylose: a non helical polymer in aqueous
solution. Polymer 12: 141-145.
Banks, W., and Greenwood, C.T. 1975. Fractionation of the starch granules, and the
fine structures of its components. In: Starch and Its Components. W. Banks and C.T.
Greenwood (Eds.), pp. 5-66. Edinburgh University Press, Edinburgh, UK.
Banks, W., and Greenwood, C.T. 1977. Mathematical models for the action of alphaamylase on amylose. Carbohydr. Res. 57: 301-315.
Bean, M.L., and Osman, E.M. 1958. Behavior of starch during food preparation. II.
Effects of different sugars on the viscosity and gel strength of starch pastes. Food
Res. 24: 665-671.
Bean, M.M., and Yamazaki, T.Y. 1978. Wheat starch gelatinization in sugar solutions
I. sucrose: Microscopy and viscosity effects. Cereal Chem. 55: 936-944.
BeMiller, J.N. 1965. Acid hydrolysis and lytic reactions of starch. In: Starch
Chemistry and Technology. Vol. 1, R.L. Whistler, and E.F. Paschall, (Eds.), pp. 495.
Academic Press New York.
BeMiller, J.N. 1997. Starch modification: challenges and prospects. Starch/Starke 49:
127-131.
Bentacur, A.D., Chel, G.L., and Canizares, H.E. 1997. Acetylation and
characterization of Canavalia ensiformis starch. J. Agric. Food Chem. 45: 378-382.
Biliaderis, C.G. 1982. Physical characteristics, enzymatic digestibility, and structure
of chemically modified smooth pea and waxy maize starches. J. Agric. Food Chem.
30: 925-930.
Biliaderis, C.G. 1991. The structure and interactions of starch with food constituents.
Canadian Journal of Physiology and Pharmacology 69: 60-78.
Biliaderis, C.G. 1992. Characterization of starch networks by small strain dynamic
rheometry. In: Developments in Carbohydrate Chemistry. R.J. Alexander, and H.F.
Zobel, (Eds.), pp. 87-135. American Association of Cereal Chemists, St. Paul, MN.
Biliaderis, C.G. 1998. Structures and phase transition of starch polymers. In:
Polysaccharide Association Structures in Food. R.H. Walter (Ed.), pp. 57-168.
Marcel Dekker. Inc, New York.
Biliaderis, C.G., 1992. Structure and phase transition of food systems. Food
Technology 46: 98-108.
Biliaderis, C.G., and Prokopowich, D.J. 1994. Effect of polyhydroxy compounds on
structure formation in waxy maize starch gelatinization. A calorimetric study.
Carbohydr. Polym. 33: 193-202.
Biliaderis, C.G., and Senaviratne, H.D. 1990. Solute effects on the thermal stability of
glycerol monostearate-amylose complex. Carbohydr. Polym. 13: 185-206.

51

Biliaderis, C.G., and Tonogai, J.R. 1991. Influence of lipids on the thermal and
mechanical properties of concentrated starch gels. J. Agric. Food Chem. 39: 833-840.
Biliaderis, C.G., Tonogai, J.R., Pereze, C.M., and Juliano, B.O. 1993.
Thermophysical properties of milled rice starches as influenced by variety and
parboiling method. Cereal Chem. 70: 512-516.
Biliaderis, C.G., Maurice, T.J., and Vose, J.R. 1980. Starch gelatinization
phenomenon studies by differential scanning calorimetry. J. Food Sci. 45: 1169-1674.
Biliaderis, C.G., Page, C.M., and Maurice, T.J. 1986a. On the multiple melting
transitions of starch/monoglycerides systems. Food Chem. 22: 279-295.
Biliaderis, C.G., Page, C.M., and Maurice, T.J. 1986b. Non-equilibrium melting of
amylose-V complexes. Carbohydr. Polym. 6: 269-288.
Biliaderis, C.G., Page, C.M., Slade, L., and Sirett, R.R. 1985. Thermal behavior of
amylose-lipid complexes. Carbohydr. Polym. 5: 367-389.
Biliaderis, C.G., Tonogai, J.R., Perez, C.M., and Juliano, B.O. 1993. Physical
properties of milled rice starch as influenced by variety and par-boiling method.
Cereal Chem. 70: 512-516.
Blanshard, J.M.V. 1986. The significance of the structure and function of the starch
granules in baked products. In: Chemistry and Physics of Baking. J.M.V. Blanshard
(Ed.), pp. 1-2. Royal Society of Chemistry.
Blanshard, J.M.V. 1987. Starch granule structure and function: A physicochemical
approach. In: Starch Properties and Potential. T. Galliard. (Ed.), pp. 16-54. Critical
Reports in Applied Chemistry, Society for Chemistry and Industry, London.
Bohlin, L., and Eliasson, A.C. 1986. Shear stress relaxation of native and modified
potato starch gels. Starch/Starke. 38: 120-124.
Boyer, C.D., Shannon, J.C. Garwood, D.L., and Creech, R.G. 1976. Changes in starch
granules size and amylose percentage during kernel development in several Zea mays
L. genotypes. Cereal Chem. 53: 327-333.
Buleon, A., Colonna, P., Planchot, V., and Ball, S. 1998. Starch granule; structure and
biosynthesis. Int. J. Biol. Macromol. 23: 85-112.
Butler, L.E., Christianson, D.D., Scheerens, J.C., and Tucson, J.W.B. 1986. Buffalo
gourd root starch. Part IV. Properties of hydroxypropyl derivatives. Starch/Starke 5:
156-159.
Carlson, T.L.G., Larsson, D., Dinh-Nguyen, N., and Krog, K. 1979. A study of the
amylose-monoglyceride complex by Raman spectroscopy. Starch 31: 222-224.
Cheetham, N.W.H., and Tao, L. 1998. Variation in crystalline type with amylose
content in maize starch granules: an X-ray powder diffraction study. Carbohydr.
Polym. 36: 277-284.

52

Chinachoti, P., Kim-Shin, M.S., Mari, F., and Lo, L. 1991. Gelatinization of wheat
starch in the presence of sucrose and sodium chloride: correlation between
gelatinization temperature and water mobility as determined by oxygen-17 nuclear
magnetic resonance. Cereal Chem. 68: 245-248.
Chiotelli, E., Rolee, A., Meste, and M.L. 2000. Effect of sucrose on the
thermomechanical behavior of concentrated wheat and waxy corn starch-water
preparations. J. Agric. Food Chem. 48: 1327- 1339.
Chiu, S.H., Chung, T.W., Giridhar, R., and Wu, W.T. 2004. Immobilization of cyclodextrin in chitosan beads for separation of cholesterol from egg yolk. Food Res.
Int. 37: 217-223.
Choi, S.G., and Kerr, W.L. 2004. Swelling characteristics of native and chemically
modified wheat starches as a function of heating temperature and time. Starch/Starke
56: 181-189.
Chungcharoen, A., and Lund, D.B. 1987. Influence of solutes and water on rice starch
gelatinization. Cereal Chem. 64: 240-243.
Collado, L.S., and Corke, H. 2004. Starch properties and functionalities. In:
Characterization of Cereals and Flours: Properties, Analysis, and Applications. G.
Kaletunc, and K. Breslauer, (Eds.), pp. 473-506. Marcel Dekker, New York
Collado, L.S., Mabesa, R.C., and Corke, H. 1999. Genetic variation in the physical
properties of sweet potato starch. J. Agric. Food Chem. 47: 4195-4201.
Colonna, P., and Mercier, C. 1985. Gelatinization and melting of maize and pea
starches with normal and high amylose genotypes. Phytochem. 24: 1167-1674.
Colonna, P., Leloup, V., and Buleon, A. 1992. Limiting factors of starch hydrolysis.
Eur. J. Clin. Nutr. (Suppl. 2) 46: S17-S32.
Cooke, D., and Gidley, M.J. 1992. Loss of crystallinity and molecular order during
starch gelatinization: origin of the enthalpic transition. Carbohydr. Res. 227: 103-112.
Craig, S.A.S., Manningat, C.C., Seib, P.A., and Hoseney, R.C. Starch paste clarity.
Cereal Chem. 66: 173-182.
Crosbie, G.B. 1991. The relationship between starch swelling properties, paste
viscosity and boiled noodle quality in wheat flours. J. Cereal Sci. 13: 145-150.
DAppolonia, B.L. 1972. Effect of bread ingredients on starch gelatinization
properties as measured by the Amylograph. Cereal Chem. 49: 532-543.
Deffenbaugh, L.B., and Walker, C.E. 1989. Comparison of starch pasting properties
in the Brabender Viscograph and the Rapid Visco-Analyzer. Cereal Chem. 66: 493499.
Dias, F.F., Tekchandani, H.K., and Mehta, D. 1997. Modified starches and their use
by food industries. Indian Food Industry 16: 33-39.

53

Donald, A.M., Waigh, T.A., Jenkins, P.J., Gidley, M.J., Debet, M., and Smith, A.
1997. Internal structure of starch granules revealed by scattering studies. In: Starch,
Structure and Functionality. P.J. Frazier, P. Richmond, and A.M. Donald, (Eds.), pp.
172-179. The Royal Society of Chemistry, Cambridge.
Donovan, A.M. 1979. Phase transitions of the starch-water system. Biopolymers 18:
263-275.
Donovan, J.W., Lorenz, K., and Kulp, K. 1983. Differential scanning calorimetry of
heat-moisture treated wheat and potato starches. Cereal Chem. 60: 381-387.
Doublier, J.L. 1990. Rheological properties of cereal carbohydrates. In: Dough
Rheology and Baked Product Texture. H. Faridi, and J.M. Faubion, (Eds.) p. 111-115.
Van Nostrand Reinhold, New York.
Doublier, J.L., Llamas, G., and Le Meur, M. 1987. A rheological investigation of
cereal starch pastes and gels. Effects of pasting procedures. Carbohydr. Polym. 7:
251-275.
Dreher, M.L., and Berry, J.W. 1983. Buffalo gourd root starch. Part I. properties and
structure. Starch 35: 76-81.
Eerlingen, R.C., Jacobs, H., Van Win, H., Delcour, J.A., 1996. Effect of hydrothermal
treatment on the gelatinization properties of potato starch as measured by differential
scanning calorimetry. J. Therm. Anal. 47: 1229-1246.
Eerlingen, R.C., Jacobs, H., and Delcour, J.A. 1994. Enzyme-resistant starch. V.
Effect of retrogradation of waxy maize starch on enzyme susceptibility. Cereal Chem.
71: 351-355.
Eliasson, A-C. 1992. A calorimetric investigation of the influence of sucrose on the
gelatinization of starch. Carbohydr. Polym. 18: 131-138.
Eliasson, A.C., and Krog, N. 1985. Physical properties of amylose-monoglyceride
complexes. J. Cereal Sci. 3: 239-248.
Eliasson, A.C. 1980. Effect of water content on gelatinization of wheat starch.
Starch/Starke 32:270-272.
Eliasson, A.C. 1985b. Retrogradation of starch as measured by differential scanning
calorimetry. In: New Approaches to Research on Cereal Carbohydrates. R.D. Hill,
and L. Munck, ( Eds.), pp. 93-98. Elsevier Science, Amsterdam.
Eliasson, A.C. 1985a. Starch gelatinization in the presence of emulsifiers: a
morphological study of wheat starch. Starch/Starke 37: 411-415.
Eliasson, A.C. 1986a. On the effects of surface active agents on the gelatinization of
starch- a calorimetric investigation. Carbohydr. Polym. 6: 463-476.
Eliasson, A.C. 1986b. Viscoelastic behavior during the gelatinization of starch. .
Comparison of wheat, maize, potato, and waxy-barley starches. J. Text. Stud. 17: 253265.
54

Eliasson, A.C., and Larsson, K. 1993. Cereals in Bread Making: A Molecular


Colloidal Approach. Marcel Dekker, New York.
Eliasson, A.C., Finstad, H., and Ljunger, G. 1988. A study of amylose-lipid
interactions for some native and modified maize starches. Starch/Starke 40: 95-100.
Eliasson, A.C. 2003. Utilization of thermal properties for understanding baking and
salting processes. In: Characterization of Cereals and Flours: Properties, Analysis,
and Applications. G. Kaletunc, and K. Breslauer, (Eds.) pp. 65-115. Marcel Dekker,
Inc., New York.
Evageliou, V., Richardson, R.K., and Morris, E.R. 2000. Effect of sucrose, glucose,
and fructose on gelation of oxidized starch. Carbohydr. Polym. 42: 261-272.
Evans, I.D., and Haisman, D.R. 1982. The effect of solutes on the gelatinization
temperature range of potato starch. Starch 7: 224-231.
Fannon, J.E., Hauber, R.J., and BeMiller, J.N. 1992. Surface pores of starch granules.
Cereal Chem. 69: 384-388.
Fannon, J.E., Shull, J.M., and BeMiller, J.N. 1993. Interior channels of starch
granules. Cereal Chem. 70: 611-613.
Ferrero, C., Martin, M.N., and Zantzky, N.E. 1994. Cornstarch-xanthan gum
interaction and its effect on the stability during storage of frozen gelatinized
suspensions. Starke 46: 300-305.
Franco, C.M.L., Ciacco, C.F., and Tavares, D.Q. 1995. Effect of heat-moisture
treatment on the enzymatic susceptibility of corn starch granules. Starch/Starke 47:
223-228.
Franco, C.M.L., Preto, S.J.R. Ciacco, C.F., and Tavares, D.Q. 1987. Studies on the
susceptibility of granular cassava and corn starches to enzymatic attack. Part. 2: Study
of the granular structure of starch. Starch/Starke 40: 29-32.
Fredriksson, H., Silverio, J., Andersson, R., Eliasson, A.C., and Aman, P. 1998. The
influence of amylose and amylopectin characteristics on gelatinization and
retrogradation properties of different starches. Carbohydr. Polym. 35:119-134.
French, D. 1984. Organization of starch granules. In: Starch Chemistry and
Technology, R.L. Whistler, J.N. BeMiller, and E.F. Paschal (Eds.), 2nd ed. pp. 183212. Academic Press, New York.
French, D., and Murphy, V.G. 1977. Computer modeling in the study of starch.
Cereal Foods World 22: 61-70.
Fuwa, H., Nakajima, M., Hamada, A., and Glover, D.V. 1997. Comparative
susceptibility to amylases of starches from different plant species and several single
endosperm mutants and their double-mutant combinations with opaque-2 inbred oh
43 maize. Cereal Chem. 54: 230-237.

55

Gallant, D.J., Bewa, M., Buy, Q.H., Bouchet, B., Szylit, O., and Sealy, L. 1982. On
ultra structural nutritional aspects of some tropical tuber starches. Starch 34: 255-262.
Gallant, D.J., Derrien, A., Aumaitre, E., and Guilbot, A. 1973. In vitro degradation of
starch. Studies by transmission and scanning electron microscopy. Starke 25: 56-64.
Galliard, T., and Bowler, P. 1987. Morphology and composition of starch. In: Starch:
Properties and Potential. T. Galliard (Ed.), pp 55-79. Wiley, Chichester, England.
Galloway, G., Biliaderis, C.G., and Stanley, D.W. 1989. Properties and structure of
amylose-glycerol complexes formed in solution or extrusion of wheat flour. J. Food
Sci. 54: 950-957.
Gelders, G.G., Duyck, J.P., Goesaert, H., and Delcour, J.A. 2005. Enzyme and acid
resistance of amylose-lipid complexes differing in amylose chain length, lipid and
complexation temperature. Carbohydr. Polym. 60: 379-389.
Gernat, C., Radosta, S., Anger, H., and Damaschun, G. 1993. Crystalline parts of
three different conformations detected in native and enzymatically degraded starches.
Starch/Starke 45: 309-314.
Gidley, M. 1990. Molecular structures and chain length effects in amylose gelation.
In: Gums and Stabilizers for Food Industry. G.O. Phillips, P.A. Williams, D.J.
Wedlock (Eds.), pp. 89-101. Oxford University Press, Oxford.
Gidley, M.J., and Bociek, S.M. 1985. Molecular organization in starches. A
13
CCP/MAS NMR study. J. Am. Chem. Soc. 107: 7040-7044.
Gidley, M.J., and Bociek, S.M. 1988. 13CCP/MAS NMR studies of amylose inclusion
complex cyclodextrins and amorphous phase of starch granules: Relationship
between glycosidic linkage shift. conformation and solid-state 13C chemical shift. J.
Am. Chem. Soc. 110: 3820-3829.
Gray, J.A., and BeMiller, J.N. 2005. Influence of reaction conditions on the location
of reactions in waxy maize starch granules reacted with a propylene oxide analog at
low substitution levels. Carbohydr. Polym. 60: 147-162.
Greenwood, C.T. 1964. Structure, properties and amylolytic degradation of starch.
Food Technol. 18: 138-141.
Greenwood, C.T., and Thompson, J.A. 1959. A comparison of the starches from
barley and malted barley. Journal of the Institute of Brewing 65: 346-353.
Gudmundsson, M., and Eliasson, A.C. 1990. Retrogradation of amylopectin and the
effects of amylose and added surfactants/emulsifiers. Carbohydr. Polym. 13: 295-315.
Gunaratne, A., and Corke, H. 2004. Starch: Analysis of quality. In: Encyclopedia of
Grain Science. C.W. Wrigley, H. Corke, and C.E. Walker, (Eds.), p. 202-212, Oxford,
Elsevier.

56

Gunaratne, A., and Hoover, R. 2002. Effect of heat-moisture treatment on the


structure and physicochemical properties of tuber and root starches. Carbohydr.
Polym. 49:425-437.
Han, J.A., and BeMiller, 2005. Rate of hydroxypropylation of starches as a function
of reaction time. Starch/Starke 57: 395-404.
Han, X-Z., and Hamaker, B.R. 2002. Partial leaching of granule-associated proteins
from rice starch during alkaline extraction and subsequent gelatinization.
Starch/Starke 54: 454-460.
Han, X-Z., Campanella, O.H., Guan, H., Keeling, P.L., and Hamaker, B.R. 2002.
Influence of maize starch granule-associated protein on the rheological properties of
starch pastes. Part I. Large deformation measurements of paste properties. Carbohydr.
Polym. 49: 315-321.
Hargin, K.D., and Morrison, W.R. 1980. The distribution of acyl lipids in the germ of
four wheat varieties. J. Sci. Food Agric. 31: 877-880.
Hauber, R., BeMiller, J.N., and Fannon, J.E. 1992. Swelling and reactivity of maize
starch granules. Starch/Starke 44: 323-327.
Hinrichs, W., Buttner, G., Steifa, M., Betzel, C.H., Zabel, V., Pfannemuller, B., and
Saenger, W. 1987. An amylose antiparallel double helix at atomic resolution. Science
238: 205-208.
Hizukuri, S. 1985. Relationship between the distribution of the chain length of
amylopectin and the crystalline structure of starch granule. Carbohydr. Res. 141: 295306.
Hizukuri, S. 1986. Polymodel distribution of the chain lengths of amylopectin and its
significance. Carbohydr. Res. 147: 342-347.
Hizukuri, S. 1996. Starch analytical aspects. In: Carbohydrates in Foods, A.C.
Eliasson, (Ed.), pp. 342-429. Marcel Dekker, New York.
Hizukuri, S., Kaneko, T., and Takeda, Y. 1983. Measurement of the chain length of
amylopectins and its relevance to the origin of crystalline polymorphism of starch
granules. Biochemica and Biophysica Acta 760: 188-191.
Hizukuri, S., Shirasaka, K., and Juliano, B.O. 1983. Phosphorus and amylose
branching in rice starch granule. Starch/Starke 348-350.
Hizukuri, S., Tabata, S., and Nikuni, Z. 1970. Studies on starch phosphates. Part I.
Estimation of glucose-6-phosphate residues in starch and the presence of other bound
phosphate(s). Starch 22: 338-343.
Hizukuri, S., Takeda, Y., Maruta, N., and Juliano, B.O. 1989. Molecular structures of
rice starch. Carbohydr. Res. 189: 227-235.
Hizukuri, S., Takeda, Y., Yasuda, M., and Suzuki, A. 1981. Multibranched nature of
amylose and the action of debranching enzymes. Carbohydr. Res. 94: 205-213.
57

Hood, L.F., and Mercier, C. 1978. Molecular structure of unmodified and chemically
modified manioc starches. Carbohydr. Res. 61: 53-66.
Hoover, R. 1995. Starch retrogrdadtion. A review. Food Reviews International. 11:
331-346.
Hoover, R. 1998. Starch lipid interactions. In: Polysaccharide Association Structures
in Food. R.H. Walter, (Ed.), pp. 227-256. Marcel Dakker, New York.
Hoover, R. 2000. Acid treated starches. A review. Food Rev. Int. 16: 369-392.
Hoover, R. 2001. Composition, molecular structure and physicochemical properties
of tuber and root starches: A review. Carbohydr. Polym. 45: 253-267.
Hoover, R. Vasanthan T., Senanayake, N.J., and Martin, A.M. 1994. The effect of
defatting and heat-moisture treatment on the retrogradation of starch gels from wheat,
oat, potato, and lentil. Carbohydr. Res. 261: 13-24.
Hoover, R., and Vasanthan, T. 1992. Studies on isolation and characterization of
starch from oat (Avena nuda) grains. Carbohydr. Polym. 19: 285-297.
Hoover, R., and Vasanthan, T. 1994. Effect of heat moisture treatment on the
structure and physicochemical properties of cereal, legume, and tuber starches.
Carbohydr. Res. 252: 33-53.
Hoover, R., and Hadziyev, D. 1981. Characterization of potato starch and its
monoglyceride complexes. Starch 33: 290-300.
Hoover, R., and Manuel, H. 1995. A comparative study of the physicochemical
properties from two lentil cultivars. Food Chem. 53: 275-284.
Hoover, R., and Manuel, H. 1996. The effect of heat-moisture treatment on the
structure and physicochemical properties of normal maize, waxy maize, dull maize
and amylomaize starches. J. Cereal Sci. 23: 153-162.
Hoover, R., and Senanayake, N. 1996. Effect of sugars on the thermal and
retrogradation properties of oat starches. J. Food Biochem. 20: 65-83.
Hoover, R., and Sosulski, F.W. 1985b. A comparative study of the effect of
acetylation on starches of Phaseolus vulgaris Biotypes. Starch/Starke 37: 397-404.
Hoover, R., and Sosulski, F.W. 1985a. Studies on the functional characteristics and
digestibility of starches from Phaseolus vulgaris biotypes. Starke 37: 181-191.
Hoover, R., Hannouz, D., and Sosulski, F.W. 1988. Effect of hydroxypropylation on
thermal properties, starch digestibility and freeze-thaw stability of field pea (Pisum
sativum cv Trapper) starch. Starke 40: 383-387.
Hoover, R., and Sosulski, F. 1986. Effect of cross-linking on functional properties of
legume starches. Starch/Starke 5: 149-155.

58

Hoseney, R.C. 1994. Principles of Cereal Science and Technology. 2nd ed., pp. 29-34.
American Association of Cereal Chemists, St. Paul, MN.
Hoseney, R.C. 1986. Principles of Cereal Science and Technology. American
Association of Cereal Chemists, St. Paul, MN.
Huber, K.C., and BeMiller, J.N. 1997. Visualization of channels and cavities of corn
and sorghum starch granules. Cereal Chem. 74: 537-541.
Hung, P.H., and Morita, N. 2005. Effects of granule sizes on physicochemical
properties of cross-linked and acetylated wheat starches. Starch/Starke 57: 413-420.
Hung, P.V., and Morita, N. Physicochemical properties of hydroxypropylated and
cross-linked starches from A-type and B-type wheat starch granules. Carbohydr.
Polym. 59: 239-246.
Imberty, A., Buleon, A., Tran, V., and Perez, S. 1991. Recent advances in knowledge
of starch structure. Starch/Starke 43: 375-384.
Inouchi, N., Glover, D.V., and Fuwa, H. 1978. Properties of maize starches following
acid hydrolysis. Starch 39: 284-288.
Islam, N., and Azemi, B.M.N.M. 1997. Rheological properties of calcium treated
hydroxypropyl rice starches. Starch/Starke 49: 136-141.
Jacobs, M., and Delcour, J.A. 1998. Hydrothermal modifications of granular starch
with retention of the granular structure. A review. J. Agric. Food Chem. 46: 28952905.
Jane, J., and Robyt, J.F. 1985. 13CNMR study of the conformation of helical
complexes of amylodextrin and amylose solution. Carbohydr. Res. 140: 21-35.
Jane, J., Kasemsuwan, T., and Chen, J.F. 1996. Phosphorus in rice and other starches.
Cereal Foods World 41: 827-832.
Jane, J., Kasemsuwan, T., Leas, S., Zobel, H.F., and Robyt, J.F. 1994. Anthology of
starch granule morphology by scanning electron microscope. Starch/Starke 46: 121129.
Jane, J., Wang, K.S., and McPherson, A.E. 1997. Branch structure difference in
starches of A and B type X-ray pattern revealed by their Naegeli dextrins. Carbohydr.
Res. 300: 219-227.
Jane, J., Xu, A., Radosavljevic, M., and Seib, P.A. 1992. Location of amylose in
normal starch granules. I. Susceptibility of amylose and amylopectin to cross-linking
reagents. Cereal Chem. 69: 405-409.
Jane, J.L., and Robyt, J.F. 1984. Structure studies of amylose-V complexes and
retrograded amylose by action of -amylase, and a new method for preparing
amylodextins. Carbohydr. Res. 132: 105-118.

59

Jarowenk, W. 1986. Acetylated starches and miscellaneous organic esters. In:


Modified Starches: Properties and Uses. O.B. Wurzburg, (Ed.), pp. 55-88. CRC Press,
Boca Raton, FL.
Jayakody, L., and Hoover, R. 2000. The effect of lintnerisation on cereal starch
granule. Food Res. Int. 38: 665-680.
Jenkins, P.J., Cameron, R.E., and Donald, A.M. 1993. A universal feature in the
structure of starch granules from different botanical sources. Starch 45: 417-420.
Johnson, J.M., Davis, E.A., and Cordon, J. 1990. Interaction of starch and sugar water
measured by electron spin resonance and differential scanning calorimetry. Cereal
Chem. 67: 286-291.
Kaletunc, G., and Breslauer, K. 2003. Calorimetry of pre-and post extruded cereal
flours. In: Characterization of Cereals and Flours: Properties, Analysis, and
Applications. G. Kaletunc, and K. Breslauer, (Eds.), pp. 1-63. Marcel Dekker, New
York.
Karkalas, J., Ma, S., Morrison, W.R., and Pethrick, R.A. 1995. Some factors
determining the thermal properties of amylose inclusion complexes with fatty acids.
Carbohydr. Res. 268: 233-247.
Kasemsuwan, T., and Jane, J. 1994. Location of amylose in normal starch granules. II
Location of phosphodiester cross-linking revealed by phosphorous-31 nuclear
magnetic resonance. Cereal Chem. 72: 169-176.
Kavitha, R. and BeMiller, J.N. 1998. Characterization of hydroxypropylated potato
starch. Carbohydr. Polym. 37: 115-121.
Kawabata, A., Takase, N., Miyoshi, E., Sawayama, S., Kimura, T., and Kudo, K.
1994. Microscopic observation and X-ray diffractometry of heat-moisture treated
starch granules. Starch 46: 463-496.
Kim, H.O., and Hill, D. 1984a. Modification of wheat flour dough characteristics by
cycloheptaamylose. Cereal Chem. 61: 406-409.
Kim, H.O., and Hill, D. 1984b. Physical characteristics of wheat starch granule
gelatinization in the presence of cycloheptaamylose. Cereal Chem. 61: 432-435.
Kim, H.R., and Eliasson, A.C. 1993. Changes in rheological properties of
hydroxypropyl potato starch pastes during freeze-thaw treatments. II. Effect of molar
substitution and cross-linking. J. Tex. Stud. 24: 199-213.
Kim, H.R., Hermansson, and A.M., Eriksson, C.E. 1992. Structural characteristics of
hydroxypropyl potato starch granules depending on their molar substitution.
Starch/Starke 44: 111-116.
Kim, R.H., and Ahn, S.Y. 1996. Gelling properties of acid-modified red bean starch
gels. Agric. Chem. Biotech. 39: 49-53.
Knightly, W.H. 1977. The staling of bread. Bakers Digest. 51: 52-56, 144.
60

Kokini, J.L., Lai, L.S., and Chedid, L.L. 1992. Effect of starch structure on starch
rheological properties. Food Technology 46: 124-139.
Krog, N. 1971. Amylose complexing effect of food grade emulsifiers. Starch 23: 206210.
Kowblansky, M. 1985. Calorimetric investigation of inclusion complexes of amylose
with long-chain aliphatic compounds containing different functional groups.
Macromolecules 18: 1776-1779.
Kuge, T., and Kitamura, S. 1985. Annealing of starch granules-warm water treatment
and heat-moisture treatment. J. Jpn. Soc. Starch Sci. 32: 65-83.
Kugimiya, M., and Donovan, J.W. 1981. Calorimeric determination of the amylose
content of starches based on formation and melting of amylose-lysolecithin complex.
J. Food Sci. 46: 765-770.
Kugimiya, M., Donovan, J.W., and Wong, R.Y. 1980. Phase transition of amyloselipid complexes in starches: a calorimetric study. Starke 32: 265-270.
Kulp, K., and Ponte, J.G. 1981. Staling of white pan bread: fundamental causes.
Critical Reviews in Food Science and Nutrition 15: 1-48.
Kulp, K., and Lorenz, K. 1981. Heat-moisture treatment of starches. 1.
Physicochemical properties. Cereal Chem. 58: 46-48.
Launay, B., Doublier, J.L., and Cuvelier, G. 1986. Flow properties of aqueous
solution and dispersions of polysaccharides. In: Functional Properties of Food
Macromolecules. J.R. Mitchell, and D.A. Led ward, (Eds.), pp. 1-78. Elsevier
Applied Science, London.
Lawal, O.S., and Adebowale, K.O. 2005. Physicochemical characteristics and thermal
properties of chemically modified jack bean (Canavalia ensiformis) starch.
Carbohydr. Polym. 60: 331-341.
Le Bail, P., Bizot, H., Ollivon, M., Keller, G., Bourgaux, C., and Buleon, A. 1999.
Monitoring the crystallization of amylose-lipid complexes during maize starch
melting by synchrotron X-ray diffraction. Biopolymers 50: 99-110.
Leloup, V.M., Colonna, P., Ring, S.G., Roberts, K., and Wells, B. 1992.
Microstructure of amylose gels. Carbohydr. Polym. 18: 189-197.
Levine, H., and Slade, L. 1988. Non-equilibrium behavior of small carbohydrate
water system. Pure and Applied Chemistry 60: 1841-1847.
Li, W.D., Huang, J.C., and Corke, H. 2000. Effect of -cyclodextrin on pasting
properties of wheat starch. Nahrung-Food 44: 164-167.
Lim, H., Setser, C.S., Paukstelis, J.V., and Sobczynska, D. 1992. 17O nuclear
magnetic resonance studies on wheat starch-sucrose water interactions with
increasing temperature. Cereal Chem. 69: 382-386.
61

Lim, S.T., Kasemsuwan, T., and Jane, J. 1994. Characterization of phosphorus in


starches using 31P-NMR spectroscopy. Cereal Chem. 71: 488-493.
Lineback, D.R. 1984. The starch granule organization and properties. Bakers Digest
58: 16-21.
Lineback, S.T., and Rasper, V.F. 1988. Wheat carbohydrates. In: Wheat Chemistry
and Technology. Y. Pomeranz, (Ed.), pp. 277-372. American Association of Cereal
Chemists, St. Paul, MN.
Liu, H.J., Ramsden, L., and Corke, H. 1997. Physical properties and enzymatic
digestibility of acetylated ae, wx and normal maize starch. Carbohydr. Polym. 34:
283-289.
Liu, H.J., Ramsden, L., and Corke, H. 1999. Physical properties of cross-linked and
acetylated normal and waxy rice starch. Starch/Starke 51: 249-252.
Liu, Q., and Thompson, D.B. 1998. Effects of moisture content and different
gelatinization heating temperatures on retrogradation of waxy-type maize starches.
Carbohydr. Res. 314: 221-235.
Lu, T.J., Jane, J.L., Keeling, P.L. 1997. Temperature effects on retrogradation rate
and crystalline structure of amylose. Carbohydr. Polym. 33: 19-26.
Liu, H.J., Ramsden, L., and Corke, H. 1999. Physical properties and enzymatic
digestibility of hydroxypropylated ae, wx, and normal maize starches. Carbohydr.
Polym. 40: 175-182.
Lorenz, K., Kulp, K. 1982. Cereal and root starch modification by heat-moisture
treatment. 1. Physicochemical properties. Starch 34: 50-54.
Maaruf, A.G., Che Man, Y.B., Asbi, B.A., Junainah, A.H., and Kennedy, J.F. 2001.
Gelatinization of sago starch in the presence of sucrose and sodium chloride as
assessed by differential scanning calorimetry. Carbohydr. Polym. 45: 335-345.
Malpezzi, L., Fronza, G., Fuganti, C., Mele, A., and Bruckner, S. 2004. Crystal
architecture conformational properties of inclusion complex, neoheperidin
dihydrochalcone-cyclomaltoheptaose (-cyclodextrin), by X-ray diffraction.
Carbohydr. Res. 339: 2117-2125.
Maningat, C.C., and Juliano, B.O. 1980. Starch lipids and their effect on rice starch
properties. Starch/Starke 32: 32-36
Manners, D.J. 1985. Structural analysis of starch components by debranching
enzymes. In: New Approaches to Research on Cereal Carbohydrates. R.D. Hill and
Munck, L. (Eds.), P. 45-60. Elsevier Science Publishers, Amsterdam.
Mazur, A.K. 1984. Mathematical models of depolymerization of amylose by amylases. Biopolymers 23: 1735-1756.
McPherson, A.E., and Jane, J. 1999. Comparison of waxy potato with other root and
tuber starches. Carbohydr. Polym. 40: 51-70.
62

Meredith, P., Dengate, H.N., and Morrison, W.R. 1978. The lipids of various sizes of
wheat starch granules. Starch/Starke 30: 119-125.
Miles, M.J., Morris, V.J., Orford, P.D., and Ring S.G. 1985. The roles of amylose and
amylopectin in the gelation and retrogradation of starch. Carbohydr. Res. 135: 271281.
Moorthy, S.N. 1994. Tuber crop starches, Sreekariyam, Thiruvanathapuram, Kerala,
India: Central Tuber Crops Research Institute (pp), 1-40
Morgan, K.R., Furneaux, R.H., and Stanley, R.A. 1995. Solid state NMR studies on
the structure of starch granules. Carbohydr. Res. 276: 387-389.
Morris, M.J. 1990. Starch gelation and retrogradation. Trends in Food Science &
Technology 1: 2-6.
Morrison, W.R. 1981. Starch lipids: A reappraisal. Starch/Starke 33: 408-410.
Morrison, W.R. 1995. Starch lipids and how they relate to starch granule structure
and functionality. Cereal Food World 40: 437-446.
Morrison, W.R., and Karkalas, J. 1990. Methods in plant biochemistry. In: Starch,
Vol. 2. pp. 323-352 Academic Press New York.
Morrison, W.R., and Laignelet, B. 1983. An improved colorimetric procedure for
determining apparent and total amylose in cereal and other starches. J. Cereal Sci. 1:
9-20.
Morrison, W.R., Law, R.V., and Snape, C.E. 1993a. Evidence for inclusion
complexes of lipid with V-amylose in maize, rice and oat starches. J. Cereal Sci. 18:
107-111.
Morrison, W.R., Snape, C.E., Gidley, M.J., and Tester, R.F. 1993b. Swelling and
gelatinization of cereal starches. IV. Some effects of lipid-complexed amylose and
free amylose in waxy and normal barley starches. Cereal Chem. 70: 385-391.
Morrison, R., Tester, R.F., Gidley, M.J., and Karkalas, J. 1993c. Resistance to acid
hydrolysis of lipid-complexed amylose and lipid-free amylose in lintnerised waxy and
non-waxy barley starches. Carbohydr. Res. 245: 289-302.
Noda, T., Takahata, Y., Sato, T., Suda, I., Morishita, T., Ishiguro, K., and Yamakawa,
O. 1998. Relationships between chain distribution of amylopectin and gelatinization
properties within the same botanical origin for sweet potato and buckwheat.
Carbohydr. Polym. 37: 153-158.
Obanni, M., and BeMiller, J.N. 1997. Properties of some starch blends. Cereal
Chem. 74: 431-436.
Ooraikul, B., Parker, G.J.K., and Hadziyev, D. 1974. Starch and pectin substances as
affected by a freeze-thaw potato granule process. Journal of Food Science 39: 169177.

63

Oostergetel, G.T., and Van Bruggen, E.F.J. 1993. The crystalline domain in potato
starch granules are arranged in a helical fashion. Carbohydr. Polym. 21: 7-12.
Orford, P.D., Ring, S.G., Carroll, V., Miles, M.J., and Morris, V.J. 1987. The effect
of concentration and botanical source on the gelatinization and retrogradation of
starch. J. Sci. Food Agric. 39: 169-177.
Ortega-Ojeda, F.E., and Eliasson, A.C. 2001. Gelatinization and retrogradation
behavior of some starch mixtures. Starch 53: 520-529.
Ortega-Ojeda, F.E., and Eliasson, A.C. 2004. Gel formation in mixtures of amylose
and high amylopectin potato starch. Carbohydr. Polym. 57: 55-56.
Pal, J., Singhal, P.S., and Kulkarni, P.R., 2002. Physicochemical properties of
hydroxypropyl derivative from corn and amaranth starch. Carbohydr. Polym. 48: 4953.
Perera, C., and Hoover, R. 1998. The reactivity of porcine pancreatic -amylase
towards native, defatted, and heat-moisture treated potato starches before and after
hydroxypropylation. Starch 50: 206-213.
Perera, C., Hoover, R., and Martin, A.M. 1997. The effect of hydroxypropylation on
the structure and physicochemical properties of native, defatted and heat-moisture
treated potato starches. Food Research International 30: 235-247.
Perez, L.A.B., Ramos, S.M.C., Aparicio, A.J ., and Lopez, O..P. 2000. Acetylation
and characterization of banana (Musa paradisiacal) starch. Acta Cientifica
Venezolana 51: 143-149.
Perry, P.A., and Donald, A.M. 2002. The effect of sugars on the gelatinization of
starch. Carbohydr. Polym. 49: 155-165.
Planchot, V., Colonna, P., and Roger, P. 2000. Suitability of starch granule porosity
with regard to starch biosynthesis and susceptibility towards amylolysis. Starch 52:
333-39.
Planchot, V., Colonna, P., Buleon, A., and Gallant, D. 1997. Amylolysis of starch
granules and -glucan crystallites. In: Starch Structure and Functionality. R.J. Frazier,
A.M. Donald, and P. Richmond, (Eds.), pp. 141-152. Royal Society of Chemistry,
Cambridge, UK.
Qi, Z.H., and Hedges, A.R. 1995. Use of cyclodextrins for flavors. In: Flavor
Technology. Vol. 610, pp. 231-243, American Chemical Society, Washington, DC.
Radosta, S., Kettlitz, B., Schierbaeum, F., and Gernat, C. 1992. Studies on rye starch
properties and modification part II. Swelling and solubility behavior of rye starch
granules. Starch 44: 8-14.
Rao, P., Sujatha, D., Raj, K.R., Vishwanathan, S., Narasimhamurthy, K., Saibaba, P.,
Rao., D.N., and Divakar, S. 2000. Safety aspects of residual -cyclodextrin in egg
treated for cholesterol removal. European Food Research and Technology 211: 393395.
64

Rappenecker, G.R., and Zugenmaier, P. 1981. Detailed refinement of crystal structure


of Vh-amylose. Carbohydr. Res. 89: 11-19.
Reddy, L., and Seib, P.A. 2000. Modified waxy wheat starch compared to modified
waxy corn starch. J. Cereal Sci. 31: 25-39.
Reguera, J., Alonso, M., Testera, A.M., Lopez, I.M., Martin, S., and RodriguezCabello, J.C. 2004. Effect of , , and -cyclodextrins on the thermo-responsive
behavior of the elastin-like polymer, poly (VPGVG). Carbohydr. Polym. 57: 293-297.
Richardson, G., Langton, M., Bark, A., and Hermansson, A-M. 2003. Wheat starch
gelatinization-the effect of sucrose, emulsifiers and the physical state of the emulsifier.
Starch 55: 150-161.
Ring, S.G. 1985. Some studies on starch gelation. Starch/Starke 37: 80-83.
Ring, S. G., Colonna, P., Ianson, K.J., Kalicheversk, M.T., Miles, M.J., Morris, V.J.,
and Orford, P.D. 1987. The gelation and crystallization of amylopectin. Carbohydr.
Res. 162: 277-293.
Ring, S.G., Gee, M.J., Whittam, M., Orford, P., and Johnson, I.T. 1988. Resistant
starch: Its chemical form in foodstuffs and effect on digestibility in vitro. Food Chem.
28: 97-109.
Robert, A.K., Linforth, S.T., Hort, J., and Taylor, A.J. 2004. Effect of -cyclodextrin
on aroma release and flavor perception. J. Agric. Food Chem. 52: 2028-2035.
Robin, J.P., Mercier, C., Duprat, F., and Guilbot, A. 1975. Lintnerized starches.
Chromatographic and enzymatic studies of insoluble residues from acid hydrolysis of
various cereal starches, particularly waxy maize starch. Starch 27: 36-45.
Robyt, J.F. 1984. Enzymes in the hydrolysis and synthesis of starch. In: Starch:
Chemistry and Technology. R.L. Whistler, J.N. BeMiller, and E.F. Paschall (Eds.), 2nd
ed., pp. 87-124. Academic Press, London.
Rohwer, R.G., and Klem, R.E. 1984. Acid-modified starch: production and uses. In:
Starch Chemistry and Technology. R.L. Whistler, J.N. BeMiller, E.F. Paschall (Eds.),
2nd ed., pp. 529-541. Academic Press, London.
Rutenberg, M.W., and Solarek, D. 1984. Starch derivatives: production and uses. In:
Starch Chemistry and Technology. R.L. Whistler, J.N. BeMiller, and Paschall, E.F.
(Eds.), 2nd ed., pp. 312-388. Academic Press, London.
Savage, H.L., and Osman, E.M., 1978. Effects of sugars and sugar alcohols on the
swelling of corn starch granules. Cereal Chem. 55: 477-454.
Schoch, T.J., and Maywald, E.C. 1968. Preparation and properties of various legume
starches. Cereal Chem. 45: 564-573.
Seow, C.C., and Thevamalar, K. 1988. Problems associated with traditional
Malaysian starch-based intermediate moisture foods. In: Food Preservation by
Moisture Control. C.C. Seow, (Ed.), pp. 233-252. Elsevier Applied science, London.
65

Seow, C.G., and Thevamalar, K. 1993. Internal plasticization of granular rice starch
by hydroxypropylation: effect on phase transition associated with gelatinization.
Starke 45: 85-88.
Shannon, J.C., and Garwood, D.L. 1984. Genetics and physiology of starch
development. In Starch: Chemistry and Technology. R.L. Whistler, J.N. BeMiller,
and E.F. Paschall (Eds.), 2nd Ed. pp. 25-86. Academic Press, London.
Shi, X., and BeMiller, J.N. 2002. Aqueous leaching of derivatized amylose from
hydroxypropylated common corn starch granules. Starch/Starke 54: 16-19.
Shi, Y.C., and Seib, P.A. 1992. The structure of four waxy starches related to
gelatinization and retrogradation. Carbohydr. Res. 227: 131-145.
Shi, Y.C., and Seib, P.A. 1995. Fine structure of maize starches from wx containing
genotypes of the W64A inbred line in relation to gelatinization and retrogradation.
Carbohydr. Polym. 26: 141-147.
Silverio, J., Fredriksson, H., Andersson, R., Eliasson, A.C., and Aman, P. 2000. The
effect of temperature cycling on the amylopectin retrogradation of starches with
different amylopectin unit-chain length distribution. Carbohydr. Polym. 42: 175-184.
Slade, L., and Levine, H. 1987. Recent advances in starch retrogradation. In: Recent
Developments in Industrial Polysaccharides. S.S. Stivala, V. Cresenzi, and I.C.M.
Dea, (Eds.), pp. 387-430. Gordon and Breach, New York.
Smith, D.M., Awad, A.C., Bennink, M.R., and Gill, J.L. 1995. Cholesterol reduction
in liquid egg-yolk using -cyclodextrin. Journal of Food Science 60: 691-669.
Sopade, P.A., Halley, P.J., and Junming, L.L. 2004. Gelatinization of starch in
mixtures of sugars. II. Application of differential scanning calorimetry. Carbohydr.
Polym. 58: 311-321.
Spies, R.D., and Hoseney, R.C. 1982. Effect of sugar on starch gelatinization. Cereal
Chem. 59: 128-131.
Steeneken, P.A.M. 1989. Rheological properties of aqueous suspensions of swollen
starch granules. Carbohydr. Polym. 11: 23-42.
Stute, R. 1992. Hydrothermal modification of starches: the difference between
annealing and heat-moisture treatment. Starch 44: 205-514.
Suzuki, A., Kaneyama, M., Shibanuma, K., Takeda, Y., Abe, J., and Hizukuri, S.
1992. Characterization of lotus starch. Cereal Chem. 69: 309-315.
Swinkles, J.J.M. 1985. Composition and properties of commercial native starches.
Starke 37: 1-5.
Tabata, S., and Hizukuri, S. 1971. Starch phosphates. Part 2. Isolation of glucose-3phosphate and maltose phosphate by acid hydrolysis. Starch 23: 267-301.

66

Takeda, Y., and Hizukuri, S. 1982. Location of phosphate groups in potato


amylopectin. Carbohydr. Res. 102: 321-337.
Takeda, Y., Hizukuri, S., Takeda, C., and Suzuki, A. 1987. Structures of branched
molecules of amyloses of various origins and the molar fractions of branched and
unbranched molecules. Carbohydr. Res. 165: 139-145.
Takeda, Y., Tokunaga, N., Takeda, C., and Hizukuri, S. 1986. Physicochemical
properties of sweet potato starches. Starch 38: 345-350.
Tang, H.R., Burn, A., and Hills, B. 2001. A proton NMR relaxation study of the
gelatinisation and acid hydrolysis of native potato starch. Carbohydr. Polym. 46: 7-18.
Tegge, G. 1979. Modifizierte starken in Lebensmitteln und technischen Industrieprodukten. Getreide, Mehl, Brot 33: 210-216.
Tester, R., and Morrison, W.R. 1990a. Swelling and gelatinization of cereal starches:
I. Effects of amylopectin, amylose, and lipids. Cereal Chem. 67: 551-557.
Tester, R.F. 1997a. Properties of damaged starch granules: composition and swelling
properties of maize, rice, pea and potato starch fractions in water at various
temperatures. Food Hydrocolloids 11: 293-301.
Tester, R.F. 1997b. Starch: the polysaccharide fractions. In: Starch: Structure and
Functionality. P.J. Frazier, P. Richmond, and A.M. Donald (Eds.), pp. 163-171. The
Royal Society of Chemistry, Cambridge.
Tester, R.F., and Qi, X. 2004. Molecular basis of the gelatinization and swelling
characteristics of waxy barley starches grown in the same location during the same
season. Part I. Composition and alpha-glucan fine structure. J. Cereal Sci. 39: 47-56.
Thirathumthavorn, D., and Charoenrein, S. 2005. Thermal and pasting properties of
acid-treated rice starches. Starch/Starke 57: 217-222.
Tufvesson, F., Wahlgren, M., and Eliasson, A.C. 2003. Formation of amylose-lipid
complexes and effects of temperature treatment. Part I. Monoglycerides.
Starch/Starke 55: 61-71.
Valetudie, J.C., Colonna, P., Bouchet, B., and Gallant, D.J. 1993. Hydrolysis of
tropical tuber starches by bacterial and pancreatic -amylases. Starke 46: 453-457.
Vasanthan, T., and Bhatty, R.S. 1996. Physicochemical properties of small and large
granule starches of waxy, regular and high amylose barleys. Cereal Chem. 73: 199207.
Vasanthan, T., Sosulski, F.W., and Hoover, R. 1995. The reactivity of native and
autoclaved starches from different origins towards acetylation and cationization.
Starch/Starke 47: 135-143.
Virtanen, T., Autio, K., Suortti, T., and Poutanen, K. 1993. Heat-induced changes in
native and acid modified oat starch pastes. J. Cereal Sci. 17: 137-145.

67

Waigh, T.A., Gidley, M.J., Komansheck, B.U., and Donald, A.M. 2000a. The phase
transformations in starch during gelatinization: a liquid crystalline approach.
Carbohydr. Res. 328: 165-176.
Waigh, T.A., Kato, L.K., Donald, A.M., Gidley, M.J., Clark, C.J., and Riekel, C.
2000b Side-chain liquid-crystalline model for starch. Starch/Starke 52: 450-460.
Wang, L., and Wang, Y.J. 2001. Structures and physicochemical properties of acidthinned corn, potato and rice starches. Starch/Starke 53: 570-576.
Wang, Y.J., and Wang, L.W. 2000. Effects of modification sequence on structures
and properties of hydroxypropylated and cross-linked waxy maize starch.
Starch/Starke 52: 406-412.
Ward, K.E.J., Hoseney, R.C., and Seib, P.A. 1994. Retrogradation of maize and
wheat starches. Cereal Chem. 71: 150-155.
Watanabe, T. 1981. Traditional techniques in Japan for food preservation by freezingthawing and drying. In: Water Activity: Influences on Food Quality. L.B. Rockland
and G.F. Stewart, (Eds.), pp. 733-742. Academic Press, New York.
Weselake, R.J., and Hill, R.D. 1983. Inhibition of -amylase-catalyzed starch granule
hydrolysis by cycloheptaamylose. Cereal Chem. 60: 98-101.
Whistler, R.L., and BeMiller, J.N. 1997. Starch. In: Carbohydrates For Food
Scientists. R.L. Whistler, and J.N. BeMiller (Eds.), pp. 117-151. American
Association of Cereal Chemists, St. Paul, MN.
Woo, K. and Seib, P.A. 1997. Cross-linking of wheat starch and hydroxypropylated
wheat starch in alkaline slurry with sodium trimetaphosphate. Carbohydr. Polym. 33:
263-271.
Wootton, M., and Bamunuarachchi, A. 1980. Application of differential scanning
calorimetry to starch gelatinization III. Effects of sucrose and sodium chloride. Starch
32: 126-129.
Wu, Y., and Seib, P.A. 1990. Acetylated and distarch phosphates from waxy barley:
Paste properties and freeze-thaw stability. Cereal Chem. 67: 202-208.
Wurzburg, O.B. 1986. Converted starches. In: Modified Starches: Properties and
Uses. O.B. Wurzburg, (Ed.), pp. 17-40. CRC Press, Boca Raton, FL.
Wurzburg, O.B., and Szymanski, C.D. 1970. Modified starches for the food industry.
J. Agric. Food Chem. 18: 997-1001.
Xu, Y., Miladinov, V., and Hanna, M. 2004. Synthesis and characterization of starch
acetates with high substitution. Cereal Chem. 81: 735-740.
Yamaguchi, M., Kainuma, K., and French, D.J. 1979. Electron microscopy
observation of waxy maize starch. J. Ultrastructure Research 69: 249-261.

68

Yen, G.C., and Tsui, L.T. 1995. Cholesterol removal from a lard-water mixture with
-cyclodextrin. Journal of Food Science 60: 561-564.
Yoneya, T., Ishibashi, K., Hironaka, K., and Yamamoto, K. 2003. Influence of crosslinked potato starch treated with POCl3 on DSC, rheological properties and granule
size. Carbohydr. Polym. 53: 447-457.
Yoo, D., and Yoo, B. 2005. Rheology of rice starch-sucrose composites.
Starch/Starke 57: 254-261.
Yuan, R.G, Thompson, D.B., and Boyer, C.D. 1993. Fine structure of amylopectin in
relation to gelatinization and retrogradation behavior of maize starches from three wxcontaining genotypes in two inbred lines. Cereal Chem. 70: 81-89.
Zeleznak, K.J., and Hoseney, R.C. 1987. Characterization of starch from bread aged
at different temperatures. Starch/ Starke 39: 231-233.
Zobel, H.F. 1992. Starch granule structure. In: Developments in Carbohydrate
Chemistry. R.J. Alexander and H.F. Zobel (Eds.), pp. 1-36. American Association of
Cereal Chemists. St. Paul, MN.
Zobel, H.F. 1988a. Starch crystal transformations and their industrial importance.
Starch 40: 1-7.
Zobel, H.F. 1988b. Molecules to granules: A comprehensive starch review. Starch 40:
44-50.
Zobel, H.F. 1984. Gelatinization of the starch-water system. In: Starch Chemistry and
Technology. R.L. Whistler, J.N. BeMiller, and Paschall, E.F. (Eds.), pp. 285-309.
Academic Press, London.

69

Chapter 2
Effect of Cycloheptaamylose (-Cyclodextrin) on Physicochemical
Properties of Cereal, Tuber and Root Starches
Abstract
Swelling factor, amylose leaching, solubility, gelatinization, and pasting properties of
cereal, tuber and root starches were examined in the presence of -cyclodextrin.
Dissociation parameters of amylose-lipid complexes of native starches were also
tested in the presence of -CD. The influence of -CD on the formation and
dissociation

of

synthesized

(amylose-lysophosphatidylcholine)

amylose-lipid

complex was examined by differential scanning calorimetry. Beta-cyclodextrin


increased swelling factor, amylose leaching, and solubility of all starches, but to a
greater extent in cereal starches. Early swelling and decreased pasting temperature
were observed only in cereal starches in the presence of -CD. Beta-cyclodextrin
decreased the H of native and synthesized amylose-lipid complex and melting
enthalpy was eliminated when amylose-lipid complex was synthesized in the
presence of -CD, however, gelatinization enthalpy of native starches was apparently
unaffected. The foregoing results are consistent with the disruption of amylose-lipid
complex by -CD within the starch granule.

70

2.1. Introduction
Beta-cyclodextrin is a cyclic molecule composed of seven glucose molecules linked
by -(14) linkage. The central hydrophobic cavity in the -CD molecule due to the
special orientation of the glucose molecules provides capability to form inclusion
complexes with various guest molecules such as fatty acids and flavor compounds.
Therefore, it can be speculated that -CD can form inclusion complexes with lipids in
native starch granules resulting in changes in physicochemical properties of the
starches. On the other hand -CD, as an oligosaccharide, could behave as other
polyols such as various sugars. Investigation on the effect of sugar on starch
properties have shown delayed gelatinization (Derby et al., 1975, Spies and Hoseney,
1982, Maaurf et al., 2001), decreased swelling (Savage and Osman, 1978), and
decreased amylose leaching in native starches (Richardson et al., 2003). Kim and Hill
(1984) have observed significant increase of swelling volume, amylose leaching and
solubility in wheat starch due to the presence of -CD, in a manner consistent with
the disruption of amylose-lipid complex. Using Rapid Visco-Analyser, Li et al.
(2001) have shown decreased peak viscosity in the presence of -CD only in low
swelling wheat starches extracted from wheat flour of different genotypes. However,
in a study of viscosity properties of wheat starch in the presence of diluted -CD,
Kim and Hill (1984) found increased peak viscosity for commercial wheat starches
tested using Brookfield viscometer. Therefore, some of the results observed for -CD
treated starches seem contradictory and have not used similar samples and
experimental conditions. In addition it is rather difficult to make strong conclusions
on the effect of -CD on physicochemical properties of starch by testing only a single
starch source employing few approaches. So far very few investigations have been
made to explore possible -CD-starch interaction and its uses in developing new

71

starch products, although -CD has additional benefits such as ability to carry flavor
compounds. The main purpose of this study was to examine the mode of -CD action
on starch using starches from diverse botanical sources.

2.2. Materials and Methods

2.2.1. Materials
Wheat, potato, and rice starches, -cyclodextrin, and L--lysophosphatidylcholine
were purchased from Sigma Chemical Co., (St. Louis, MO). Normal maize starch
was obtained from Starch Australasia Limited (Lane Cove, Australia). Sweet potato
and yam starches were processed in the lab from local samples.

2.2.2. Methods

Total lipids and amylose


Total lipids in starches were extracted with 2:1 chloroform:methanol solvent for 4 hr
using an extraction/desolventizing unit (Soxtec System HT6 Tecator, Sweden). Total
amylose was determined according to the iodine colorimetric method described by
Chrastil (1987).

Amylose leaching
Distilled water (10 mL) was added to starch (20 mg, db) in a screw cap tube. Tubes
were then heated at different temperatures (50-100 C) for 30 min. After cooling to
ambient temperature, samples were centrifuged at 2000 g 10 min. Amylose content
of supernatant (0.1 mL) was estimated as described by Chrastil (1987).

72

Swelling factor
Swelling factor, the ratio of the volume of swollen starch granules to the volume of
dry starch was determined by the method of Tester and Morrison (1990a), when
starch (50 mg, db) was heated at different temperatures (50-100 C) in 5 mL water.

Solubility
The method of Subramanian et al. (1994) with minor modification was used to
determine the solubility of starch. Starch (0.5 g) was heated at different temperatures
(50-100 C) in water (15 mL) with continuous stirring to prevent lump formation. The
slurry was then centrifuged at 3000 g for 10 min. A known aliquot of supernatant
was dried at 130 C overnight. The weight of oven-dried supernatant was back
calculated to the volume of supernatant and the initial weight of dry starch and
express as percent soluble starch.

Synthesis of amylose-lipid complex


The method of Biliaderis (1985) with slight modification was employed to synthesize
the amylose-lipid complex. Potato amylose (400 mg) was dissolved in DMSO (5 mL).
The mixture was then diluted with distilled water at 100 C to 0.8% (w/v)
concentration. To maintain the weight ratio of (5:1) amylose to lipid, 80 mg of lysophosphatidylcholin dissolved in water (4 mL) at 60 C was added to the amylose
solution. The mixture was then heated at 90 C in a water bath with occasional
stirring for three hours. After cooling to room temperature slowly, the suspension
was allowed to stand for three days. Complex recovered by centrifugation at 3000 g
was then washed with water and freeze-dried at 55 C for two days.

Differential scanning calorimetry

73

Gelatinization and dissociation parameters of amylose-lipid complex were measured


using a TA 2920 Modulated DSC Thermal Analyzer differential scanning calorimeter
equipped with a thermal analysis data station (TA Instruments, Newcastle, DE).
Starch (3 mg) was directly measured onto the aluminum DSC pan and distilled water
(3 l) was added with a microsyringe and mixed for homogenization. Pans were
sealed, and allowed to stand for 1 hr at room temperature for even distribution of
water. The scanning temperature and the heating rates were 30-140 C and 5 C/min
respectively. An empty pan was used as reference for all measurements.

Differential scanning calorimetry of synthesized amylose-lipid complex


Distilled water (3 l) was added to synthesized amylose-lipid complex (3 mg)
directly into the aluminum DSC pan and mixed for homogenization. Pans were then
sealed and allowed to stand for 1 hr for equal distribution of water. The scanning
temperature and heating rate were 30-140 C and 5 C/min respectively. An additional
series of samples were prepared in order to form amylose-lipid complex in the DSC
pan as follows. Pure potato amylose (1 mg), -lysophosphatidylcholin (0.2 mg), and
distilled water (6 l) were added to a DSC pan. After sealing, pans were heated at 90
C in an oven for 3 hrs to form the amylose-lipid complex in the DSC pan in the
presence of -CD. Samples were then cooled to room temperature and allowed to
stand for 3 hr. The scanning temperature and the rate of heating were 30-140 C and
10 C/min respectively.

Pasting properties
Pasting properties of starches were determined using a Rapid Visco-Analyser (RVA)
model 3D (Newport Scientific, Warriewood, Australia). Distilled water (25.5 g) was
74

added to starch (2.5 g, db) in a RVA canister to obtain a total constant sample weight
of 28 g. The slurry was then manually homogenized using the plastic paddle to avoid
lump formation before the RVA running. A programmed heating and cooling cycle
was set for 22 minutes, where it was first held at 50 C for 1.0 min, heated to 95 C in
7.5 min, further held at 95 C for 5 min, cooled back to 50 C within 7.5 min and held
at 50 C for 1 min.
Note: -cyclodextrin 1% (w/v) was added instead of water in all above experiments
where appropriate.

2.2.3. Results and Discussion

Swelling factor and solubility


Swelling factor and solubility of native and -CD treated starches heated at different
temperatures (50-90 C) are presented in Fig. 2 and Fig. 1 respectively. Both native
and -CD treated starches showed increased swelling with rise of temperature and it
was more pronounced in cereal starches. Among the native starches, potato starch
exhibited highest swelling as usual. In general, cereal starches swelled less than tuber
and root starches except maize starch (Fig. 2).

-CD, as an oligosaccharide should reduce the starch swelling as shown for other
polyols such as different sugars (Savage and Osman, 1978). However, swelling factor
was strongly increased in all cereal starches (wheat > rice > maize) in the presence of
-CD, while the swelling of tuber and root starches remained unchanged or slightly
increased upon addition of -CD (Fig. 2). Kim and Hill (1984) have demonstrated
that -CD dramatically increased the swelling volume of wheat starch likely by

75

disrupting the amylose-lipid complex. Our results confirm this finding because any
other factors responsible for starch swelling, that could be affected by -CD, should
have significant effects on cereal as well as tuber and root starches. Root and tuber
starches were virtually lipid free (Table 1) and did not show the presence of amyloselipid complex in the differential scanning analysis (Fig. 4). Becker et al. (2001) found
that amylose-lipid complexing can decrease swelling and solubility, restricting water
penetration to starch granules, while Tester and Morrison (1990a) observed an
elevation of swelling factor by 30% after half of the lipid removal from wheat starch.
Solubility of cereal, tuber and root starches was affected similarly to swelling and
amylose leaching by the presence of -CD.

The solubility of all cereal starches was strongly increased by the presence of -CD,
whereas tuber and root starches showed a slight increase of solubility (Fig.1). Higher
solubility increased in cereal starches could be due to the higher starch swelling along
with amylose leaching upon addition of -CD (Fig. 2 and 3).

Amylose leaching
All starches tested showed increased amylose leaching with increasing temperature
with or without -CD. Differences of amylose leaching among native starches could
be attributed to the differences in lipid complexed with amylose, chain interaction of
amylose in the starch granule, chain length, amount of amylose (Table 1), and the
location of amylose in starch granule.

-CD significantly increased the amylose leaching in all cereal starches, whereas it
had only a slight effect on tuber and root starches (Fig. 3). In relation to the factors
that affect amylose leaching in native starches, -CD could probably increase the
76

amylose leaching altering the amylose-lipid complex in cereal starch granules,


because any other effect should similarly influence amylose leaching in tuber and root
starches. It has been reported that lipid-complexed amylose decreased amylose
leaching, in contrast to lipid-free amylose (Becker et al., 2001). Cereal starches are
generally characterized as higher lipid content, while tuber and root starches are
virtually fat-free. The presence of amylose-lipid complex, the V-polymorph in wheat,
maize, and rice starches have been confirmed from

13

CCP/MAS-NMR studies

(Morrison et al., 1993; Morgan et al., 1995). In this study we also clearly observed the
presence of amylose-lipid complex in wheat, maize and rice starch but not in tuber
and root starches (Fig. 4). Kim and Hill (1984) showed that -CD can promote
amylose leaching significantly at the second stage of gelatinization in wheat starch.
We also observed a greater effect on amylose leaching after the gelatinization
temperature of wheat, maize, and rice starches, (Table 1), and more pronounced in
maize starch (Fig. 3).

Defatting was performed to examine whether the effect of -CD on amylose leaching
can be eliminated by lipid removal from the starch granule. Increased amylose
leaching was observed in all cereal starches after defatting (Fig. 3). However,
defatting did not fully eliminate the effect of -CD, still showing increased amylose
leaching in defatted starches by -CD. Morrison (1981) reported that defatting cannot
be fully accomplished, particularly for bound lipids, until the starch granule is
denatured. This could explain that the effect of -CD on defatted starch could be due
to the starch bound lipid, which is mostly found in the amylose-lipid complex.
Defatting slightly increased the amylose leaching in sweet potato starch whereas a
slight decrease was shown in yam starch (Fig. 3). In general, neither defatting nor CD had a major influence on amylose leaching in tuber and root starches, in contrast
77

to cereal starches (Fig. 3). This supports the view that greatly increased amylose
leaching in cereal starches is likely due to the disruption of amylose-lipid complex,
also notable in both native and -CD treated starches is that there is a strong
correlation between swelling factor and amylose leaching (Fig.2 and 3). Tester and
Morrison (1990b) also observed a greater correlation between swelling factor and
leaching of soluble carbohydrates in cereal starches.

Differential scanning calorimetry


Gelatinization parameters and scanning curves of native and -CD treated starches
are presented in Table 2 and Fig. 4 respectively. Differential scanning calorimetric
curves of synthesized amylose-lipid complex using -L- lysophosphatidylcholine and
potato amylose with and without -CD are shown in Fig. 5 and results summarized in
Table 3. Highest gelatinization temperature (Tp) was observed in yam starch, and the
lowest value in wheat starch. Differences in crystallinity, granular architecture, and
composition of starch e.g. amount of amylose, amount of amylose complexed with
lipids, and molecular characteristics of amylopectin may cause the observed
differences of gelatinization in native starches. Among the native starches highest H
was shown in potato starch followed by yam starch and lowest transition enthalpy
was exhibited in rice starch. Tester and Morrison (1990b) have postulated that the
transition enthalpy reflects the quality and the amount of starch crystallites, while
Cooke and Gidley (1992) have proposed the melting enthalpy primarily reflects the
lose of double helical order. Differences in gelatinization temperature range (Tc-To)
suggest that the homogeneity of crystallites in native starches, with lesser values
indicating greater homogeneity thus melting over a narrow temperature range.

78

Slight increase of gelatinization temperature resulted in all starches upon addition of


-CD. However, gelatinization enthalpy is likely unaffected in the presence of -CD
(Table 1). Although -CD, a polyol, slightly delayed the gelatinization of all starches,
swelling, amylose leaching and solubility were increased (more pronounced in cereal
starches), suggesting that the -CD- starch interaction is significantly different to that
of other polyols such as sugars which have been shown to eliminate the starch
swelling (Savage and Osman, 1978), soluble carbohydrate leaching (Richardson et al.,
2003) and solubility.

Differential scanning calorimetric curves obtained at (1:1) starch to water ratio


displayed three transition endotherms in all cereal starches while tuber and root
starches showed only two endotherms except for yam starch which showed only a
single endotherm (Fig. 4). First two endotherms (a double endotherm) that is
dissociated at the temperature range around 50 C-100 C (Table 2) reflects melting of
starch crystals that are formed by amylopectin polymer causing starch gelatinization
enthalpy (Eliasson, 2003). High temperature endotherm, which was always above the
gelatinization temperature range ~110 C (Table 3) and only seen in cereal starches
could be attributed to the melting of amylose-lipid complex (Kugimiya et al., 1980).
The presence of two peaks (double endotherm) at moderate moisture content (25-65%,
in this study 50%) could be due to non-equilibrium moisture distribution or the
differences in the thermal stability of starch crystallites, which are melted over
different temperatures (Kaletunc and Breslauer, 2003). According to the above
explanation crystallites of yam starches, which showed apparently a single endotherm,
should be more homogeneous in terms of thermal stability (cooperative melting
causing a single endotherm) than other starches used in this study.

79

Significant decreases of melting enthalpy of amylose-lipid complex in wheat, maize,


and rice by ~35%, 30% and 20% respectively in the presence of -CD strongly
suggests the disruption of amylose-lipid complex by -CD (Table 3). Decreased
melting enthalpy of the synthesized amylose-lipid complex produced by -Llysophosphatidylcholine and potato amylose further confirms this claim. As observed
in endothermic transition of amylopectin starch crystals, -CD apparently did not
change the transition temperatures in the dissociation of amylose-lipid complex in
both native and synthesized complexes. The observed melting temperature of 104.1
C for synthesized amylose-lipid complex (amyloselysophosphatidylcholine) in the
DSC pan is comparable with the value of 104 C observed at higher water content by
Biliaderis (1985). Elimination of H in amylose-lysophosphatidylcholine complex
that was synthesized in the DSC pan in the presence of -CD indicates that -CD can
prevent the formation of amylose-lipid complex (Table 3).

Pasting properties
Pasting properties of native and -CD treated starches are shown in Table 4. Among
native starches potato starch showed the highest peak viscosity followed by sweet
potato and lowest in yam. Differences in swelling (Fig. 2) release of soluble
carbohydrate (Fig. 3), and the resistance to the disintegration of swollen granules at
higher temperature under shear force may cause the observed differences of peak
viscosities in native starches. In the pasting behavior of cereal, tuber and root starches
in the presence of -CD, all cereal starches showed a similar pattern in the pasting
behavior and similarly tuber and root starches showed a similar pattern among them,
different to that of cereal starches (Fig. 6). Early swelling, decreased pasting
temperature, were only shown in cereals starches with -CD (Fig. 6). Takahashi and
Seib (1988) removed lipid from wheat and maize starches by boiling in 75% ethanol
80

and reduced pasting temperature by 15 C and 4 C respectively, while Doublier et al.


(1987) reported that internal lipid-free oat starch exhibited early swelling and early
solubilization of amylose compared with its native counterpart. Nearly complete lipid
removal in wheat and oat starches defatting with hot propanol:water (3:1) was found
to decrease the pasting temperature (Hoover and Vasanthan, 1992). Apart from this,
decreased peak viscosity was reported for defatted wheat starch (Doublier, 1987,
Hoover and Vasanthan, 1992; Melvin et al., 1979). On the other hand Lorenz and
Kulp (1983) showed that lipid removal from potato starch using 80% methanol did
not significantly change the consistency of the amylogram. This indicates that both
defatting using different solvents and -CD treatment have similar effects such as
reduction of pasting temperature, peak viscosity and early swelling. Kim and Hill
(1984) however, reported that -CD promotes the peak viscosity of wheat starch. Li
et al. (2001) showed that high-swelling wheat starch increased in peak viscosity in the
presence of -CD, while low-swelling wheat starch decreased in peak viscosity, they
further speculated that the differences in lipid distribution in low-and high-swelling
wheat starch granules could be the reason for the above outcome. Highly swollen
starch granules could be more susceptible to disintegration at higher temperature
under shear force which may explain the decreased peak viscosity in wheat and rice
starch by the presence of -CD, on the other hand lipid removal particularly
disruption of amylose-lipid complex may have weakened the rigidity of starch
granule. However, very marginal increase of peak viscosity in -CD treated maize
starch was observed (Table 4). The extent of maize starch swelling in the presence of
-CD was not so remarkable compared with that of wheat and rice starch (Fig. 2),
perhaps due to greater resistance of maize starch towards granular breakdown. These
results for hot paste viscosity and the extent of granular breakdown of cereal starches
in the presence of -CD are similar. Not only disruption of amylose-lipid complex,
81

but other factors such as crystallinity, and properties of amylose and amylopectin also
could affect the starch swelling. Tester and Morrison (1990b) have postulated that
starch swelling is primarily a property of amylopectin. This would explain the
differences of the magnitude of wheat, maize, and rice starch swelling in the presence
of -CD. Slightly increased swelling (Fig. 2) and amylose leaching (Fig. 3) of tuber
and root starches with -CD may contribute to the increased peak viscosity of tuber
and root starches (Table 4), in addition, -CD itself can provide viscosity to the starch
slurry at higher temperature.

Cold paste viscosity of cereal starches were increased by -CD except rice starch and
the extent of this increase was more pronounced in wheat starch (Table 4), in contrast
to the decreased or unchanged cold paste viscosity in all tuber and root starches in the
presence of -CD. In general, the development of cold paste viscosity is correlated
with the reassociation of amylose chains in starch paste on cooling. According to
Miles et al. (1985) the three-dimensional network involved in amylose gelation
requires an entanglement of amylose chains in the starch gel on cooling, and swollen
granules are embedded in the continuous phase of amylose and amylopectin matrix
(Ring, 1985). Hansen et al. (1991) have shown that soluble amylopectin hinders gel
formation ability, while amylose and granular swelling promote the gelation. It can be
concluded that starch gelation could affect the properties and molecular
characteristics of amylose and amylopectin, swelling, and degree of starch chain
interaction. Differences in these factors in native starches could change the extent of
cold paste viscosity and set back observed, but addition of -CD may have different
impacts on starch gelation, such as lipid removal that can affect the amylose
entanglement due to increased hydrophilicity, the extent of swelling, amylose

82

leaching and granular breakdown (shear-thinning). On the other hand possible -CD
starch chains (amylose and amylopectin) interaction could affect the reassociation of
amylose polymer. More precisely, the interplay of all these factors may cause the
differences of cold paste viscosity and set back observed in -CD treated starches
(Table 4).

2.2.4. Conclusions
Very significant increase of swelling factor, amylose leaching, and solubility of
higher lipid content starches (cereals) compared with low lipid content starches (tuber
and root starches) indicates that -CD probably can disrupt the amylose-lipid
complex in cereal starches. Decreased H in both native and synthesized amyloselipid complex in the presence of -CD strongly supports the above claim.
Furthermore, decreased pasting temperature and early swelling, but only in cereal
starches, in the presence of -CD suggests that -CD could remove lipid from native
starch granules. In addition, decreased H in amylose-lipid complex, synthesized in
the presence of -CD, suggests that -CD could prevent the formation of amyloselipid complex.

2.2.5. References
Becker, A., Hill, S.E., and Mitchell, J.R. 2001. Relevance of amylose-lipid complexes
to the behavior of thermally processed starches. Starch/Starke 53: 121-130.
Biliaderis, C.G. 1985. Thermal behaviour of amylose-lipid complexes. Carbohydr.
Polym. 5: 367-389.
Chrastil, J. 1987. Improved colorimetric determination of amylose in starches or
flours. Carbohydr. Res. 159: 154-158.
Cooke, D., and Gidley, M.J. 1992. Loss of crystalline and molecular order during
starch gelatinization: origin of the enthalpic transition. Carbohydr. Res. 227: 103-112.
83

Derby, R.I., Miller, B.S., Miller, B.F., and Trimbo, H.B. 1975. Visual observations of
wheat starch gelatinization in limited water systems. Cereal Chem. 52: 702-713.
Doublier, J.-L., Paton, D., and Llamas, G. 1987. A rheological investigation of oat
starch pastes. Cereal Chem. 64: 21-26.
Eliasson, A.C. 2003. Utilization of thermal properties for understanding baking and
salting processes. In: Characterization of Cereals and Flours: Properties, Analysis,
and Applications. G. Kaletunc, and K. Breslauer, (Eds.), pp. 65-115. Marcel Dekker,
Inc. New York.
Hansen, L.M., Hoseney, R.C., and Faubion, J.M. 1991. Oscillatory rheometry of
starch-water systems: Effect of starch concentration and temperature. Cereal Chem.
68: 347-351.
Hoover, R., and Vasanthan, T. 1992. Studies on isolation and characterization of
starch from oat (Avena nuda) grains. Carbohydr. Polym. 19: 285-297.
Kaletunc, G., and Breslauer, K. 2003. Calorimetry of pre-and posttextruded cereal
flours. In: Characterization of Cereals and Flours: Properties, Analysis, and
Applications. G. Kaletunc, and K. Breslauer, (Eds.) pp. 1-63. Marcel Dekker, Inc.
New York.
Kim, H.O., and Hill, D. 1984. Physical characteristics of wheat starch granule
gelatinization in the presence of cycloheptaamylose. Cereal Chem. 61: 432-435.
Kugimiya, M., Donovan, J.W., and Wong, R.Y. 1980. Phase transition of amyloselipid complexes in starches: a calorimetric study. Starch/Starke 32: 265-270.
Li, W.D., Huang, J.C., and Corke, H. (2000). Effect of -cyclodextrin on pasting
properties of wheat starch. Nahrung-Food 44: 164-167.
Lorenz, K., and Kulp, K. 1983. Physicochemical properties of defatted and heatmoisture treated starches. Starch/Starke 35: 123-129.
Maaruf, A.G., Che Man, Y.B., Asbi, B.A., Junainah, A.H., and Kennedy, J.F. 2001.
Gelatinization of sago starch in the presence of sucrose and sodium chloride as
assessed by differential scanning calorimetry. Carbohydr. Polym. 45: 335-345.
Melvin, M.A. 1979. The effect of extractable lipid on the viscosity characteristics of
corn and wheat starches. J. Sc. Food Agric. 30: 731-735.
Miles, M.J., Morris, V.J., and Ring, S.G. 1985. Gelation of amylose. Carbohydr. Res.
135: 257-269.
Morgan, K.R., Furneaux, R.H., and Larsen, N.G. 1995. Solid -state NMR studies on
the structure of starch granules. Carbohydr. Res. 276: 387-399.
Morrison, W.R. 1981. Starch lipids: A reappraisal. Starch/Starke 33: 408-410.
Morrison, W.R., Law, R.V., and Snape, C.E. 1993. Evidence for inclusion complex of
lipids with V-amylose in maize, rice, and oat starches. J. Cereal Sci. 18: 107-109.
84

Richardson, G., Langton, M., Bark, A., and Hermansson, A.M. 2003. Wheat starch
gelatinization-the effects of sucrose, emulsifier and the physical state of the emulsifier.
Starch/Starke 55: 150-161.
Ring, S.G. 1985. Observations on the crystallization of amylopectin from aqueous
solution. Int. J. Biol. Macromol. 7: 253-254.
Savage, H.L., and Osman, E.M. 1978. Effects of certain sugars and sugar alcohols on
the swelling of corn starch granules. Cereal Chem. 55: 447- 454.
Spies, R.D., and Hoseney, R.C. 1982. Effect of sugar on starch retrogradation. Cereal
Chem. 59: 128-131.
Subramanian, V., Hoseney, R.C., and Bramel-Cox, P. 1994. Shear thinning properties
of sorghum and corn starches. Cereal Chem. 71: 272-275.
Takahashi, S., and Seib, P.A. 1988. Paste and gel properties of prime corn and wheat
starches with and without native lipids. Cereal Chem. 65: 474-483.
Tester, R., and Morrison, W.R. 1990a. Swelling and gelatinization of cereal starches.
. Effects of amylopectin, amylose, and lipids. Cereal Chem. 67: 551-557.
Tester, R.F., and Morrison, W.M. 1990b. Swelling and gelatinization of cereal
starches. . Waxy rice starches. Cereal Chem. 67: 558-563.

85

Table 1. Amylose and total lipid content of cereal, tuber and root starches
Starch
Amylose (%)
Total lipid (%)
Wheat
24.50.2
1.00.03
Maize

24.30.4

0.80.08

Rice

23.10.2

0.70.06

Potato

24.80.1

0.10.03

Sweet potato

22.30.3

0.20.01

Yam
24.10.2
0.10.02
Values are means of triplicate determinations standard deviation.

86

Table 2. Gelatinization parameters of cereal, tuber and root starches with and without -cyclodextrin (1% w/v)
Starch
Treatment
To (C)
Tp (C)
Tc (C)
H (J/g)
Wheat
Water
54.90.2
59.20.1
86.90.1
11.00.1
-CD
55.10.1
59.50.1
86.70.1
10.80.2
Maize

Water
-CD

64.00.3
64.60.2

68.30.1
68.80.1

90.70.1
90.10.1

12.20.1
12.10.1

Rice

Water
-CD

55.80.1
57.00.1

65.20.2
66.70.2

89.30.2
91.60.1

8.90.2
9.00.1

Potato

Water
-CD

58.20.1
58.50.1

62.00.1
62.50.1

80.30.2
80.40.2

13.60.2
14.00.1

Sweet potato

Water
-CD

59.60.2
59.70.1

65.40.1
65.60.2

95.00.1
95.50.1

11.70.2
12.10.1

81.40.2
82.30.1

12.80.1
13.10.3

Yam

Water
64.30.2
69.30.1
-CD
64.40.1
69.60.1
To=onset temperature; Tp=peak temperature; Tc=conclusion temperature; H=enthalpy.
Values are means of triplicate determinations standard deviation.

87

Table 3. Dissociation parameters of amylose-lipid complexes of native wheat,


maize, rice, synthesized and synthesized in DSC pan with and without cyclodextrin (1% w/v)
Source
Treatment To (C)
Tp (C)
Tc (C)
H (J/g)
Wheat

Water
-CD

103.30.1
104.80.3

107.60.3
108.90.1

112.90.3
113.20.1

1.40.2
0.90.1

Maize

Water
-CD

94.40.3
94.00.1

105.20.3
106.50.2

111.90.4
110.80.2

2.30.1
1.60.1

Rice

Water
- CD

97.60.3
98.80.3

108.00.2
108.80.1

114.80.5
115.00.1

2.80.1
2.20.1

Synthesized
AMLC

Water

95.50.2

103.40.4

113.20.6

12.90.4

-CD

96.20.3

103.30.2

113.40.3

10.80.2

Water

101.50.1

104.10.1

108.20.2

10.10.1

Synthesized
AMLC in DSC
pan

-CD
101.60.2 104.30.2 108.70.1
8.80.3
To=onset temperature; Tp=peak temperature; Tc=conclusion temperature.
AMLC=amylose-lipid complex.
Values are means of triplicate determinations standard deviation.

88

Table 4. Pasting properties of cereal, tuber and root starches with and without -cyclodextrin (1% w/v)
Starch
Treatment
PV
HPV
BD
CPV
Wheat
Water
150 0.3
113 0.2
37 0.2
211 0.2
-CD
143 0.5
98 0.3
45 0.7
220 0.6

SB
95 0.8
123 0.3

Maize

Water
-CD

213 0.2
215 0.3

110 0.2
114 0.4

105 0.4
103 0.2

223 0.2
225 0.7

113 0.2
110 0.3

Rice

Water
-CD

114 0.5
110 0.2

74 0.4
68 0.8

41 0.3
45 0.1

137 0.2
133 0.3

61 0.5
67 0.5

Potato

Water
-CD

664 0.3
675 0.6

190 0.2
181 0.4

475 0.2
490 0.3

273 0.4
262 0.6

83 0.2
79 0.5

Sweet
potato

Water

405 0.2

187 0.2

218 0.8

276 0.7

89 0.3

-CD

425 0.7

193 0.6

230 0.7

277 0.2

83 0.3

Water
52 0.2
9 0.5
43 0.2
14 0.1
-CD
57 0.8
11 0.6
46 0.1
16 0.3
PV=peak viscosity; HPV=hot paste viscosity; BD=breakdown; CPV=cold paste viscosity; SB=setback.
Values are means of triplicate determinations standard deviation.
Yam

5 0.2
5 0.5

89

16

20

15

20

14

12

15

10

10

10

4
2

Solubility (%)

50

60

14
12
10
8
6
4
2
0

70

80

90

100

50

60

70

80

90

100

100

100

80

90

100

5
0

90

90

10

80

80

15

70

70

20

60

60

25

5
4

50

50

50

60

70

80

90

100

50

60

70

Temperature (C)
Fig. 1. Solubility of wheat (A); maize (B); rice (C); potato (D); sweet potato (E); and yam (F), with and without -CD. =with water;
=with -CD.

90

50
45
40
35
30
25
20
15
10
5
0

40

35
30
25
20
15
10
5

Swelling factor

0
50

60

70

80

90

100

40

30
25
20
15
10
5
0

50

60

70

80

90

15

20

80

90

100

80

90

100

15

20

30

70

20

25

40

60

25

30

50

50

100

35

60

35

10

10

10

0
50

60

70

80

90

100

50

60

70

80

90

100

50

60

70

Temperature (oC)
Fig. 2. Swelling factor of wheat (A); maize (B); rice (C); potato (D); sweet potato (E), and yam (F), with and without -CD. =with water;
=with -CD.

91

25

30

25

25

20

20
15

15

15

10

10

10

Amylose leaching (%)

0
50

60

70

80

90

100

50

25

16

20

14
12

20

60

70

80

90

100

0
50

15

10

6
4

10
5

2
0
50

60

70

80

90

100

110

20

10
8

90

25

15

70

0
50

60

70

80

90

100

50

60

70

80

90

100

Temperature (oC)
Fig. 3. Amylose leaching of wheat (A); maize (B); rice (C); potato (D); sweet potato (E), and yam (F) starches with and without -CD.
=native with water; =defatted with water; =native with -CD; =defatted with -CD.

92

-0.2

Heat flow (W/g)

-0.3

-0.3

-0.4

b
c

-0.4

i
d

-0.5

-0.5

-0.6

-0.6

f
40

60

80

100

120

40

60

80

100

120

Temperature (C)
Fig. 4. DSC curves of cereal (A), tuber and root (B) starches with and without -CD; a=wheat with water;
b=wheat with -CD; c=maize with water; d=maize with -CD; e=rice with water; f=rice with -CD;
g=potato with water; h=potato with -CD; i=sweet potato with water; j= sweet potato with -CD; k=yam
with water; l=yam with -CD.

93

-3.4
-0.34

-0.36

-3.6

Heat flow (W/g)

-0.38

-3.8

-0.40

-0.42

-4.0

-0.44

-0.46

-4.2

-0.48

-0.50
70

80

90

100

110

120

130

-4.4

98

100

102

104

106

108

110

Temperature (C)

Fig. 5. DSC curves of synthesized amylose-lysophosphatidylcholine complex (A) and the same complex
synthesized in DSC pan (B) with and without -CD. a=with water; b=with -CD.

94

100

800

h
c

80

b
a

150

80
600

Temperature (C)

Viscosity (RVU)

200

100

60

60
400

100

40

200

50

40

10

15

20

20

0
25

Time (Mins)

20

l
0

k
10

15

20

0
25

Fig. 6. Pasting properties of cereal (A), tuber and root (B) starches with and without -CD: a=wheat with water; b=wheat with -CD;
c=maize with water; d=maize with -CD; e=rice with water; f=rice with -CD; g=potato with water; h=potato with -CD; I=sweet
potato with water; j=sweet potato with -CD; k=yam with water; l=yam with -CD.

95

Chapter 3
Influence of Unmodified and Modified Cycloheptaamylose (Cyclodextrin) on Physical Properties and Enzymatic Hydrolysis of
Cereal, Tuber and Root Starches
Abstract
Some physicochemical properties of cereal, tuber and root starches in the presence of
-cyclodextrin (-CD) and hydroxypropyl -cyclodextrin (HP-CD) were examined.
Both cyclodextrin compounds significantly increased swelling factor, amylose
leaching, and solubility of cereal starches while tuber and root starches were less
affected. Gelatinization enthalpy in cereal starches was slightly decreased in the
presence of -CD and HP-CD but in tuber and root starches was not affected. Both
-CD and HP-CD decreased dissociation energy of native (wheat, maize, and rice)
and synthesized (amylose-lysophosphatidylcholine and amylose-stearic acid)
amylose-lipid complex. Reformation of native amylose-lipid complex in cereal
starches was decreased by both -CD and HP-CD. Only in cereal starches the
presence of -CD and HP-CD in starch pasting result in early swelling, and
decreased pasting temperature. Gel hardness was decreased in all starches in the
presence of -CD and HP-CD, but the effect was most pronounced in wheat starch.
All these results are consistent with the disruption of amylose-lipid complex by both
cyclodextrin compounds. Neither -CD nor HP-CD inhibited -amylase hydrolysis.
In comparison with -CD, HP-CD increased swelling factor and solubility more
than did -CD, but decreased gel hardness, transition energy of gelatinization and
amylose-lipid complex in both native and synthesized, and reformation of amyloselipid complex in native starches. However, -CD showed increased amylose leaching
compared to HP-CD, but this was not consistent with results for swelling and
96

solubility. This could be due to the interruption caused by larger HP-CD molecules
in amylose-iodine interaction, which may hinder the development of blue colour. The
greater impact on swelling factor, solubility, dissociation of amylose-lipid complex in
native and synthesized, and reformation of amylose-lipid complex in the presence of
HP-CD compared to -CD indicates that HP-CD has a greater effect on disrupting
the amylose-lipid complex than its unmodified counterpart.

97

3.1. Introduction
Unmodified -cyclodextrin, an oligosaccharide, has been shown to increase amylose
leaching, swelling power, and solubility of wheat starch in a manner consistent with
the disruption of amylose-lipid complex within the starch granules (Kim and Hill,
1984b). These authors further reported that the effect is more pronounced at the
second stage of gelatinization. Prior investigation of this study also showed that
increased amylose leaching, swelling factor, and solubility occurred in cereal starches
in the presence of -CD, but the effect was very slight in tuber and root starches. The
decreased transition energy, both in amylose-lipid complexes of native cereal starches
(wheat, rice and maize) and synthesized amylose-lipid complex (using potato
amylose and lysophophatidylcholine) in the presence of -CD further confirmed the
disruption of amylose-lipid complex by -CD.

Understanding of complex formation and dissociation and their role in determining


functional properties of starch is technologically important for the quality of starchbased food products. Higher amylose leaching and swelling factor in the presence of
-CD could influence gel textural properties such as hardness and cohesiveness. It
has been reported that -CD can inhibit hydrolysis by -amylase with a slight
increase of peak viscosity in wheat starch with -amylase when -CD is present (Kim
and Hill, 1984a). However, Li et al. (2000) reported decreased viscosity for wheat
starch with added bacterial -amylase in the presence of -CD, but increased peak
viscosity for wheat flour under the same conditions, suggesting that -CD has
strongly inhibited wheat endosperm -amylase.

98

So far all previous studies carried out to investigate cyclodextrin-starch interaction


have been on unmodified -CD. However, the effect of hydroxypropyl cyclodextrin, a chemically modified form of -CD with greater solubility in water
than -CD may interact differently with starch. HP-CD is prepared by substituting
hydroxypropyl groups to the hydroxyl groups of the seven glucose molecules that are
oriented to form a hydrophobic core in the center of the molecule. This hydrophobic
core is capable of complexing with guest molecules such as fatty acids and flavor
compounds. Substitution thus may have an effect on complex-forming ability because
of greater hydrophilicity and larger molecular size. This study aimed to compare the
influence of -CD with its modified counterpart, HP-CD on starch properties,
enzymatic hydrolysis (using different types of amylases), and starch gel textural
properties, using starches of diverse botanical origin.

3.2. Materials and Methods

3.2.1. Materials
Wheat, potato, and rice starch, -cyclodextrin (-CD), hydroxypropyl -cyclodextrin
(HP-CD), bacterial and fungal -amylase, -amylase, L--lysophosphatidylcholine,
and stearic acid were purchased from Sigma Chemical Co., (St. Louis, MO,). Normal
(25% amylose) maize starch was obtained from Starch Australasia Limited (Lane
Cove, Australia). Sweet potato and yam starches were processed from local market
samples.

3.2.2. Methods

99

Amylose leaching
Distilled water (10 mL) was added to starch (20 mg, db) in a screw cap tube. Tubes
were then heated at different temperatures (50-100 C) for 30 min. After cooling to
ambient temperature, samples were centrifuged at 2000 g for 10 min. Amylose
content of supernatant (0.1 mL) was estimated as described by Chrastil (1987).

Swelling factor
Swelling factor, the ratio of the volume of swollen starch granules to the volume of
dry starch was determined by the method of Tester and Morrison (1990a), when
starch (50 mg, db) was heated at different temperatures (50-100 C) in 5 mL of water.

Solubility
The method of Subramanian et al. (1994) with minor modification was used to
determine the solubility of starch. Starch (0.5 g) was heated at 80 C temperature in
water (15 mL) with continuous stirring to prevent lump formation. The slurry was
then centrifuged at 3,000 g for 10 min. A known aliquot of supernatant was dried at
130 C overnight. The weight of oven-dried supernatant was back-calculated to the
volume of supernatant and the initial weight of dry starch and expressed as percent
soluble starch.

Synthesis of amylose-lipid complex


The method described by Biliaderis (1985) with slight modifications was employed to
synthesize the amylose-lipid complex. Potato amylose (400 mg) was dissolved in
DMSO (5 mL). The mixture was then diluted with distilled water at 100 C to 0.8%
(w/v) concentration. To maintain the weight ratio of 5:1 amylose to lipid, 80 mg of lysophosphatidylcholine and stearic acid dissolved in water and 4 mL at 60 C was

100

added to the amylose solution. In case of stearic acid, 80 mg of fatty acid was
dissolved in DMSO 4 mL. The mixture was then heated at 90 C in a water bath with
occasional stirring for three hours. After slowly cooling to room temperature, the
suspension was allowed to stand for 3 days. Complex recovered by centrifugation at
8,000 g was then washed with water and freeze-dried at 55 C for two days.

Differential scanning calorimetry


Gelatinization and dissociation parameters of amylose-lipid complex were measured
using a TA 2920 Modulated DSC Thermal Analyzer differential scanning calorimeter
equipped with a thermal analysis data station (TA Instruments, Newcastle, DE).
Starch (3 mg) was directly measured onto the aluminum DSC pan and distilled water
(3 L) was added with a microsyringe and mixed for homogenization. Pans were
sealed, and allowed to stand for 1 hr at room temperature for even distribution of
water. The scanning temperature and the heating rates were 30-140 C and 5 C/min
respectively. An empty pan was used as reference for all measurements.

Differential scanning calorimetry of synthesized amylose-lipid complex


Distilled water (6 l) was added to synthesized amylose-lipid complex (1 mg)
directly into the aluminum DSC pan and mixed for homogenization. Pans were sealed
and allowed to stand for 1 hr for uniform distribution of water. The scanning
temperature and heating rate were 30-140 C and 5 C/min respectively.

Pasting properties
Pasting properties of starches were determined using a Rapid Visco-Analyser (RVA)
model 3D (Newport Scientific, Warriewood, Australia). Distilled water (25.5 g) was

101

added to starch (2.5 g, db) in the RVA canister to obtain a total constant sample
weight of 28 g. The slurry was then manually homogenized using the plastic paddle
to avoid lump formation before the RVA run. A programmed heating and cooling
cycle was set for 22 minutes, where it was first held at 50 C for 1.0 min, heated to 95
C in 7.5 min, further held at 95 C for 5 min, cooled to 50 C within 7.5 min and
held at 50 C for 1 min.

Gel hardness
Gel hardness was determined on the starch gel made in the RVA testing using a TAXT2 Texture Analyzer (Stable Micro Systems, Godalming, Surrey, England). After
RVA testing, the paddle was removed and the starch paste in the canister was covered
by Parafilm and stored at room temperature (20-22 C) for 20 hr. The gel was
compressed at a speed of 0.5 mm/sec to a distance of 10 mm with a 6 mm cylindrical
probe, and gel hardness and adhesiveness were noted. The maximum force peak in
the TPA profile represents the gel hardness.

Enzymatic hydrolysis
In order to examine the effect of -cyclodextrin and hydroxypropyl -cyclodextrin on
enzymatic hydrolysis of starch, a viscoamylometric method was used as described by
Li et al. (2000) with slight modifications. 100 units of different types of amylases
(bacterial from Bacillus species and fungal -amylases from Aspergillus oriyzae and
-amylase from barley) were added to starch slurry in the canister containing cyclodextrin and hydroxypropyl -cyclodextrin just before the RVA testing. The
development of peak viscosity was then noted.
Note: -cyclodextrin or hydroxypropyl -cyclodextrin (MS=0.8) 0.01M was added
instead of water in all above experiments where appropriate.
102

3.3. Results and Discussion


Swelling factor and solubility
All the starches tested showed increased swelling factor with increasing temperature
with or without added cylodextrins (Fig. 1). For native starches, tuber and root
starches showed higher swelling compared to cereal starches. This could be due to the
differences in structural properties of amylose and amylopectin, amount of amylose,
interaction of amylose and amylopectin, crystallinity, amount of phosphorus, and
amount of lipid complexed with amylose. Both -CD and HP-CD significantly
increased swelling factor of all cereal starches whereas tuber and root starches
showed slightly increased or unchanged swelling (Fig. 1). The effect of -CD and
HP-CD on swelling is more pronounced at the second stage of gelatinization
indicating that extensive hydration of starch granules occurred at this stage. Tester
and Morrison (1990a) have shown that granular swelling is primarily a property of
amylopectin and that amylose is a diluent. However, the other factors indicated above
could also influence starch granular swelling. In previous results of this study on the
effect of -CD on cereal, tuber and root starches showed that the increase of swelling
factor by -CD is consistent with disruption of amylose-lipid complex, because other
factors responsible for starch swelling that could be affected by -CD should be
significant in cereal as well as tuber and root starches. Kim and Hill (1984b) showed
that -CD dramatically increases swelling volume of wheat starch in the presence of
-CD, likely by disrupting the amylose-lipid complex. Presence of amylose-lipid
complex which can form a rigid structure in cereal starch granules can restrict both
swelling and amylose leaching preventing water absorption to the starch granules
(Becker et al., 2001). Biliaderis (1998) reported that the amylose-lipid complex in
cereal starches limits granular swelling compared to waxy or lipid free starches.
103

It can be speculated that greater hydrophilicity of HP-CD can inhibit starch swelling
by binding more water molecules, thus limiting the starch hydration, but HP-CD
increased granular swelling, only in cereal starches, and to an even greater extent than
did -CD (Fig. 1). This indicates that HP-CD also capable of disrupting the
amylose-lipid complex as did -CD, perhaps to a greater extent. This claim is further
supported by increased solubility (Table 1), and less energy requirement to dissociate
both native and synthesized amylose-lipid complex in the presence of HP-CD
compared to -CD (Table 2). For tuber and root starches HP-CD slightly increased
the swelling of sweet potato and yam but slightly decreased that of potato starch.

Amylose leaching
Amylose leaching in cereal starches was significantly increased especially above the
gelatinization temperature by both -CD and HP-CD while very slight effects were
observed in tuber and root starches similar to the situation for swelling and solubility.
Differences of amylose leaching of native starches could also be attributed to
differences in amount of amylose, location of amylose, lipid complexed with amylose,
chain length of amylose, and inter-chain interaction of amylose and amylopectin in
the starch granules. It is clear that the significant increase of amylose leaching only in
cereal starches should be due to disruption of amylose-lipid complex, because other
factors should similarly influence tuber and root starches.

Generally cereal starches are characterized as higher lipid content starch in contrast to
tuber and root starches. Many authors have reported that lipid-free amylose typically
leaches out much easier than that of lipid-complexed amylose (Becker et al., 2001;
Tester and Morrison, 1990a; Hoover and Hadziyer, 1981). Prior investigation showed
that defatting increased amylose leaching in cereal starches and also decreased the
104

effect of -CD on amylose leaching. HP-CD slightly eased amylose leaching in all
cereal starches compared with -CD except for maize starch (Fig. 2). This result was
not consistent with those for swelling factor and solubility because HP-CD
increased swelling and solubility of all cereal starches more than did -CD, thus HPCD should increase amylose leaching than that of -CD. Quantitative determination
of amylose leaching was carried out using an iodometric method. The large HP-CD
molecule may interfer with the formation of amylose-iodine complex inhibiting the
development of blue colour which to some extent may explain the observed lower
amylose leaching.

DSC parameters
Gelatinization onset temperature and gelatinization peak temperature were slightly
increased for all starches upon addition of -CD and HP-CD, although the effect of
HP-CD was greater (Table 1). Hydrophilic -CD and HP-CD can limit the water
availability, forming hydrogen bonding with water molecules resulting in delayed
starch gelatinization. However, greater amylose leaching, swelling, solubility, and
decreased onset of pasting (Table 3) are not consistent with inhibition of starch
hydration. This may suggest that -CD and HP-CD interaction with starch is
different from other polyols such as sugars which have been shown to increase the
gelatinization temperature while decreasing swelling and amylose leaching (Savage
and Osman, 1978; Richardson et al., 2001).

All starches showed two transition endotherms during gelatinization and a third,
higher temperature endotherm that showed only in cereal starches and is believed to
be due to the dissociation of amylose-lipid complex (Fig. 3). The low temperature

105

endotherms represent the melting of starch crystals made by amylopectin. The


presence of a double endotherm at moderate moisture content (50% in this study)
could be attributed to the uneven moisture distribution or the presence of crystallites
differing in thermal stability (Kaletunc and Breslauer, 2003). Gelatinization enthalpy
of all cereal starches was slightly decreased in the presence of -CD and HP-CD,
however a previous study carried out at 1% concentration showed that -CD did not
change the gelatinization energy. Extensive structural changes in the amorphous
region (higher amylose leaching and swelling) due to the dissociation of amyloselipid complex in the presence of -CD and HP-CD could facilitate the
destabilization of amylopectin helices in the crystalline region requiring less energy
for the gelatinization process, and could thus explain the slight decrease of H by CD and HP-CD at the concentration of 0.01M (1%<), perhaps extensive hydration
of starch granules could reduce the energy requirement to disrupt the starch structure.

Thermal transition of amylose-lipid complex in native starches in the presence of


cyclodextrins showed that both cyclodextrins decreased the dissociation energy
requirement (Table 2 and Fig. 3). Starch lipid could be either surface or internal. The
internal lipids in the cereals are mostly monoacyl lipids in which lysophospholipids
represent the major component (Morrison, 1981; Hargin and Morrison; 1980).
According to Meredith et al. (1978) wheat starch lipids contain at least 90%
lysophospholipids. Monoacyl lipids that occur both on surface and internally can be
complexed with amylose (Buleon et al., 1998), thus an analysis of transition
parameters of lysophosphatidyl-amylose complex in the presence of cyclodextrin will
provide useful data in understanding the interaction between cyclodextrin and
amylose-lipid complex. Amylose-lipid complex can be crystalline or amorphous
depending on the temperature at which they form (Biliaderis, 1992). Amorphous
106

amylose-lipid complexes are believed to be present in native starch granules, next to


free amylose and lipids (Morrison et al., 1993). Decreased H in synthesized
(amylose-lysophosphatidylcholine and amylose-stearic acid complex) amylose-lipid
complexes (Table 2) in the presence of -CD and HP-CD further confirmed the
dissociation of amylose-lipid complex by -CD and HP-CD. However, transition
temperatures

of

amylose-lysophosphatidylcholine

complex

were

apparently

unchanged in the presence of -CD and HP-CD and also showed only a single
endotherm (one endotherm is an agreement with Biliaderis, 1985) while two
endotherms were shown in amylose-stearic acid complex (Fig. 4) with decreased
onset and peak temperatures. The first endotherm, which dissociates at low
temperature, is due to the uncomplexed fatty acids (Serap and Jackson, 2002) and the
second endotherm dissociated at higher temperature reflects the crystalline amyloselipid complex. The magnitude of the decrease of H in both synthesized and native
amylose-lipid complex (Table 2) was greater with HP-CD than that of -CD,
indicating that HP-CD is more capable of disrupting amylose-lipid complex.
However, decrease of transition temperatures was not consistent.

Dissociated amylose-lipid complex can be reformed on cooling. This concept was


used to examine whether -CD and HP-CD can form new cyclodextrin starch-lipid
complex preventing the formation of amylose-lipid complex in native starch. It has
been reported that fatty acids can form inclusion complex with cycloheptaamylose
(Schlenk and Sand, 1961). Amylose-lipid complex of native starches was dissociated
by heating samples in DSC pans at 120 C in the presence of -CD or HP-CD
followed by rescanning after cooling at room temperature for 20 minutes. In wheat
and maize starches, the observed decrease of H in reformed amylose-lipid complex

107

in the presence of -CD or HP-CD indicated that both cyclodextrin molecules are
capable of complexing with starch lipids, although HP-CD has more effective
(Table 2). Dissociation of reformed amylose-lipid complex in both maize and rice
starch showed a very small endotherm at around 130 C (not shown in the figure)
under the given conditions. This could be attributed to either the dissociation of
reformed amylose-lipid complex or retrograded amylose polymer formed at cooling.

Pasting properties
Pasting behavior of all cereal starches in the presence of -CD or HP-CD showed a
similar trend, while tuber and root starches as a group also showed similar behavior
(Fig. 5). Both -CD and HP-CD caused early swelling, decreased onset of pasting
temperature, and decreased peak viscosity in cereal starches while tuber and root
starches showed unchanged onset of pasting temperature, and early swelling (Fig. 5).
However, -CD increased the peak viscosity of all tuber and root starches but HPCD decreased peak viscosity. Previous results of this study showed that pasting
characteristics of cereal starches were similar to those of defatted starches in the
presence of -CD such as early swelling and decreased pasting temperature. HP-CD
also, caused the same general effects as -CD, although the magnitude of the pasting
properties differed. Decreased peak viscosity by HP-CD compared to -CD is
consistent with granular swelling (Fig. 1) because highly swollen granules are more
liable to disintegrate at higher temperature under shear, perhaps severe disruption of
amylose-lipid complex may weaken the starch structure causing more disintegration
by HP-CD. In addition, Fig. 5 shows that HP-CD causes early onset of pasting
temperature indicating early hydration of starch granules compared to -CD. The
main contributor to the cold paste viscosity and setback is the association of amylose

108

chains leached out from starch granules, thus higher amylose leaching in all cereal
starches in the presence of -CD and HP-CD should increase both cold paste
viscosity and set back. However, only wheat starch showed substantial increase of
cold paste viscosity and setback in the presence of -CD and HP-CD while maize
and rice starch showed slight decreases (Table 3). Lipid removal, especially from the
amylose by -CD and HP-CD may be the reason for higher setback in wheat starch,
since lipid-free amylsoe can associate much easier in contrast to amylose complexed
with lipids. Large molecules of -CD and HP-CD could interfere with the
association of amylose chains, and their physical characteristics such as chain length
may partly explain the decreased cold paste viscosity and setback in maize and rice
starch. The cold paste viscosity and setback of potato starch were decreased in the
presence of -CD and HP-CD while yam and sweet potato starches showed slight
increase or no change.

Gel hardness
When a starch aqueous suspension is subjected to cooling a phase separation in the
water, amylose and amylopectin system occurs. A three-dimensional network, where
the swollen granules are embedded in a continuous matrix of entangled amylose
(Ring, 1985) is then formed as a result of nucleation. In this solid-liquid system
extensive hydrogen bonding between adjacent amylose chains occurs. Theoretically,
amylose leaching could increase the gel hardness while granular swelling decreases it
because highly swollen granules may hinder the association of amylose chains
leading to a weaker gel (Hoover and Vasanthan, 1994). Among the native starches
wheat starch formed the strongest gel and rice starch the weakest gel. Yam starch did
not produce a defined gel for which measurement of hardness was possible. Gel
hardness of all starches decreased in the presence of -CD and HP-CD although
109

both cyclodextrins significantly increased amylose leaching. The extent of gel


hardness decrease was greater in the presence HP-CD, as clearly exhibited in wheat
starch (Table 3). Disruption of amylose-lipid complex and subsequent higher swelling
could weaken the rigidity of swollen starch granules. This can disrupt the integrity of
the starch granules. The decrease in gel hardness with increasing amylose leaching,
therefore could be due to the increasing disruption of the integrity of the starch
granule that can not function as a hard filler phase with reinforcing effects on the
composite strength.

Enzymatic hydrolysis
Kim and Hill (1984a) reported that -CD can inhibit -amylase hydrolysis, finding a
slight increase of peak viscosity of wheat starch containing -amylase. However,
using a Rapid Visco-Analyser Li et al. (2000) showed that -CD did not inhibit amylase hydrolysis of wheat starch, but it did have an inhibitory effect on wheat flour.
In this study, we tested different types of -amylase (fungal and bacteria), and amylase in the presence of -CD and HP-CD. Peak viscosity of all starches was
decreased by the amylases. Among these, bacterial -amylase showed the highest
hydrolysis of starch under the given conditions followed by fungal -amylase and amylase (Table 4). Decrease of peak viscosity was additive in the presence of both CD and HP-CD for all starches with -amylases. This indicates that neither -CD
nor HP-CD could inhibit the -amylase hydrolysis, however, except wheat starch,
peak viscosity was increased in other starches containing -amylase in the presence of
-CD or HP-CD than their corresponding -amylase added samples (Table 5). The
mode of action of -amylase requires an end group to initiate the hydrolysis of (14) linkage. Therefore -amylase cannot hydrolyze either type of cyclodextrin
compounds. Unhydrolyzed cyclodextrin molecules could hinder the reaction between
110

starch and -amylase probably by inhibiting the reaction site of -amylase. This could
be the reason why peak viscosity of starches containing -amylase slightly increased
in the presence of both -CD and HP-CD than their corresponding starches
containing -amylase. However, -amylases that do not need end group to initiate the
hydrolysis process can possibly hydrolysis the cyclodextrin molecules. The
development of peak viscosities in starches containing amylase enzymes in the
presence of -CD and HP-CD was not consistent, thus it is difficult to show which
one has generally more effect on enzyme hydrolysis (Table 4).

3.4. Conclusions
This study clearly demonstrated the disruption of amylose-lipid complex by both CD HP-CD by complexing with starch lipids. The mode of action of -CD and
HP-CD in interacting with starch granules (disruption of amylose-lipid complex)
appeared similar, however, higher magnitude of swelling factor, solubility, early
swelling, decreased transition energy of both native and synthesized amylose-lipid
complex, and complex forming ability with native starch lipids by HP-CD than that
of -CD may indicate HP-CD has greater interaction with starch granules in the
disruption of amylose-lipid complex and formation of a new cyclodextrin-starch lipid
inclusion.

3.5. References
Becker, A., Hill, S.E., and Mitchell, J.R. 2001. Relevance of amylose-lipid complex
to the behavior of thermally processed starches. Starch/Starke 53: 121-130.
Biliaderis, C.G. 1985. Thermal behavior of amylose-lipid complexes. Carbohydr.
Polym. 5: 367-389.

111

Biliaderis, C.G. 1992. Structures and phase transitions of starch in food systems.
Food Technol. 6: 98-109.
Biliaderis, C.G. 1998. Structures and phase transition of starch polymers. In:
Polysaccharide Association Structures in Food. R.H. Walter (Ed.), pp. 57-168.
Marcel Dekker, New York.
Buleon, A., Colonna, P., Planchot, V., and Ball, S. 1998. Starch granule: structure and
biosynthesis. Int. J. Biol. Macromol. 23: 85-112.
Chrastil, J. 1987. Improved colorimetric determination of amylose in starches or
flours. Carbohydr. Res. 159: 154-158.
Hargin, K.D., and Morrison, W.R. 1980. The distribution of acyl lipids in the germ of
four wheat varieties. J. Sci. Food Agric. 31: 877-880.
Hoover, R., and Hadziyev, D. 1981. Characterization of potato starch and its
monoglyceride complexes. Starch 33: 290-300.
Hoover, R., and Vasanthan, T. 1994. Effect of heat-moisture treatment on the
structure and physicochemical properties of cereal, legume, and tuber starches.
Carbohydr. Res. 252: 33-53.
Kaletunc, G., and Breslauer, K. 2003. Calorimetry of pre-and post extruded cereal
flours. In: Characterization of Cereals and Flours: Properties, Analysis, and
Applications. G. Kaletunc, and K. Breslauer, (Eds.), pp. 1-63. Marcel Dekker, New
York.
Kim, H.O., and Hill, D. 1984a. Modification of wheat flour dough characteristics by
cycloheptaamylose. Cereal Chem. 61: 406-409.
Kim, H.O., and Hill, D. 1984b. Physical characteristics of wheat starch granule
gelatinization in the presence of cycloheptaamylose. Cereal Chem. 61: 432-435.
Li, W.D., Huang, J.C., and Corke, H. 2000. Effect of -cyclodextrin on pasting
properties of wheat starch. Nahrung-Food 44: 164-167.
Meredith, P., Dengate, H.N., and Morrison, W.R. 1978. The lipids of various sizes of
wheat starch granules. Starch/Starke 30: 119-125.
Morrison, W.R. 1981. Starch lipids: A reappraisal. Starch/Starke 12: 408-410.
Morrison, W.R., Law, R.V., and Snape, 1993. Evidence for inclusion complex of
lipids with V-amylose in maize, rice, and oat starches. J. Cereal Sci. 18: 107-109.
Richardson, G., Langton, M., Bark, A., and Hermansson, A.M. 2003. Wheat starch
gelatinization-the effects of sucrose, emulsifier and the physical state of the emulsifier.
Starch/Starke 55: 150-161.
Savage, H.L., and Osman, E.M. 1978. Effects of certain sugars and sugar alcohols on
the swelling of corn starch granules. Cereal Chem. 55: 447-454.
112

Ring, S.G. 1985. Observations on the crystallization of amylopectin from aqueous


solution. Int. J. Biol. Macromol. 7: 253-254.
Schlenk, H., and Sand, D.M. 1961. The association of - and -cyclodextrins with
organic acids. J. Am. Chem. Soc. 83: 2312-2316.
Serap, O., and Jackson, D.S. 2002. The impact of thermal events on amylose-fatty
acid complexes. Starch/Starke 54: 593-602.
Subramanian, V., Hoseney, R.C. and Bramel-Cox, P. 1994. Shear thinning properties
of sorghum and corn starches. Cereal Chem. 71: 272-275.
Tester, R., and Morrison, W.M. 1990a. Swelling and gelatinization of cereal starches.
1. Effects of amylopectin, amylose, and lipids. Cereal Chem. 67: 551-557.

113

Table 1. Gelatinization parameters and solubility of cereal, tuber and root


starches in the presence of -CD and HP-CD
Starch
Treatment To (C)
Tp (C)
Tc (C)
Solubility
H (J/g)
(%) at 80 C
Wheat

Water
-CD
HP-CD

54.70.1
55.30.2
55.50.3

59.00.3
59.40.2
59.60.2

86.70.2 10.90.6
86.40.3 9.10.5
86.10.3 8.70.3

5.10.3
16.20.2
18.80.4

Maize

Water
-CD
HP-CD

63.90.5
64.10.1
64.80.4

68.20.1
68.40.3
69.00.1

90.20.1 12.10.3
90.10.4 11.50.2
89.50.1 11.10.3

6.50.2
15.20.1
16.10.1

Rice

Water
-CD
HP-CD

56.20.1
57.40.6
57.40.1

64.70.1
65.20.2
65.30.3

89.40.3 7.70.1
90.30.2 6.90.1
90.20.3 6.70.2

5.50.3
20.30.1
22.10.3

Potato

Water
-CD
HP-CD

57.30.3
58.50.4
58.50.3

61.70.1
62.30.4
62.50.2

80.50.2 14.10.4
80.40.1 14.30.3
80.10.2 14.40.1

5.50.1
7.50.1
8.10.2

Sweet
potato

Water

59.60.5

65.40.3

95.20.1 12.00.3

3.90.1

-CD
HP-CD

59.30.1
60.00.2

64.90.3
65.20.2

95.60.1 12.20.2
96.10.2 12.10.1

4.80.2
5.40.3

Yam

Water
64.30.1
69.30.2
81.40.2 13.20.3
-CD
63.90.2
69.10.3
82.60.4 13.30.1
HP-CD
64.10.3
69.40.1
82.90.2 13.50.1
To=onset; Tp=peak; Tc=conclusion; H=enthalpy.
Values are means of triplicate determinations standard deviation.

15.10.2
17.60.1
18.20.3

114

Table 2. Dissociation parameters of amylose-lipid complexes of native cereal


starches and amylose-lipid complex synthesized in DSC pan; and reformation of
amylose-lipid complex in native starches in the presence of -CDa and HP-CDa
Source
Treatment To (C)
Tp (C)
Tc (C)
H (J/g)
Wheat
Water
103.30.3 107.90.2 112.60.2 1.50.1
100.90.2 105.80.1 113.20.3 0.820.1
-CD
101.70.1 106.50.3 113.60.2 0.520.2
HP-CD
Maize

Water
-CD
HP-CD

95.00.2
93.50.2
93.20.3

105.20.3
105.60.2
102.70.4

112.10.3 2.20.2
110.30.4 1.80.1
110.10.3 1.40.2

Rice

Water
-CD
HP-CD

96.70.4
94.10.1
93.30.1

107.90.2
106.20.1
105.70.4

114.50.3 2.80.3
114.90.5 1.90.2
114.20.2 1.30.1

Synthesized
Water
(lysophospholipid)
-CD
HP-CD

98.10.3

100.80.1

106.80.1 13.00.1

97.40.1
97.90.1

100.50.1
101.00.2

107.10.3 10.50.3
107.80.3 9.40.2

Synthesized
(stearic acid)

Water

94.20.2

97.30.3

103.20.4 8.30.2

-CD
HP-CD

90.90.5
91.70.1

96.80.2
96.80.3

103.10.3 7.10.1
103.20.3 6.30.2

Water
-CD
HP-CD

105.90.1
106.90.2
104.80.3

110.30.1
111.70.1
108.40.2

113.60.5 0.780.1
115.10.3 0.610.2
112.30.4 0.440.2

Re-formation of
native AML
Wheat

Maize

Water
123.50.1 127.70.3 131.20.3 0.660.3
123.50.2 127.50.2 133.10.4 0.500.2
-CD
123.50.5 128.60.3 132.20.3 0.360.1
HP-CD
To=onset; Tp=peak; Tc=conclusion; H=enthalpy.
AML= amylose-lipid complex; a=0.01M.
Values are means of triplicate determinations standard deviation.

115

Table 3. Pasting properties of cereal, tuber and root starches in the presence of CDa and HP-CDa
Starch Treatment PV
HPV
BD
CPV
SB
GH
Wheat

Water
-CD
HP-CD

1480.8
1350.7
1280.7

1120.7 350.5
990.3 360.8
980.6 290.5

Maize

Water
-CD
HP-CD

2120.7
2090.8
2050.6

1101.2 991.3 2200.7 1090.6 460.5


1080.4 1000.6 2180.4 1070.5 460.7
1060.9 990.5 2160.4 1100.3 350.7

Rice

Water
-CD
HP-CD

1100.7
1080.6
1080.9

720.6
680.4
680.4

1381.1 660.8
1291.2 600.6
1331.1 650.5

140.4
120.4
90.6

Potato

Water
-CD
HP-CD

6590.9
6651.2
6400.8

1870.6 4750.4 2740.5 800.4


1800.6 4860.4 2541.3 730.5
6390.9 4591.1 2550.5 750.6

390.4
380.6
350.4

Sweet
potato

Water

4060.6

1860.7 2070.5 2720.4 860.8

440.3

-CD
HP-CD

4180.9
4001.2

1920.7 2201.2 2811.3 880.8


1820.9 2140.8 2670.4 860.5

420.6
480.7

390.8
400.7
410.8

2041.2 930.8 640.7


2210.7 1210.9 530.6
2161.2 1180.7 460.4

Yam

Water
480.6
90.2
370.9 110.7 50.3
510.4
100.4 390.5 140.5 40.2
-CD
440.7
80.5
350.4 130.7 40.5
HP-CD
PV=peak viscosity; HPV=hot paste viscosity; BD=breakdown; CPV=cold paste
viscosity; SB=setback; GH=gel hardness (g); a=0.01M.
Values are means of triplicate determinations standard deviation.

116

Table 4. Development of peak viscosity of cereal, tuber and root starches with -amylase (fungal and bacterial) and
-amylase
Starch
Starch (no enzyme)
-amylase
-amylase
-amylase
(bacterial)
(fungal)
Wheat
1460.8
90.8
320.8
740.7
+-CD
1350.7
90.5
180.6
700.6
+HP-CDa
1280.7
90.6
190.4
660.5
Maize
+-CD
+HP-CD

2120.7
2090.8
2050.6

80.4
80.4
80.9

960.3
910.7
920.6

1781.3
1910.6
1880.8

Rice
+-CD
+HP-CD

1107
1086
1089

60.3
60.3
60.7

120.4
90.5
90.4

590.9
730.6
650.8

Potato
+-CD
+HP-CD

6580.9
6701.2
6400.8

70.2
50.6
50.8

740.6
710.5
780.4

3261.6
3660.4
3451.1

Sweet potato
+-CD
+HP-CD

3880.6
3970.9
3981.2

480.7
400.4
430.6

2501.2
2450.8
2380.6

3461.6
3601.8
3570.9

Yam
520.4
30.4
120.6
+-CD
550.8
30.7
100.5
+HP-CD
480.9
30.5
90.4
Values are means of at least duplicate determinations standard deviation.

160.4
220.3
180.5

117

60

50
45
40
35
30
25
20
15
10
5
0

WHEAT

50
40
30
20

Swelling factor

10
0
50

60

70

80

90

100

60

60

MAIZE

40
30
20
10
0
50

35

POTATO

50

60

70

80

90

70

80

90

100

70

80

90

100

5
0

60

100

10

50

90

15

10

80

20

15

10

70

YAM

25

20

20

60

30

25

30

50

100

SWEET
POTATO

30

40

RICE

50

50

60

70

80

90

100

50

60

Temperature (C)
Fig. 1. Swelling factor of starches in the presence of -CD and HP-CD. =with water; =with-CD; =with HP-CD.

118

Amylose leaching (%)

25

25

WHEAT

25

MAIZE

20

20

20

15

15

15

10

10

10

0
50

25

60

70

80

90

100

50

70

80

90

100

SWEET
POTATO

14

20

60

12

80

90

100

80

90

100

15
10

6
4

70

YAM

10

60

20

10

15

50

25

16

POTATO

RICE

0
50

60

70

80

90

100

0
50

60

70

80

90

100

50

60

70

Temperature (C)
Fig. 2. Amylose leaching of starches in the presence of -CD and HP-CD. = with water; =with -CD =with HP-CD.

119

-0.2

-0.3

-0.4

Heat flow (W/g)

-0.3

-0.5

b
-0.4

c
c

-0.6

-0.5
-0.7

-0.6
40

60

80

100

120

-0.8
40

60

80

100

120

Temperature (C)
Fig. 3. DSC thermograms of representative (A) cereal (wheat), (B) tuber and root (potato) starches in the presence of -CD and HP -CD.
a=with water; b=with -CD; c=with HP -CD.

120

-0.5

-1.0

-1.2

-1.2

-1.0

-1.3

-1.4

Heat flow (W/g)

-1.6

-1.8

-1.5

-1.4

-2.0

-1.5

-2.0

-2.5

-1.6

-2.2

-2.4
60

70

80

90

100

110

120

130

-3.0

40

60

80

100

120

-1.7

85

90

95

100

105

110

115

120

125

Temperature (C)

Fig. 4. DSC curves of (A) synthesized amyloselysophosphatidylcholine , (B) amylose-stearic acid , and (C) reformed amylose-lipid
complex in wheat starch in the presence of water (a), -CD (b), and HP -CD (c).

121

B
100

250

750

100

b
600

60

450

60

40

300

40

20

150

20

80

150

80

100

Temperature (C)

Viscosity (RVU)

200

c
50

0
0

10

15

20

25

0
0

10

15

20

25

Time (Mins)
Fig. 5. Pasting curves of representative (A) cereal (wheat), (B) tuber and root (potato) starches in the presence of -CD and HP-CD.
a=with water; b=with -CD; c=with HP-CD.

122

Chapter 4
Thermal, Pasting and Gelling Properties of Wheat and Potato
Starches in the Presence of Sucrose, Glucose, Glycerol and
Hydroxypropyl -Cyclodextrin
Abstract
Thermal, pasting and gelling properties of wheat and potato starches were studied in
the presence of sucrose, glucose, glycerol, and hydroxypropyl -cyclodextrin (HPCD). Swelling factor of both starches slightly increased up to 20% sucrose and
glucose but decreased at 40% concentration (sucrose > glucose). Glycerol did not
affect swelling factor of wheat starch even at 40% concentration but decreased in
potato starch. Amylose leaching of wheat and potato starches tends to decline above
5% and 1% concentration of sucrose and glucose, respectively. However, similar to
swelling factor the extent of amylose leaching in wheat starch was unaffected in the
presence of glycerol. Gelatinization temperature and enthalpy of both starches were
increased by sucrose, glucose and glycerol in the order of sucrose > glucose >
glycerol. Glucose increased peak viscosity of the two starches more than other
polyols. Cold paste viscosity increased in wheat starch following the order: glucose >
sucrose > glycerol but sucrose was most effective in potato starch. Gel hardness of
wheat starch was increased following the order glucose > sucrose > glycerol but
sucrose was more effective in potato starch. All above results indicate the occurrence
of starch-polyhydroxy interaction which reinforces the starch granules depending on
the botanical source of starch and the type and concentration of polyhydroxy
compound. The influence of HP-CD on the swelling factor, amylose leaching, and
dissociation parameters of amylose-lipid complex in wheat starch is consistent with
the disruption of amylose-lipid complex. This greatly affects the gelatinizing, pasting
and gelling properties of wheat starch.

123

4.1. Introduction
Gelatinization, pasting, and subsequent gel properties of starch are key functional
properties that determine many applications of starch in the food industry. Usually
these properties are not optimal in native starch, and thus need to be modified by
various techniques to suit the relevant end product.

Additives such as sugars are commonly used in starch based-foods in order to


optimize the process operation and cause some textural modification in addition to its
role as a sweetening agent. Employing various techniques, many researchers have
reported that sugar and other related polyhydroxy compounds can delay the
gelatinization process. Suggestions proposed for the mechanism of this delay are (1)
that sugars reduce the water activity in the system due to competition for water with
starch molecules which always favors sugars (DAppolonia, 1972; Spies and Hoseney,
1982), (2) that there is a sugar-starch interaction (Spies and Hoseney, 1982; Chiotelli
et al., 2000; Hoover and Senanayaka, 1996; Baek et al., 2004), (3) that there is less
plasticizing effect of sugar than water (Levine and Slade, 1988; Perry and Donald,
2002). Different sugars, however, substantially differ in the magnitude that they can
increase the gelatinization temperature. Disaccharides increase gelatinization
temperature more than monosaccharides (Baek et al., 2004; Savage and Osman, 1978;
Ahmad and Williams, 1999). Sugars restrict the swelling and amylose leaching of
starch depending on the type of sugar and the concentration (Hoover and Senanayake,
1996; Ahmad and Williams, 1999; Savage and Osman, 1978; Richardson et al., 2003).
Influence of sugars on rheological properties of starch has also been discussed by
several researchers. With increase of sugar concentration, peak viscosity and pasting
temperature of wheat starch increase substantially (DAppolonia, 1972; Bean and
Yamazaki, 1978; Richardson et al., 2003). By promoting polymer-polymer

124

association, sugars can markedly change the conformational ordering and


intermolecular association of starch polymers. This greatly affects the physical and
mechanical properties of starch gel. The effectiveness of promoting conformational
ordering and intermolecular association depends on the type of sugar molecule
(Evageliou et al., 2000).

Although there have been many reports on the effect of sugars on starch
gelatinization, pasting and gelling properties, little or no information is available for
the detailed analysis of these properties involving non-sugar polyols such as glycerol
and hydroxypropyl -cyclodextrin (HP-CD). HP-CD is produced from cyclodextrin by hydroxypropylation of the hydroxyl groups which enhances water
solubility. HP-CD is a permitted food additive in some countries and usually used as
a flavor carrier, extraction of cholesterol from egg yolk, and processing aid. Previous
investigations of this study on -CD and starch interaction showed that the structural
configuration of both modified and unmodified -cyclodextrin can complex with
starch lipids causing disruption of amylose-lipid complex. But not much detailed
information is available on modified -CD-starch interaction. In addition to new
information of HP-CD-starch interaction and its influence on starch functional
properties, this study will lead to deeper insight into sugar-starch interaction by
comparing the HP-CD-starch interaction with other polyol-starch interaction.

4.2. Materials and Methods

4.2.1. Materials

125

Wheat and potato starches, hydroxypropyl -cyclodextrin (HP-CD), D-glucose and


sucrose were purchased from Sigma Chemical Co., (St. Louis, MO). Glycerol was
from Fluka Chemie. Different concentrations of sucrose, glucose, glycerol, and HPCD solutions were prepared as w/v solutions in dw.

4.2.2. Methods

Swelling factor
Swelling factor, the ratio of the volume of swollen starch granules to the volume of
dry starch was determined by the method of Tester and Morrison (1990a), when
starch (50 mg, db) was heated at 85 C in 5 mL water.

Amylose leaching
Distilled water or solution (10 mL) was added to starch (20 mg, db) in a screw cap
tube. Tubes were then heated at 85 C for 30 min. After cooling to ambient
temperature, samples were centrifuged at 2000 g for 10 min. Amylose content of
supernatant (0.1 mL) was estimated as described by Chrastil (1987).

Differential scanning calorimetry


Gelatinization and dissociation parameters of amylose-lipid complex were measured
using a TA 2920 Modulated DSC Thermal Analyzer differential scanning calorimeter
equipped with a thermal analysis data station (TA Instruments, Newcastle, DE).
Starch (3 mg) was directly measured onto the aluminum DSC pan and distilled water
or solution (3 L) was added with a microsyringe and mixed for homogenization.
Pans were sealed, and allowed to stand for 1 hr at room temperature for even

126

distribution ofliquid. The scanning temperature and heating rates were 30-140 C and
5 C/min respectively. An empty pan was used as reference for all measurements.

Pasting properties
Pasting properties of starches were determined using a Rapid Visco-Analyser (RVA)
model 3D (Newport Scientific, Warriewood, Australia). Distilled water or solution
(25.5 g) was added to starch (2.5 g, db) in the RVA canister to obtain a total constant
sample weight of 28 g. The slurry was then manually homogenized using the plastic
paddle to avoid lump formation before the RVA run. A programmed heating and
cooling cycle was set for 22 minutes, where it was first held at 50 C for 1.0 min,
heated to 95 C in 7.5 min, further held at 95 C for 5 min, cooled to 50 C within 7.5
min and held at 50 C for 1 min.

Gel textural analysis


Gel textural properties were determined on the starch gel made in the RVA testing
using a TA-XT2 Texture Analyzer (Stable Micro Systems, Godalming, Surrey,
England). After RVA testing, the paddle was removed and the starch paste in the
canister was covered by Parafilm and stored at 4 C for 7 hr. The gel was compressed
at a speed of 0.5 mm/sec to a distance of 10 mm with a 6 mm cylindrical probe, and
gel hardness and adhesiveness, chewiness, and gumminess were noted. The
maximum force peak in the TPA profile represents the gel hardness while negative
area of the curve provides the adhesiveness, when the probe pulls back to the initial
position. Gumminess and chewiness were also recorded from the results of TPA
profile.

4.3. Results and Discussion


127

Swelling factor and amylose leaching


Swelling factor and amylose leaching of wheat and potato starches at 85 C in the
presence of sucrose, glucose, glycerol and HP-CD are presented (Fig. 1). Consistent
with previous investigations potato starch showed a higher swelling ability than
native wheat starch. Swelling factor of the two starches was slightly increased up to
20% sucrose and glucose concentration but substantially decreased at 40%
concentration with sucrose showing more effect than glucose. The reduction of
swelling factor at higher concentration was greater in wheat starch. These results are
comparable with previous investigations on the effect of sugars on starch swelling
(Ahmad and Williams, 1999; Savage and Osman, 1978). Glycerol did not affect
swelling factor of wheat starch even at 40% concentration, but its effect was more
pronounced in potato starch at 40% concentration.

HP-CD greatly increased swelling factor of wheat starch even at 1% concentration.


For potato starch, HP-CD inhibited swelling similar to the higher concentration of
sucrose, glucose and glycerol but required lower concentration to achieve this. Spice
and Hoseney (1982) proposed that sugar molecules can restrict the mobility and
flexibility of starch chains hence stabilizing the amorphous region of the starch
granule, and that longer sugar chains can bridge more starch chains than shorter
chains. Formation of a bridge between starch chains and sugar molecules restricts
swelling ability depending on the type, concentration, size, and the flexibility of the
polyhydroxy compounds and thus the decreasing order of swelling factor (sucrose >
glucose > glycerol) shown for wheat starch can be explained based on size of the
molecules. This kind of polyol-starch chain interaction could be the reason for
observed amylose leaching reduction (Fig. 1) by tested polyhydroxy compounds
because cross bonded amylose chains are difficult to leach out in contrast to free

128

amylose. However, consistent with earlier investigations, HP-CD significantly


increased amylose leaching of wheat starch in a manner consistent with the disruption
of amylose-lipid complex. This is further supported by higher swelling and decreased
transition parameters of amylose-lipid complex of wheat starch in the presence of
HP-CD (Table 2). Similar to other tested polyol-starch interactions there could be a
HP-CD-starch chain interaction that could negatively affect wheat starch swelling
but increased swelling factor of wheat starch suggests that disruption of amylose-lipid
complex and subsequent hydration of starch granule by HP-CD can dominate the
negative effect of HP-CD-starch chain cross-linking on wheat starch swelling.

Gelatinization
Gelatinization parameters of potato and wheat starches in the presence of different
concentrations of sucrose, glucose, glycerol, and HP-CD are presented (Table1) and
some representative DSC curves are shown (Fig. 2 A & B). Starch gelatinization in
the presence of sugars has been extensively studied by using different techniques such
as DSC (Eliasson, 1992; Wootton and Bamunuarchchi, 1980; Kohyama and Nishinari,
1991), light microscopy (Bean and Yamazaki, 1978), and NMR (Chinachoti et al.,
1991). Among the different methods DSC has been the most widely used.

Inclusion of sucrose, glucose, glycerol and HP-CD increased the gelatinization


temperature of both starches as the concentration increased in the order of HP-CD >
sucrose > glucose > glycerol. However, HP-CD slightly affected the gelatinization
temperature of wheat starch at lower concentration (up to 5%) compared with other
tested polyols and at 40% concentration its effect was lower than that of sucrose but
higher than glucose and glycerol. Recent explanations proposed for the delaying
effect on starch gelatinization upon addition of sugars are decreased water activity,

129

sugar-starch interaction and less plasticization effect of sugar than water. Spice and
Hoseney (1982) proposed that in addition to lowering available water content, sugars
can stabilize the amorphous region by cross linking starch chains (as described under
swelling factor and amylose leaching). A similar suggestion was made by Baek et al.
(2004), Chiotelli et al. (2000) and Hoover and Senanayake (1996) for the existence of
sugar-starch interaction that can stabilize the amorphous region of the starch granule.
These results also showed that delayed gelatinization in the presence of tested polyols
except HP-CD could relate to a possible polyol-starch chain interaction which
appears to stabilize the amorphous region of the starch granules. Decreased swelling
factor, amylose leaching and granular breakdown, particularly at higher concentration
of polyol, indicates the existence polyol-starch chain interaction. The higher the
concentration of polyol the greater the gelatinization temperature increased. After
studying different sugars and their alcoholic derivatives, Baek et al. (2004) suggested
that the number of hydroxyl groups in a sugar molecule plays a key role in
determining the gelatinization temperature. Watase et al. (1992) reported that the
content of equatorial hydroxyl groups of sugar molecules determines the relative
effectiveness of different sugars in stabilizing the intermolecular association of
biopolymers. Ahmad and Williams (1999) also reported the importance of axial or
equatorial hydroxyl groups in the effectiveness of sugar-starch interaction. Hoover
and Senanayake (1996) have shown that sucrose has more interactions with starch
chains than glucose and fructose as it contains more binding sites. The above
investigations could explain the observed differences in the extent of the increased
gelatinization temperature in the presence of sucrose and glucose (sucrose > glucose).
Perhaps similar explanations could be valid for glycerol and HP-CD, where the
smaller molecule of glycerol that has fewer binding sites showed the lowest impact
on starch gelatinization temperature and the larger molecule of HP-CD with more

130

hydroxyl groups showed the highest gelatinization temperature increase. However, at


40% the effect of HP-CD was less than sucrose. This may be due to the problems
associated with preparation of 40% HP-CD where the real concentration may be less
than the expected. Previous investigations have generally shown that disaccharides
increase gelatinization temperature more than monosaccharides. This is in agreement
with observed results for both wheat and potato starches.

In case of HP-CD reduction of gelatinization temperature of wheat starch was not


very obvious at lower concentrations compared with other tested polyols but at higher
concentration it delayed the gelatinization more than other polyols. As we have
discussed before extensive hydration of wheat starch granules due to the disruption of
amylose-lipid complex could require less energy for gelatinization. The way that
HP-CD affects swelling factor, amylose leaching, and gelatinization of potato starch
was similar to that for other polyols. This indicates that HP-CD has weaker
interaction with potato starch.

Conflicting results were reported for the changes of gelatinization enthalpy in the
presence of sugars and other polyols. Decreased (Wootton and Bamunuarachchi,
1980; Chungcharoen and Lund, 1987), unchanged (Eliasson, 1992; Maaruf et al.,
2001; Evans and Haisman, 1982) or increased (Baek et al., 2004; Chiotelli et al., 2000;
Ahmad and Williams, 1999; Sopade et al., 2004) gelatinization enthalpies were
reported in the presence of sugars. In this study increased gelatinization enthalpy of
wheat starch followed the order: sucrose > glucose > glycerol > control > HP-CD,
whereas for potato starch, the order was HP--CD > sucrose > glucose > glycerol. An
increased enthalpy indicates more energy requirement to disrupt the starch structure.
Gelatinization is a swelling-driven process in which expansion of the amorphous

131

region disrupts the crystalline domain due to the stress applied on it (Donovan, 1979).
Thus restricted granular swelling caused by sugars and polyols may require more
energy to pull the crystals apart during gelatinization. Similarly increased swelling
due to the disruption of amylose-lipid complex by HP-CD may need less energy to
destabilize the starch crystalline domain. However, increased gelatinization
temperature cannot be explained solely on the polyols-starch interaction and its
relation to low swelling. For example with up to 20% concentration of sucrose,
glucose and glycerol swelling factor was increased but gelatinization temperature was
not decreased (Table 1). Furthermore gelatinization of wheat starch was delayed at
higher concentration of HP-CD while increasing the swelling. Thus it is difficult to
explain the delayed gelatinization by sugars and other polyols considering a one
factor alone, so that other factors should be considered too. Thus increased
gelatinization temperature and enthalpy by sugar and other polyols could be due to
the interplay of decreased water activity, sugar-starch interaction, change of water
structure, and reduced plasticizing effect of solvent due to sugar-starch interaction.
However, the magnitude of each factor in delaying the gelatinization temperature may
substantially vary depending on the botanical source of starch, types of polyols, and
the sugar-starch-water system.

It has been reported that when sugar concentration is increased in low water content
starch gelatinization systems (1:1 starch water ratio), the double endotherm shifts to a
single endotherm (Eliasson, 1992; Maaruf et al., 2001) but in this study the existence
of double endotherm or biphasic character at higher concentration of all tested polyols
in 1:1 starch:water ratio was observed (Fig 2 a and b). Similar results were reported
for the presence of a double endotherm at higher concentration of sugars in low water
content starch gelatinization systems (Ahmad and Williams, 1999; Sopade et al.,

132

2004). This discrepancy could be due to differences in the heating rate, type of starch
or the way that the sugar:starch:water system prepared. An uneven shape of
gelatinization endotherm shown particularly in wheat starch at higher concentration
of HP-CD could be due to incomplete gelatinization. Substitution of the hydroxyl
groups of -cyclodextrin by hydroxypropyl groups disrupts the network of hydrogen
bonding around the rim of the -CD which increases the interaction of hydroxyl
groups with water. Thus at higher concentration of HP-CD it could bind more water
molecules limiting the available water for gelatinization leading to an incomplete
gelatinization.

For the amylose-lipid complex of wheat starch in the presence of polyols, the melting
temperature was shifted to higher temperature as the concentration of sucrose,
glucose and glycerol increased (sucrose > glucose > glycerol) but corresponding
melting enthalpies seemed slightly decreased with increased concentration (Fig. 2a
and Table 2). This is in agreement with the finding of Eliasson (1992) for the
transition parameters of normal maize amylose-lipid complex in the presence of
sucrose. In contrast, HP-CD dramatically decreases both melting temperature and
enthalpy of amylose-lipid complex with increased concentration. Even at 1% HPCD, the melting enthalpy of amylose-lipid complex was substantially decreased
indicating its capability to destabilize the amylose-lipid complex (Table 2).

Pasting and gel textural properties


Pasting curves of wheat and potato starches at different concentrations of sucrose,
glucose, glycerol, and HP-CD are presented (Fig. 3 a & b) and results summarized
(Table 3). Peak viscosity of wheat and potato starches increased as the concentration
of sucrose, glucose, and glycerol increased. However, addition of HP-CD reduced

133

peak viscosity in both starches. For both starches, peak viscosity increased in the
following order: glucose > sucrose > glycerol. When the concentration of sucrose,
glucose, glycerol, and HP-CD was increased the early onset of viscosity
development was observed in wheat starch but generally potato starch showed
delayed first viscosity increase as the concentration of all tested polyols increased,
particularly after 5% concentration sucrose, glucose, and glycerol and 1% HP-CD.

Pasting properties of starch are primarily related to the swelling and rupture of the
starch granules. Richardson et al. (2003) demonstrated by microscopic analysis that
starch granules can keep together better in sucrose solution than in water. They
further explained that early viscosity increase with the addition of sugar could be
attributed to the influence of sugar on the close packing concentration of swollen
starch granule as described by Doublier et al. (1987). However, onset of viscosity
increase in potato starch was slightly affected up to 5% concentration after that it was
increased significantly by all tested polyols (Fig 3b). This might be attributed to the
differences in the influence of tested polyols on close packing of potato and wheat
starch granules due to their differences in swelling properties. After testing influence
of sucrose on wheat starch gelatinization and pasting properties DAppolonia (1972)
reported that the increased onset pasting, gelatinization temperatures and higher peak
viscosity were related to the inhibition of starch granule hydration by sucrose
molecules. Steeneken (1989) proposed that in dilute paste (8.9%), viscosity is
governed by the volume fraction of starch granules, and as the paste becomes more
concentrated it is controlled by particle rigidity. Interaction of sugar molecules with
starch chains as reported by Spies and Hoseney (1982) could increase the rigidity of
starch granules. This could permit granules to swell for a longer time achieving
higher peak viscosity before some granular breakdown occurs. Monosaccharides have

134

been reported to increase peak viscosity more than disaccharides (Bean and Osman,
1958). This was consistent with the results of this study for both starches. HP-CD
reduced peak viscosity of wheat starch. Disruption of amylose-lipid complex and
subsequent extensive hydration of wheat starch granules in the presence of HP-CD
could weaken the starch structure. Thus weaker granules can undergo greater shear
thinning at higher temperature. Potato is a higher swelling starch thus little increase of
swelling was caused by HP-CD at 1% concentration, which may influence on the
shear thinning at higher temperature.

The two starches showed an increased cold paste viscosity (CPV) in the presence of
all ployols but to a different extent. Sugars can create junction zones on amylose
chains facilitating realignment of amylose in cold paste. For wheat starch the increase
of CPV followed the order: glucose > sucrose > glycerol but sucrose was more
effective on potato starch followed by glucose and glycerol. This implies that the
conformational ordering and intermolecular association of starch polymers depend on
type of polyol as well as botanical source of the starch.

The observed increased gel hardness in wheat and potato starch in the presence of
sucrose, glucose and glycerol is consistent with the increased CPV. Textural
properties of wheat and potato starch gel (Table 4) were determined after 24 hr at 4
C. Thus textural properties should be predominantly affected by the short term
amylose gelation as this time is generally not enough for longer term amylopectin
association. Starch gels are metastable and non-equilibrium systems and therefore
undergo structural changes during storage (Ferrero et al., 1994). Miles et al. (1985)
and Ring et al. (1987) attributed the initial gel firmness during retrogradation to the
formation of an amylose matrix gel and the subsequent slow increase in gel firmness

135

to reversible crystallization of amylopectin. Morris (1990) explained that starch gel


properties relate to the characteristics of gel matrix, the amylose, the deformable
fillers (swollen granules) that are embedded on the continuous amylose matrix,
volume fraction of the filler, and the filler-matrix interaction. Doublier et al. (1987)
suggested the main structural parameters of a starch gel involved are the
deformability of swollen starch particles and the amylose concentration in the
continuous network. Increased gel hardness of wheat and potato starch by sucrose,
glucose, and glycerol could therefore be attributed to the formation of a strong
amylose gel matrix network via changes of conformational ordering and
intermolecular association of amylose chains probably by creating more junction
zones on amylose chains. Influence of polyols on the deformability (rigidity) of
swollen particles due to starch chain-polyol interaction could also affect the gel
hardness. Decreased gel hardness at higher concentration (40%) of sucrose and
glucose in wheat starch could therefore be attributed to the greater reduction of
amylose leaching that reduces the amylose concentration in the continuous network.
However, recent evidence in the literature argues that another mechanism may be at
play. Polysaccharides (agarose, -carrageenan, deacylated gellan) forming gels on the
basis of a double helix arrangement, as for amylose, are increasingly unable to
support intermolecular order in the presence of sugars at levels of 40% or beyond
(Kasapis and Al-Marhoobi, 2003). According to this argument, decreased gel
hardness could be due to reduced amylose order as opposed to reduced amylose
concentration in the network. Even at 40% concentration, glycerol did not decrease
the gel hardness of wheat starch because it did not reduce the amylose leaching and
probably it has no effect on the amylose order even at 40% concentration. However,
gel hardness of potato starch was not reduced despite amylose leaching was decreased
at higher concentration of sucrose, glucose, and glycerol indicating that creation of

136

more junction zones and the granular deformability (rigidity) of potato starch at
higher concentrations could override the negative effect of decreased amylose
leaching to form a harder gel structure. In the gelation of oxidized starch, Evageliou
et al. (2000) found that addition of sucrose, glucose and fructose increases the rate of
conformational ordering of oxidized starch polymers during cooling in the order of
sucrose > glucose > fructose but in subsequent holding at 5 C the increase of storage
module was least in the presence of sucrose followed by glucose and fructose,
indicating fructose caused rapid gelation and sucrose had least effect. Accordingly,
they suggested that sucrose is very effective in the rate of ordering but less effective
in network formation that required aggregation of polymers. This might be one of the
reasons for the observed order of gel hardness (glucose> sucrose> glycerol) in wheat
starch but this reasoning can not be applied for the outcome for potato starch.
However, these results suggest that the conformational ordering and intermolecular
association of starch chains not only depend on the types of sugar molecules but also
on the botanical source of the starch.

Although both amylose leaching and clod paste viscosity were dramatically increased
by HP-CD in wheat starch, the resulting gel was soft. Disruption of amylose-lipid
complex accompanied by the extensive swelling of starch granules may alter the
rigidity (the deformability) and close packing of the swollen particles that are
embedded in the amylose matrix. However, increased gel hardness of potato starch at
higher concentration of HP-CD implies that more interaction between HP-CD and
starch chains would preserve more granules from disintegration (Table 3) affecting
the shape and the deformability (rigidity) of the starch granules. Table 4 shows that
the other textural properties, especially in wheat starch, the adhesiveness, chewiness,
and gumminess also followed a similar trend as exhibited for gel hardness in the

137

presence of tested polyols. Except HP-CD all other tested polyols increased the
chewiness and formed more adhesive gel.

4.4. Conclusions
Polyhydroxy compounds could interact with starch molecules depending on size of
the molecule and the type of starch, altering thermal, pasting and gelling properties.
However, it seems that polyol-starch interaction is not the sole factor that determines
the gelatinization properties of starch in the presence of polyhydroxy compounds. Gel
hardness and other textural properties of starch paste, particularly in wheat starch
consistently increased up to 20% of sucrose and glucose despite decreasing amylose
leaching. This would suggest that in addition to the amylose concentration in the
continuous medium other factors such as rigidity of swollen granules, effectiveness of
the conformation ordering and intermolecular association of starch polymers affect
the textural properties of starch gel. Significant increase of amylose leaching, and
swelling factor and greater reduction of melting parameters of amylose-lipid complex
were apparent only in wheat starch in the presence of HP-CD, suggesting the
disruption of amylose-lipid complex and its influence on the thermal and pasting
properties were very significant. Further useful results would be provided by a
comparative study of different sugars with HP-CD of different degree of
modification on thermal and pasting properties of starch.

4.5. References
Ahmad, F.B., and Williams, P.A. 1999. Effect of sugars on the thermal and
rheological properties of sago starch. Biopolymers 50: 401-412.
Baek, M.H., Yoo, B., and Lim, S-T. 2004. Effects of sugars and sugar alcohols on the
thermal transition and cold stability of corn starch gel. Food Hydrocolloids 18: 133142.

138

Bean, M.L., and Osman, E.M. 1958. Behavior of starch during food preparation. II.
Effects of different sugars on the viscosity and gel strength of starch pastes. Food
Res. 24: 665-671.
Bean, M.M., and Yamazaki, T.Y. 1978. Wheat starch gelatinization in sugar solutions
I. sucrose: Microscopy and viscosity effects. Cereal Chem. 936-944.
Chinachoti, P., Kim-Shin, M.S., Mari, F., and Lo, L. 1991. Gelatinization of wheat
starch in the presence of sucrose and sodium chloride: correlation between
gelatinization temperature and water mobility as determined by oxygen-17 nuclear
magnetic resonance. Cereal Chem. 68: 245-248.
Chiotelli, E., Rolee, A., Meste, M.L. 2000. Effect of sucrose on the thermomechanical
behavior of concentrated wheat and waxy corn starch-water preparations. J. Agric.
Food Chem. 48: 1327- 1339.
Chrastil, J. 1987. Improved colorimetric determination of amylose in starches or
flours. Carbohydr. Res. 159: 367-389.
Chungcharoen, A., and Lund, D.B. 1987. Influence of solutes and water on rice starch
gelatinization. Cereal Chem. 64: 240-243.
DAppolonia, B.L. 1972. Effect of bread ingredients on starch gelatinization
properties as measured by the Amylograph. Cereal Chem. 49: 532-543.
Donovan, A.M. 1979. Phase transitions of the starch water system. Biopolymers 18:
263-275.
Doublier, J.L., Llamas, G., and Le Meur, M. 1987. A rheological investigation of
cereal starch pastes and gels. Effects of pasting procedures. Carbohydr. Polym. 7:
251-275.
Eliasson, A-C. 1992. A calorimetric investigation of the influence of sucrose on the
gelatinization of starch. Carbohydr. Polym. 18: 131-138.
Evageliou, V., Richardson, R.K., and Morris, E.R. 2000. Effect of sucrose, glucose,
and fructose on gelation of oxidized starch. Crabohydr. Polym. 42: 261-272.
Evans, I.D., and Haisman, D.R. 1982. The effect of solutes on the gelatinization
temperature range of potato starch. Starch 7: 224-231.
Ferrero, C., Martin, M.N., and Zantzky, N.E. 1994. Cornstarch-xanthan gum
interaction and its effect on the stability during storage of frozen gelatinized
suspension. Starch 46: 300-305.
Hoover, R., and Senanayake, N. 1996. Effect of sugars on the thermal and
retrogradation properties of oat starches. J. Food Biochem. 20: 65-83.
Kasapis, S., and Al-Marhoobi, I.M. 2003. Gelatin vs polysaccharide in mixture with
sugar. Biomacromolecules 4: 1142-1149.

139

Kohyama, K., and Nishinari, K. 1991. Effect of soluble sugars on gelatinization and
retrogradation of sweet potato starch. J. Agric. Food Chem. 39: 1406- 1410.
Levine, H., and Slade, L. 1988. Non equilibrium behavior of small carbohydrate
water system. Pure and Applied Chemistry 60: 1841-1847.
Maaurf, A.G., Che Man, Y.B., Asbi, B.A., Junainah, A.H., and Kennedy, J.F. 2001.
Gelatinization of sago starch in the presence of sucrose and sodium chloride as
assessed by differential scanning calorometry. Carbohydr. Polym. 45: 335-345.
Miles, M.J., Morris, V.J., Orford, P.D., and Ring, S.G. 1985. The roles of amylose
and amylopectin in the gelation and retrogradation of starch. Carbohydr. Res. 135:
271-281.
Morris, M.J. 1990. Starch gelation and retrogradation. Trends in Food Science &
Technology. 1: 2-6.
Perry, P.A., and Donald, A.M. 2002. The effect of sugars on the gelatinization of
starch. Carbohydr. Polym. 49: 155-165.
Richardson, G., Langton, M., Bark, A., and Hermansson, A-M. 2003. Wheat starch
gelatinization-the effect of sucrose, emulsifiers and the physical state of the emulsifier.
Starch 55: 150-161.
Ring, S.G., Colonna, P., Ianson, K.J., Kalicheversk, M.T., Miles, M.J., Morris, V.J.,
and Orford, P.D. 1987. The gelation and crystallization of amylopectin. Carbohydr.
Res. 162: 277-293.
Savage, H.L., and Osman, E.M., 1978. Effects of sugars and sugar alcohols on the
swelling of corn starch granules. Cereal Chem. 55: 477-454.
Sopade, P.A., Halley, P.J., and Junming, L.L. 2004. Gelatinization of starch in
mixtures of sugars. II. Application of differential scanning calorimetry. Carbohydr.
Polym. 58: 311-321.
Spies, R.D., and Hoseney, R.C. 1982. Effect of sugar on starch gelatinization. Cereal
Chem. 59: 128-131.
Steeneken, P.A.M. 1989. Rheological properties of aqueous suspensions of swollen
starch granules. Carbohydr. Polym. 11: 23-42.
Tester, R., and Morrison, W.R. 1990a. Swelling and gelatinization of cereal starches:
I. Effects of amylopectin, amylose, and lipids. Cereal Chem. 67: 551-557.
Watase, M., Kohyama, K., and Nishinari, K. 1992. Effects of sugars and polyols on
the gel-sol transition of agarose by differential scanning calorimetry. Thermochemica
Acta 206: 163-173.
Wootton, M., and Bamunuarachchi, A. 1980. Application of differential scanning
calorimetry to starch gelatinization III. Effects of sucrose and sodium chloride. Starch
32: 126-129.

140

Table 1. Gelatinization temperature (Tp) and enthalpy (H) of wheat and potato starches in the presence of sucrose, glucose, glycerol, and HP-CD
Starch Concentration
Sucrose
Glucose
Glycerol
HP-CD
(%)
Tp (C)
H (J/g)
Tp (C)
H (J/g)
Tp (C)
H (J/g)
Tp (C)
H (J/g)
Wheat

Potato

0
1
5
10
20
40

59.20.2
60.80.1
62.10.3
62.40.4
66.10.1
76.40.3

11.00.4
11.10.2
11.20.1
11.50.3
11.80.5
12.40.2

59.20.5
59.30.1
60.20.3
61.20.2
63.70.3
71.00.1

0
62.00.4
14.60.1
62.00.4
1
62.80.5
14.70.2
62.20.4
5
64.50.1
15.00.1
62.60.3
10
65.70.4
15.20.4
64.30.2
20
67.50.6
15.40.1
66.60.1
40
76.50.3
15.90.3
72.50.3
Values are means of triplicate determinations standard deviation.

11.00.2
11.00.2
11.10.1
11.30.3
11.50.4
12.00.5

59.20.2
59.40.1
60.10.1
61.20.2
63.40.6
68.50.3

11.00.5
11.00.1
11.20.3
11.30.4
11.50.4
11.80.3

59.20.2
59.40.3
59.60.1
62.70.1
66.50.3
76.20.5

11.00.2
8.50.3
8.2 0.4
6.30.2
6.20.3
5.80.1

14.60.2
14.70.1
14.90.3
15.00.4
15.20.3
15.50.2

62.00.4
62.10.1
62.90.4
63.80.5
65.90.3
70.80.2

14.60.6
14.60.5
14.60.3
14.90.2
15.00.1
15.20.2

62.00.1
62.80.1
64.70.4
66.50.1
68.30.1
75.30.3

14.60.3
14.70.3
14.90.1
15.30.4
15.50.1
16.80.3

141

Table 2. Melting temperature (Tm) and enthalpy (H) of amylose-lipid complex


of wheat starch in the presence of sucrose, glucose, glycerol, and HP-CD
Treatment
Concentration
Tm (C)
H (J/g)
(%)
Sucrose
0
107.60.8
1.60.1
1
107.70.2
1.60.2
5
107.90.5
1.60.1
10
109.20.3
1.40.3
20
112.20.7
1.40.1
40
121.70.2
1.30.1
Glucose

1
5
10
20
40

107.70.2
108.10.1
109.10.3
111.20.1
118.10.1

1.60.2
1.60.1
1.50.2
1.50.1
1.30.1

Glycerol

1
5
10
20
40

107.50.4
107.90.5
108.30.2
110.50.3
112.80.4

1.60.1
1.60.3
1.60.3
1.40.2
1.40.1

HP-CD

1
5
10
20
40

106.70.2
105.80.1
100.60.3
ND
ND

0.90.2
0.60.3
0.30.4
ND
ND

ND= not determined.


Values are means of triplicate determinations standard deviation.

142

Table 3. Pasting properties of wheat and potato starches in the presence of sucrose,
glucose, glycerol, and HP-CD
Starch
Treatment Con. (%)
PV
HPV
BD
CPV
SB
Wheat
Control
0
150
112
38
212
100
Sucrose
1
160
118
41
219
101
5
209
147
60
258
110
10
268
183
85
305
120
20
365
284
78
425
140
40
438
422
16
530
112
Glucose
1
160
116
43
221
103
5
210
140
70
256
116
10
273
168
105
297
128
20
380
276
104
428
150
40
526
502
24
655
153
Glycerol
1
162
117
45
221
102
5
200
133
67
250
116
10
249
151
98
275
124
20
325
197
127
319
122
40
430
369
60
450
75
HP-CD
1
134
98
35
220
125
5
120
104
17
235
130
10
132
119
12
275
158
20
138
134
4
340
206
40
ND
ND
ND
ND
ND
Potato
Control
0
678
190
489
270
74
Sucrose
1
680
193
484
275
79
5
704
211
490
287
75
10
740
230
506
310
80
20
773
284
488
389
100
40
ND
ND
ND
ND
ND
Glucose
1
685
194
490
269
73
5
707
198
509
275
76
10
749
215
533
296
80
20
780
279
499
376
98
40
ND
ND
ND
ND
ND
Glycerol
1
680
194
488
270
75
5
703
210
492
285
76
10
722
230
494
305
78
20
760
286
475
372
84
40
ND
ND
ND
ND
ND
HP-CD
1
669
185
480
266
91
5
637
191
446
267
75
10
636
199
437
275
76
20
634
227
409
333
106
40
558
314
243
416
102
PV=peak viscosity; HPV=hot paste viscosity; BD=breakdown; CPV=cold paste
viscosity; SB=setback.
ND=not determined as the consistent of the profile was not constant.
Values are means of at least duplicate determinations.

143

Table 4. Textural properties of wheat and potato starch gel prepared with sucrose, glucose, glycerol, and HP-CD
Starch
Con.
Sucrose
Glucose
Glycerol
HP-CD
(w/v)
HD AD GU
CH
HD
AD
GU
CH
HD
AD GU CH
HD AD
Wheat
0
77
197 32
30
77
197
32
30
77
197 32
30
77 197
1
80
200 34
37
80
164
31
30
71
218 26
23
42 162
5
82
113 40
32
85
181
35
34
78
227 32
30
48 176
10
118 182 43
74
124 227
37
35
90
242 36
35
56 184
20
146 280 58
71
155 300
60
75
103 230 40
40
63 185
40
78
150 35
62
146 109
60
61
138 166 48
86
ND ND
Potato

0
38
1
19
19
38
1
19
1
38
5
20
18
35
3
19
5
39
5
20
20
38
5
20
10
53
6
25
35
45
14
22
20
67
8
31
30
54
30
27
40
86
12
48
49
70
38
34
HD=hardness; AD=adhesiveness; GU=gumminess; CH=chewiness.
ND=not determine.
Values are mean of at least duplicate determinations.

19
25
20
24
27
32

38
42
44
44
45
51

1
22
55
95
102
150

19
18
18
18
20
26

19
18
18
18
19
24

38
34
27
30
45
53

1
12
15
29
164
167

GU
32
12
15
18
21
ND

CH
30
14
17
18
19
ND

19
14
16
11
21
24

19
13
15
10
19
22

144

60

60

50

Swelling factor

40

sucrose
glucose

30

glycerol
HP-CD

20
10

40

sucrose
glucose

30

glycerol
HP-CD

20
10

10

20

30

40

50

10

20

30

40

50

16

25

A
Amylose leaching (%)

50

14

20

12
sucrose

15

glucose
glycerol
10

HP-CD

10

sucrose
glucose

glycerol

HP-CD

2
0

0
0

10

20

30

40

50

10

20

30

40

50

Concentration (W/V)
Fig. 1. Swelling factor and amylose leaching of wheat (A) and potato (B) starches at different concentrations of sucrose, glucose, glycerol,
and HP-CD.

145

0.0

-0 .3

-0 .4

-0.2

-0 .5

-0 .6
-0.4

Heat Flow (W/g)

-0 .7

-0 .8

0.0

40

B
60

80

100

-0.6

12 0

-0.3

40

60

80

100

120

-0.4

-0.2
-0.5

-0.6

-0.4

-0.7

C
-0.6

D
-0.8

40

60

80

100

120

40

60

80

100

120

Temperature (C)
Fig. 2a. DSC curves of wheat starch in the presence of 0%, 5%, 10%, 20%, and 40% sucrose (A), glucose (B), glycerol (C), and HPCD (D); curves represent increasing concentration from top to bottom.

146

-0 .3

-0.1

-0 .4
-0.2

-0 .5
-0.3

-0 .6
-0.4

-0 .7

Heat Flow (W/g)

-0 .8

-0 .9

40

60

80

1 00

12 0

-0.6

-0.2

-0 .3

-0.3

-0 .4

-0.4

-0 .5

-0.5

-0 .6

-0.6

-0.7

50

40

-0 .7

C
40

-0.5

60

70

80

90

100

110

-0 .8

50

60

70

80

90

100

110

D
40

50

60

70

80

90

100

110

Temperature (C)
Fig. 2b. DSC curves of potato starch in the presence of 0%, 5%, 10%, 20%, and 40% sucrose (A), glucose (B), glycerol (C), and HP-CD
(D); curves represent increasing concentration from top to bottom.

147

100
600

450

100

B
)

80

E
450

60

300

80

60

300
40

150

40

20

Viscosity (RVU)

D
0

10

B
15

20

25

0
0

10

15

20

25

100

100
320

400

Temperature (C)

150
20

80

80
240

300

60

60

E
160
200

40

40

E
80

100

20

0
0

10

15

20

25

10

20

A
15

20

25

Time (Mins)

Fig. 3a. RVA curves of wheat starch in different concentrations of sucrose (A), glucose (B), glycerol (C), and HP-CD (D); A=native,
B=1%, C=5%, D=10%, E=20%, F=40%.

148

100

100

80

60

D
400

80

600

600

60

400
40

200

Viscosity (RVU)

20

A
0

0
0

10

15

20

25

0
0

10

15

20

25

100

750

80

600

60

450

60

40

300

40

20

600

100

80

Temperature (C)

200

40

D
E
400

200
20

0
5

10

20

0
0

150

15

20

25

0
0

10

15

20

25

Time (Mins)
Fig. 3b. RVA curves of potato starch in different concentrations of sucrose (A), glucose (B), glycerol (C), and HP-CD (D); A=native, B=1%,
C=5%, D=10%, E=20%, F=40%.
149

Chapter 5
Functional Properties and Retrogradation of Heat-Moisture Treated
Wheat and Potato Starches in the Presence of Hydroxypropyl Cyclodextrin
Abstract
Some functional and retrogradation properties of native and heat-moisture treated
potato and wheat starches were examined in the presence of hydroxypropyl cyclodextrin (HP-CD). HP-CD increased swelling factor, amylose leaching, and
solubility of both native and heat-moisture treated wheat starches but it had less
impact on corresponding potato starches. Gelatinization enthalpy of native wheat
starch was decreased in the presence of HP-CD but it was increased in potato starch
with increasing the concentration. Reduction of amylose-lipid complex endotherm in
both native and heat-moisture treated wheat starch was noticed in the presence of
HP-CD. Heat-moisture treatment did not change the transition parameters of
amylose-lipid complex showing its resistance to hydrothermal treatment. HP-CD
greatly decreased the pasting temperature of wheat starch. Cold paste viscosity of
both native and heat-moisture treated wheat starch was increased by HP-CD to a
greater extent than corresponding potato starch. Amylopectin retrogradation of all the
starches was unaffected in the presence of HP-CD but heat-moisture treatment
slightly decreased retrogradation of potato starch. These results suggest that HP-CD
can disrupt the amylose-lipid complex within the starch granule in both native and
heat-moisture treated wheat starch but has no influence on amylopectin retrogradation.
However, greatly increased wheat starch setback with HP-CD indicates its greater
effect on wheat starch amylose retrogradation.

150

5.1. Introduction
Various additives are used for different purposes in starch-related food products such
as stabilizers, sweeteners, and textural modifying agents. There is growing interest in
searching for new functional additives that are readily available and non-toxic. cyclodextrin, a cyclic oligosaccharide which has a hydrophobic core in the center of
the molecule is widely used in the pharmaceutical industry because of its ability to
form complexes with various organic molecules. However, its applications in the
starch-related food industry have not been extensively investigated although it has
potential as a flavor carrier and as a possible non-toxic starch modification agent. cyclodextrin can complex with native starch lipids altering the starch structure and
functional properties.

Hydroxypropyl -cyclodextrin is a modified version of -CD. Substitution of


hydroxypropyl groups onto the hydroxyl groups of -cyclodextrin increases its
solubility in water. The key to utilization of cyclodextrin in starch industry is the
understanding of its interaction with starch molecules and the subsequent effects on
starch functional properties. Very little research on cyclodextrin-starch interaction has
been conducted, and has been limited to native starches. My previous work on -CD
and HP-CD-starch interaction has shown that both cyclodxtrins can promote
amylose leaching, and increase swelling factor and solubility by disrupting the
amylose-lipid complex.

Heat-moisture treatment is a physical modification method that can be employed to


modify starch structure and functional properties (Hoover and Vasanthan, 1994;
Gunaratne and Hoover, 2002; Stute, 1992; Lorenz and Kulp, 1982). Heat-moisture
treatment of potato and wheat starches decreases the apparent amylose content,

151

indicating additional interaction between native starch lipid and amylose chains
(Hoover and Vasanthan, 1994), and my previous results showed that HP-CD is
capable of defatting starch granules by complexing with starch lipids. Interaction of
starch chains within the amorphous region and reorientation and disruption of starch
crystallites during heat-moisture treatment could have some specific interaction with
the larger HP-CD molecule affecting the starch structure and functional properties.
The aim of this study was to examine starch properties of both native and heatmoisture treated (structurally altered) starches in the presence of HP-CD. This may
reveal insights into -CD- starch interaction, and show how heat-moisture treatment
affects the -CD starch interaction.

5.2. Materials and Methods

5.2.1. Materials
Wheat and potato starch, and hydroxypropyl -cyclodextrin (HP-CD), were
purchased from Sigma Chemical Co., (St. Louis, MO).

5.2.2. Methods

Amylose leaching
Distilled water (10 mL) was added to starch (20 mg, db) in a screw cap tube. Tubes
were then heated at different temperatures (50-100 C) for 30 min. After cooling to
ambient temperature, samples were centrifuged at 2000 g for 10 min. Amylose
content of supernatant (0.1 mL) was estimated as described by Chrastil (1987).

152

Swelling factor
Swelling factor, the ratio of the volume of swollen starch granules to the volume of
dry starch was determined by the method of Tester and Morrison (1990a), when
starch (50 mg, db) was heated at different temperatures (50-100 C) in 5 mL water.

Solubility
The method of Subramanian et al. (1994) with minor modification was used to
determine the solubility of starch. Starch (0.5 g) was heated at different temperatures
(50-100 C) in water (15 mL) with continuous stirring to prevent lump formation. The
slurry was then centrifuged at 3,000 g for 10 min. A known aliquot of supernatant
was dried at 130 C overnight. The weight of oven-dried supernatant was back
calculated to the volume of supernatant and the initial weight of dry starch and
expressed as percent soluble starch.

Differential scanning calorimetry


Gelatinization and dissociation parameters of amylose-lipid complex were measured
using a TA 2920 Modulated DSC Thermal Analyzer differential scanning calorimeter
equipped with a thermal analysis data station (TA Instruments, Newcastle, DE).
Starch (3 mg) was directly measured onto the aluminum DSC pan and distilled water
(3 L) was added with a microsyringe and mixed for homogenization. Pans were
sealed, and allowed to stand for 1 hr at room temperature for even distribution of
water. The scanning temperature and the heating rates were 30-140 C and 5 C/min
respectively. An empty pan was used as reference for all measurements.

Pasting properties

153

Pasting properties of starches were determined using a Rapid Visco-Analyser (RVA)


model 3D (Newport Scientific, Warriewood, Australia). Distilled water (25.5 g) was
added to starch (2.5 g, db) in the RVA canister to obtain a total constant sample
weight of 28 g. The slurry was then manually homogenized using the plastic paddle
to avoid lump formation before the RVA run. A programmed heating and cooling
cycle was set for 22 minutes, where it was first held at 50 C for 1.0 min, heated to 95
C in 7.5 min, further held at 95 C for 5 min, cooled to 50 C within 7.5 min and
held at 50 C for 1 min.
Retrogradation
Starch gel was prepared in the DSC pan using 1:1 starch to water ratio scanning the
sample to 120 C. Sample was then stored at 4 C for 24 hr to initiate nucleation.
After that samples were kept at 35 C for 10 days before rescanning. Temperature
range and heating rate were 30-120 C and 10 C/min respectively.

Heat-moisture treatment
Moisture content of starch samples (20 g) was brought up to 30 % in tightly capped
bottles. Sample was then kept at room temperature for 24 hr for equal distribution of
moisture content before heating at 100 C for 10 hr.
Note: Hydroxypropyl -cyclodextrin (MS=0.8) 1%, 5% or 10% (W/V) in dw was
added instead of water in all above experiments where appropriate.

5.3. Results and Discussion

Swelling factor

154

Heat-moisture treatment decreased swelling ability of both wheat and potato starches,
but to a greater extent in potato starch (Table 1). Tester and Morrison (1990a)
claimed that swelling is primarily a property of amylopectin and that amylose is a
diluent. Cooke and Gidley (1992) have suggested that the forces holding the granule
together are mainly at the double helical level and that the starch crystallinity
functions as a means of achieving dense packing rather than as a primary provider of
structure. Thus the unraveling of amylopectin double helices during heat-moisture
treatment could destabilize the starch structure weakening the swelling ability. Also
additional starch chain interaction within the amorphous region during heat-moisture
treatment may restrict the swelling ability of the starch granule.

Swelling factor of both native and heat-moisture treated wheat starches was greatly
increased

in the presence of HP-CD but

it was slightly increased

in the

corresponding potato starches, showing that HP-CD has less impact on potato starch
swelling (Table 1). Increased swelling in the presence of HP-CD only in wheat
starch even after heat-moisture treatment shows that HP-CD swells wheat starch
granules in a specific manner based on a factor not common to both starches. This
result further confirmed my earlier finding that HP-CD increased swelling only in
cereal starches by disrupting the amylose-lipid complex.

Amylose leaching
Amylose leaching of native and heat-moisture treated potato and wheat starches at
different temperatures in the presence of HP-CD is shown (Fig. 1). Addition of HPCD dramatically increased the extent of amylose leaching of both native and heatmoisture treated wheat starches, but corresponding potato starches showed little or no
effect in the presence of HP-CD. In the case of native potato starch, a slight decrease

155

of amylose leaching was noted when HP-CD concentration was increased above 1%.
After heat-moisture treatment amylose leaching of both starches was decreased
probably due to the additional starch polymer interaction during heat-moisture
treatment but the effect was greater in potato starch. Addition of HP-CD greatly
increased the amylose leaching of heat-moisture treated wheat starch, to a greater
extent similar to native starch. Previous results of this study showed that disruption of
amylose-lipid complex by HP-CD facilitates more amylose to come out from the
granules. This claim is further supported by the observed results of higher amylose
leaching before and after heat-moisture treatment of wheat starch in the presence of
HP-CD. Lesser impact on amylose leaching of both native and heat-moisture treated
potato starch in the presence of HP-CD is therefore an expected result because
virtually lipid-free potato starch has weaker interaction with HP-CD. Amylose
leaching was slightly reduced in potato starch above 1% HP-CD. Greater water
affinity towards the HP-CD molecule could limit the available water in the system.
This could restrict the mobility of amylose chains and also it was possible for HPCD-starch chain interaction to reduce the amylose leaching. Higher solubility of
wheat starch compared with corresponding potato starch in the presence of HP-CD
is consistent with increased amylose leaching and swelling factor (Table 1).

Gelatinization
Gelatinization temperature was increased in all the starches in the presence of HPCD (Table 2). The higher the concentration of HP-CD, the more the gelatinization
temperature increased. HP-CD reduced gelatinization enthalpy (H) of native wheat
starch but did not affect it in HMT wheat starch. HP-CD increased H of native
potato starch with increasing the concentration, but it had no effect on HMT potato
starch. Gelatinization is primarily a swelling driven process and thus decreased wheat

156

starch gelatinization enthalpy is consistent with the increased swelling in the presence
of HP-CD. Increased gelatinization enthalpy of native potato starch in the presence
of HP-CD therefore could relate to the reduced swelling by HP-CD. However, after
heat-moisture treatment gelatinization enthalpy of wheat starch was unaffected by
HP-CD, although HP-CD increased the swelling ability of heat-moisture treated
wheat starch. This could be due to some structural changes occurring in the
amorphous and crystalline region during heat-moisture treatment. Increased
gelatinization temperature of native wheat starch in the presence of HP-CD is not
consistent with the increased swelling. Generally, greater granular swelling should
reduce the gelatinization temperature. This could suggest that HP-CD has a unique
interaction with wheat starch. It has been proposed that polyhydric compounds such
as sugars increased gelatinization temperature as a result of competition between
sugar and starch for water lowering water availability in the system for gelatinization
(DAppolonia, 1972), sugar-starch interaction (Spies and Hoseney, 1982) or
decreased plasticizing effect of sugar (Perry and Donald, 2002).

Heat-moisture treatment increased gelatinization temperature while decreasing


gelatinization enthalpy. Hoover and Vasanthan (1994) and Gunaratne and Hoover
(2002) explained that decreased swelling after heat-moisture treatment could decrease
the destabilizing effect of the amorphous region on the crystalline domain requiring
more energy for melting. Lim et al. (2001) reported that the increased transition
temperature after heat-moisture treatment could be attributed to the possible
annealing of starch crystalline region during heat-moisture treatment. Weaker
amylopectin crystals could degrade during heat-moisture treatment leaving thermally
more stable crystals that melt at higher temperature (Miyoshi 2002). The decrease in
enthalpy on heat-moisture treatment suggests that some of the double helices present

157

in crystalline and in non-crystalline regions of the granule may have disrupted under
the conditions prevailing during heat-moisture treatment. Thus, fewer double helices
would unravel and melt during gelatinization of heat-moisture treated starches. A
greater reduction of gelatinization transition enthalpy (Table 2) of heat-moisture
treated potato starch indicates that potato starch double helices are more susceptible
to heat-moisture treatment than those of wheat starch. Similar results were reported
for heat-moisture treated potato starch (Gunaratne and Hoover 2002; Hoover and
Vasanthan, 1994). In B-type starches (potato), the packing of helices is less compact
than in A-type starches (cereal starches) (Gidley, 1987). Therefore double helices of
potato starch would be more mobile and more prone to disruption than those in wheat
starch. The double endotherm pattern observed in gelatinization at moderate water
content (50%) was more distinguished after heat-moisture treatment particularly in
wheat starch (Fig. 2).

Heat-moisture treatment apparently did not change the transition parameters of


amylose-lipid complex (Table 2 and Fig. 2). This is in agreement with Donovan et al.,
(1983). However, amylose-lipid complex of both native and heat-moisture treated
wheat starches greatly decreased in the presence of HP-CD with the increase of
concentration (~ 80% at 10% concentration), indicating disruption of amylose-lipid
complex by HP-CD (Table 3 and Fig. 2). Transition temperatures of amylose-lipid
complex also decreased as the concentration of HP-CD increased.

Pasting properties
The pasting profile of potato starch showed a typical A-type profile with higher peak
viscosity followed by rapid breakdown at higher holding temperature while wheat
starch exhibited B-type pasting profile characterized by lower peak and much less

158

thinning (Schoch and Maywald, 1968). After heat-moisture treatment the original
pasting profile of potato starch was much more changed than wheat starch shifting
more close to a B-type profile (Fig. 3). After heat-moisture treatment both starches
began to paste at a higher temperature and had lower subsequent viscosities than the
corresponding unmodified starches. This was more in potato starch. This could be
attributed to the decreased swelling ability of heat-moisture treated starch granules.
Similar observations have been reported for potato starch (Kulp and Lorenz, 1981;
Stute, 1992) and wheat starch (Kulp and Lorenz, 1981). Addition of HP-CD to
native and heat-moisture treated wheat starches decreased the pasting onset
temperature probably due to the occurrence of early swelling in the presence of HPCD. Heat-moisture treated starches showed a greater resistance to shear thinning.
Potato starch showed no granular breakdown after heat-moisture treatment. Granular
rigidity resulted from heat-moisture treatment due to the extensive starch chain
interaction which could impart resistance to shear thinning. Cold paste viscosity and
set back which are primarily related to the reassociation of amylose chains during the
cooling stage were progressively increased with increase of concentration only in
wheat starch while potato starch showed unchanged or decreased cold paste viscosity
and setback. As shown in previous results, HP-CD could remove lipid from amylose
polymer. Lipid free amylose can easily reassociate (retrograde), in contrast to lipid
complexed amylose. Perhaps HP-CD can create more junction zones on amylose
polymer facilitating the realignment of amylose. There is a positive correlation
between severity of amylose-lipid complex disruption and the concentration of HPCD and cold paste viscosity development in both native and heat-moisture treated
wheat starches (Table 1, 3, and 4). Thus it seems the amylose-lipid complex that can
reform on cooling can possibly affect properties of the starch gel but further research
is needed to clarify this claim.

159

Retrogradation
The behavior of gelatinized starches on cooling and storage is generally termed
retrogradation. This phenomenon is of great interest to food scientists and
technologists since it profoundly affects the quality of starch-based foods. Short- term
amylose association is far stronger than long-term amylopectin association.
Retrograded amylose crystals dissociate at higher temperature compared to
retrograded amylopectin crystals thus DSC can not be applied to measure amylose
retrogradation until special conditions are employed to protect the DSC pan from
bursting at higher temperature (Gunaratne and Corke, 2004). Usually retrogradation
causes undesirable characteristics in starch-related food products.

My aim of this study was to detect whether the hydrophilic molecule of HP-CD has
an affect on amylopectin retrogradation. Unchanged melting enthalpy of both
retrograded native and heat-moisture treated wheat starch crystals in the presence of
HP-CD suggests that neither heat-moisture treatment nor HP-CD affect the
retrogradation of wheat starch. However, heat-moisture treatment slightly decreased
potato starch retrogradation but it was unaffected in the presence of HP-CD similar
to wheat starch. During heat-moisture treatment possible degradation of longer outer
branches of amylopectin in potato starch compared to that of shorter branches of
wheat starch could reduce the average unit chain length of amylopectin weakening
the lateral association of amylopectin.

The differences of To, Tp and Tc of retrograded starch crystals

compared with

corresponding native starch crystals (Table 2 and 5) indicate that the nature (quality
and perfection of crystals) of retrograded starch crystals is different from that of
native amylopectin starch crystals. Eliasson (2003) reported that melting temperature

160

range of recrystallized amylopectin is rather similar to the gelatinization temperature


range. However, in this study recrystallized amylopectin (first stored at 4 C and then
at 40 C), melted over a narrow temperature range compared with corresponding
native starch gelatinization transition. Quality of retrograded amylopectin crystallites
can be improved by storing first at low temperature (6 C) followed by higher
temperature (30-40 C). These crystals melted at higher temperature over a narrow
range of temperature as they are more homogenized. This could be the reason for
observed

differences

in

melting

temperatures

between

gelatinization

and

corresponding melting temperatures of recrystallized amylopectin.

5.4. Conclusions
Greater increase of granular swelling, amylose leaching and solubility of native and
heat-moisture treated wheat starches in the presence of HP-CD is consistent with the
disruption of amylose-lipid complex. Heat-moisture treatment does not affect the
transition parameters of amylose-lipid complex but presence of HP-CD drastically
decreases transition parameters of amylose-lipid complex in both native and heatmoisture treated starches. This further proved the disruption of amylose-lipid
complex by HP-CD within the starch granules. HP-CD has no effect on starch
retrogradation but heat-moisture treatment slightly decreased retrogradation of potato
starch. Cold paste viscosity of wheat starch dramatically increased as the
concentration of HP-CD increased indicating its influence on wheat starch amylose
retrogradation. For further explanation of cyclodextrin-starch interaction it is worth
examining the thermal and pasting properties of chemically modified starch such as
acetylated starch in the presence of cyclodextrin. Perhaps there could be a specific
interaction between hydrophilic HP-CD molecules with substituted acetyl groups
creating novel starch properties.
161

5.5. References
Chrastil, J. 1987. Improved colorimetric determination of amylose in starches or
flours. Carbohydr. Res. 159: 154-158.
Cooke, D., and Gidley, M.J. 1992. Loss of crystallinity and molecular order during
starch gelatinization: origin of the enthalpic transition. Carbohydr. Res. 227: 103-112.
D Appolonia, B.L. 1972. Effect of bread ingredients on starch gelatinization
properties as measured by the amylograph. Cereal Chem. 49: 532-543.
Donovan, J.W., Lorenz, K., and Kulp, K. 1983. Differential scanning calorimetry of
heat-moisture treated wheat and potato starches. Cereal Chem. 60: 381-387.
Eliasson, A.C. 2003. Utilization of thermal properties for understanding baking and
salting processes. In: Characterization of Cereals and Flours: Properties, Analysis,
and Applications. G. Kaletunc and K. Breslauer, (Eds.), pp. 65-115. Marcel Dekker,
New York.
Gidley, M.J. 1987. Factors affecting the crystalline type (A-C) of native starches and
model compounds. A rationalization of observed effects in terms of polymorphic
structure. Carbohydr. Res. 161: 301-304.
Gunaratne, A., and Corke, H. 2004. Starch: analysis of quality. In: Encyclopedia of
Grain Science. C.W. Wrigley, H. Corke, and C.E. Walker, (Eds.), p. 202-212,
Elsevier, Oxford,
Gunaratne, A., and Hoover, R. 2002. Effect of heat-moisture treatment on the
structure and physicochemical properties of tuber and root starches. Carbohydr.
Polym. 49: 425-437.
Hoover, R., and Vasanthan, T. 1994. Effect of heat moisture treatment on the
structure and physicochemical properties of cereal, legume, and tuber starches.
Carbohydr. Res. 252: 33-53.
Kulp, K., Lorenz, K. 1981. Heat-moisture treatment of starches. 1. Physicochemical
properties. Cereal Chem. 58: 46-48.
Lim, S.-T., Chang, E.-H., and Chung, H.-J. 2001. Thermal transition characteristics of
heat-moisture treated corn and potato starches. Carbohydr. Polym. 46: 107Lorenz, K., Kulp, K. 1982. Cereal and root starch modification by heat-moisture
treatment. 1. Physicochemical properties. Starch 34: 50-54.
Miyoshi, E. 2002. Effect of heat-moisture treatment and lipids on gelatinization and
retrogradation of maize and potato starches. Cereal Chem. 79: 72-77.
Perry, P.A., and Donald, A.M. 2002. The effect of sugars on the gelatinization of
starch. Carbohydr. Polym. 49: 155-165.

162

Schoch, T.J., and Maywald, E.C. 1968. Preparation and properties of various legume
starches. Cereal Chem. 45: 564-573.
Spies, R.D., Hoseney, R.C. 1982. Effect of sugar on starch retrogradation. Cereal
Chem. 59: 128-132.
Stute, R. 1992. Hydrothermal modification of starches: the difference between
annealing and heat-moisture treatment. Starch 44: 205-514.
Subramanian, V., Hoseney, R.C. and Bramel-Cox, P. 1994. Shear thinning properties
of sorghum and corn starches. Cereal Chem. 71: 272-275.
Tester, R., and Morrison, W.M. 1990a. Swelling and gelatinization of cereal starches.
1. Effects of amylopectin, amylose, and lipids. Cereal Chem. 67: 551-557.

163

Table 1. Swelling factor and solubility of native and heat-moisture treated wheat
and potato starches in the presence of HP-CD
Source
Treatment
Swelling factor
Solubility
(80 C)
( 80 C)
Wheat
Native+water
15.20.2
5.30.1
Native+1% HP-CD
53.10.1
16.20.3
Native+5% HP-CD
42.60.3
20.10.4
Native+10% HP-CD
41.20.1
19.80.2

Potato

HMT+water
HMT+1% HP-CD
HMT+5% HP-CD
HMT+10% HP-CD

12.10.2
22.30.1
24.10.2
38.20.3

3.20.3
10.20.2
13.80.1
14.10.3

Native+water
Native+1% HP-CD
Native+5% HP-CD
Native+10% HP-CD

47.80.1
49.10.4
40.10.2
37.80.3

9.30.3
11.50.2
12.20.3
14.10.1

HMT+water
7.60.1
1.30.1
HMT+1% HP-CD
8.90.2
2.70.3
HMT+5% HP-CD
10.50.3
4.30.3
HMT+10 HP-CD
9.40.1
4.20.1
HMT=heat-moisture treated.
Values are means of triplicate determinations standard deviation.

164

Table 2. Gelatinization parameters of heat-moisture treated wheat and potato starches in the presence of HP-CD
Starch
Treatment
To (C)
Tp (C)
Tc (C)
Tc-To (C)
Wheat
Native+water
54.90.2
59.20.1
86.90.1
32.30.4
Native+1% HP-CD
55.00.1
59.20.1
81.20.2
26.40.3
Native+5% HP-CD
55.80.2
59.90.2
82.60.3
26.50.2
Native+10% HP-CD 57.80.1
62.20.1
83.30.1
25.60.5

H (J/g)
11.00.1
8.50.1
8.20.2
6.50.1

HMT+water
HMT+1% HP-CD
HMT+5% HP-CD
HMT+10% HP-CD

67.50.2
68.10.1
68.50.3
71.00.1

71.50.2
71.70.2
72.00.1
74.40.1

88.90.1
88.80.2
90.10.2
93.40.1

21.20.3
20.50.2
21.40.2
22.60.1

6.20.3
6.40.1
6.30.1
6.30.1

Native+water
Native+1% HP-CD
Native+5% HP-CD
Native+10% HP-CD

58.20.2
58.20.3
59.80.1
60.70.1

62.00.1
62.80.2
64.20.1
66.50.2

80.30.1
80.10.1
81.10.1
81.20.1

22.10.5
21.70.3
21.50.4
20.30.1

13.60.1
13.80.2
14.00.1
14.70.2

HMT+water
71.10.1
75.50.1
HMT+1% HP-CD
71.40.1
75.90.1
HMT+5% HP-CD
72.80.1
78.20.2
HMT+10% HP-CD
74.50.2
79.70.1
To=onset, Tp=peak, Tc=conclusion, H=gelatinization enthalpy.
HMT=heat-moisture treatment.
Values are means of triplicate determinations standard deviation.

94.00.1
94.30.2
95.30.1
96.80.1

22.70.2
22.90.5
22.40.1
22.60.2

5.70.2
6.10.1
6.30.1
6.20.1

Potato

165

Table 3. Dissociation parameters of amylose-lipid complexes of native and heatmoisture treated wheat starches in the presence of HP-CD
Starch Treatment
To (C)
Tp (C)
Tc (C)
H (J/g)
Wheat Native+water

103.30.1 107.60.3

112.90.3

1.50.1

Native+1% HP-CD

101.20.2 106.00.1

111.20.1

0.90.2

Native+5% HP-CD

101.50.3 105.40.1

111.00.1

0.60.2

Native+10% HP-CD

95.70.2

100.60.2

105.60.2

0.30.1

HMT+water

103.10.1 107.30.3

112.00.2

1.60.1

HMT+1% HP-CD

100.00.3 106.00.2

111.50.1

0.90.1

HMT+5% HP-CD

100.00.1 106.10.1

111.20.1

0.50.2

HMT+10% HP-CD 96.10.2 102.10.3 109.10.1 0.40.1


Melting parameters of amylose-lipid complex: To= onset, Tc=conclusion,
H=transition energy.
HMT=heat-moisture treatment.
Values are means of triplicate determinations standard deviation.

166

Table 4. Pasting parameters of native and heat-moisture treated wheat and potato starches in the presence of HP-CD
Starch
Treatment
PV
HPV
BD
CPV
Wheat
Native+water
1480.5
1110.4
370.4
2080.7
Native+1%HP-CD
1340.6
980.5
350.4
2241.4
Native+5% HP-CD
1200.4
1041.3
170.6
2350.5
Native+10% HP-CD
1390.6
1240.6
150.5
2900.7

Potato

SB
960.3
1251.1
1300.6
1661.3

HMT+water
HMT+1% HP-CD
HMT+5% HP-CD
HMT+10% HP-CD

910.7
690.8
780.4
880.8

760.6
660.5
770.5
870.7

140.8
30.1
10.3
10.1

1130.4
1210.3
1450.6
1690.7

360.7
570.4
670.4
820.8

Native+water
Native+1% HP-CD
Native+5% HP-CD
Native+10% HP-CD

6821.4
6691.6
6370.9
6370.5

1910.5
1850.7
1910.8
1990.8

4880.6
4800.8
4460.5
4371.3

2720.8
2660.7
2671.3
2751.5

810.9
910.7
750.7
761.2

2050.7
2000.6
1890.4
1660.9

620.2
680.3
470.4
180.2

HMT+water
1420.7
1420.7
0
HMT+1% HP-CD
1300.5
1300.5
0
HMT+5% HP-CD
1400.6
1400.6
0
HMT+10% HP-CD
1481.1
1481.1
0
PV=peak viscosity, HPV=hot paste viscosity, BD=breakdown, CPV=cold paste viscosity, SB=setback.
HMT=heat-moisture treated.
Values are means of at least duplicate determinations standard deviation.

167

Table 5. DSC parameters of retrograded native and heat-moisture treated wheat and potato starches in the presence of HP-CD
Starch
Treatment
To (C)
Tp (C)
Tc (C)
Tc-To (C)
H (J/g)
Wheat
Native+water
58.80.2
64.40.1
73.00.3
14.40.3
5.10.4
Native+1% HP-CD
59.00.4
64.50.2
73.00.3
14.20.3
4.90.1
Native+5% HP-CD
59.00.1
64.60.5
73.10.7
14.30.1
5.10.2
Native+10% HP-CD
58.60.6
64.40.3
73.20.1
14.40.4
5.00.3
HMT+water
HMT+1% HP-CD
HMT+5% HP-CD
HMT+10% HP-CD

58.90.4
59.00.8
59.00.3
59.10.2

64.70.1
65.00.2
65.00.2
65.10.3

73.00.3
73.00.4
73.10.1
73.60.7

14.50.1
14.20.2
14.00.3
14.30.2

4.90.6
4.90.1
4.80.2
5.00.4

Native+water
Native+ 1% HP-CD
Native+5% HP-CD
Native+10% HP-CD

63.20.1
63.80.2
63.70.3
62.90.2

70.90.4
70.60.3
70.70.1
70.0.6

82.10.4
82.00.6
82.00.6
81.90.4

18.40.2
18.50.1
18.60.2
19.10.5

7.90.3
7.80.4
7.90.1
8.10.2

HMT+water
63.80.5
70.50.2
HMT+1% HP-CD
63.20.1
70.60.1
HMT+5% HP-CD
63.50.1
70.60.5
HMT+10% HP-CD
62.50.1
70.30.4
To=onset, Tp=peak, Tc=conclusion, H=gelatinization enthalpy.
HMT=heat-moisture treatment.
Values are means of triplicate determinations standard deviation.

83.00.2
81.50.5
83.00.3
81.30.8

19.10.2
18.30.1
19.20.1
18.90.1

6.70.4
6.90.4
6.80.3
7.00.2

Potato

168

25

30

Amylose leaching (%)

25

20

20
15

15
10

10
5

5
0

50

60

70

80

90

100

50

60

70

80

90

100

Temperature (C)
Fig. 1. Amylose leaching of native and heat-moisture treated wheat (A) and potato (B) starches in the presence of HP-CD, =Native+water;
=Native+1%; =Native+5%; =Native+10%; =HMT+water; -=HMT+1%; +=HMT+5%; =HMT+10%.

169

-0.3

-0.2

N+water

N+water
-0.4

-0.4

Heat flow (w/g)

N+5%
N+5%
-0.5

-0.6

HMT+water

-0.6

HMT+water
HMT+5%
-0.8

HMT+5%

-0.7

-0.8

40

60

80

100

120

-1.0

40

50

60

70

80

90

100

110

Temperature (C)
Fig. 2. Representative (at 5% concentration of HP-CD) DSC curves of native and heat-moisture treated wheat (A) and potato (B)
starches in the presence of HP-CD. N=native; HMT=heat-moisture treatment.

170

0.2

-0.1

N+water

-0.2

0.0

Heat flow (w/g)

N+water

N+5%

-0.3

N+5%

HMT+water

-0.4

HMt+water

-0.2

-0.4
-0.5

HMT+5%
HMT+5%
-0.6
-0.6

-0.8

40

50

60

70

80

90

100

-0.7

40

60

80

100

120

Temperature (C)
Fig. 3. Representative (at 5% concentration of HP-CD) DSC curves of retrograded native and heat-moisture treated wheat (A) and potato (B)
starches in the presence of HP-CD. N=native; HMT=heat-moisture treated.

171

100

300

750

100

N+
water

240

80

N+1%

600

80

N+10%
N+water

180

N+1%

N+
5%

120

N+
10%

60

HMT+10%
HMT+water

HMT+
5%

60

450

40

300

20

150

60

HMT+
water

HMT+1%

20

HMT+10%

HMT+
5%

HMT+1%
0

0
0

10

15

20

25

40

Temperature (C)

Viscosity (RVU)

N+5%

0
0

10

15

20

25

Time (Mins)
Fig. 4. Pasting profiles of native and heat-moisture treated wheat (A) and potato (B) in the presence of HP-CD. N=native;
HMT=heat-moisture treated.

172

Chapter 6
Physical Properties and Digestibility of Acetylated Wheat, Potato,
Waxy Maize, and High-Amylose Maize Starches as Affected by
Hydroxypropyl -Cyclodextrin
Abstract
The effect of hydroxpropyl -cyclodextrin (HP-CD) on physical properties and
digestibility of wheat, potato, waxy maize and high-amylose maize starches before
and after acetylation was studied. Effect of HP-CD on amylose-lipid complexes in
native and acetylated potato starches synthesized using -lysophosphatidylcholine
was also studied. Acetylation increased swelling factor, amylose leaching, peak
viscosity and susceptibility to -amylase hydrolysis but decreased gelatinization
temperature and enthalpy, retrogradation and gel hardness in all starches. HP-CD
markedly increased swelling factor and amylose leaching in native and acetylated
wheat starches but had little or no impact on other starches. Wheat starch
gelatinization enthalpy decreased in the presence of HP-CD but gelatinization
temperature of all the starches was slightly increased. HP-CD had no influence on
amylopectin retrogradation and enzymatic hydrolysis. Melting enthalpy of amyloselipid complex in both native and acetylated wheat starches was decreased by HP-CD.
Acetylation also decreased the melting enthalpy of amylose-lipid complex in wheat
starch. Similar trend of thermal transitions was observed in the presence of HP-CD
for the amylose-lipid complexes synthesized in potato starch. Acetylation reduces the
complex formation ability of amylose polymer. Similar to gelatinization, acetylation
widened the melting temperature range of amylose-lipid complexes while shifting it
to a lower temperature. Higher swelling and amylose leaching, and decreased
gelatinization temperature and enthalpy resulting from acetylation of wheat starch is
consistent with its influence on starch hydration. Similar effects resulting from the

173

inclusion of HP-CD were consistent with the disruption of amylose-lipid complex


by HP-CD which promotes granular hydration.

174

6.1. Introduction
Previous results of this study showed that hydroxypropyl -cyclodextrin (HP-CD)
increased amylose leaching, swelling and solubility of cereal starches in a manner
consistent with the disruption of amylose-lipid complex within the starch granule by
complexing with starch lipids. In food systems, behavior of the amylose-lipid
complex is of technological interest because it can affect the quality of starch-based
food products. For example decreased swelling, solubilization and thickening power
of starch (Galliard and Bowler, 1987), retardation of starch retrogradation and bread
firming (Krog, 1971; Biliaderis and Tonogai, 1991), prevention of stickiness of dried
potato (Hoover and Hadziyev, 1981), and improvement of structural integrity of
cereal kernels during cooking (e.g., parboiled rice) [Biliaderis et al., 1993] have been
reported in the presence or formation of the amylose-lipid complex. Studies have
shown that in the process of enzymatic hydrolysis of wheat starch to glucose, the
presence of amylose-lipid complexes decreased swelling and dissolving capacity and
the water binding capacity of starch and thereby delays the access of amylolytic
enzymes into the starch granules (Nebesnky et al., 2002, 2004). Furthermore, the
complexes negatively influence the color, transparency and aroma of starch
hydrolysates as well as their filtration rate (Master and Steeneken, 1998a and b). The
resistance towards -amylase hydrolysis increased when amylose complexed with
lipids. This could reduce the effectiveness of starch hydrolysis requiring more
enzyme or longer treatment. Previous results of this study showed that modified -CD
is more effective in disrupting native cereal starch amylose-lipid complex than its
unmodified counterpart.

Acetylation is one of the commonly used techniques for making modified starch.
Substituting acetyl groups onto starch chains loosens the starch structure by

175

preventing interchain association of adjacent starch chains facilitating starch granule


hydration (Liu et al., 1999 and 1997). Extensively hydrated starch granules required
less energy to reach gelatinization. By preventing interchain association acetylation
decreased retrogradation providing greater freeze-thaw stability (Liu et al., 1999 and
1997). It can be expected that after substituting more bulky acetyl groups onto starch
chains, there could be some specific interaction with larger HP-CD molecules
creating some novel starch functional properties. Acetylation could also affect the
amylose-lipid complex and the complex forming ability of amylose chains. This
study therefore aimed to characterize the interaction between HP-CD and acetylated
starches and its effect on some structural and functional properties of starch.

6.2. Materials and Methods

6.2.1. Materials
Potato

and

wheat

starches,

hydroxypropyl

-cyclodextrin

(MS=0.8),

lysophosphatidylcholine and fungal -amylase were from Sigma Chemical Co., (St.
Louis, MO, USA). Waxy maize and high-amylose maize starches were obtained from
Starch Australasia Limited. Acetic anhydride was from Merck Co., Germany.

6.2.2. Methods

Total amylose content


An amylose/amylopectin assay kit from Megazyme International Co. Wicklow,
Ireland, was used to estimate the total amylose content of all the starches. This assay

176

is based on the principle of specific formation of amylopectin complex with


Concanavalin-A (Con A), after a pre-treatment to remove lipids.

Total lipids
Total lipids in starches were extracted with 2:1 chloroform:methanol solvent for 5 hr
using an extraction/desolventizing unit (Soxtec System HT6, Tecator, Sweden).

Swelling factor
Swelling factor, the ratio of the volume of swollen starch granules to the volume of
dry starch was determined by the method of Tester and Morrison (1990a), when
starch (50 mg, db) was heated at 85 C in 5 mL of water.

Amylose leaching
Distilled water or solution (10 mL) was added to starch (20 mg, db) in a screw cap
tube. Tubes were then heated at 85 C for 30 min. After cooling to ambient
temperature, samples were centrifuged at 2000 g for 10 min. Amylose content of
supernatant (0.1 mL) was estimated as described by Chrastil (1987).

Differential scanning calorimetry


Gelatinization and dissociation parameters were measured using a TA 2920
Modulated DSC Thermal Analyzer differential scanning calorimeter equipped with a
thermal analysis data station (TA Instruments, Newcastle, DE). Starch (3 mg) was
directly measured on to the aluminum DSC pan and distilled water (3 L) was added
with a microsyringe and mixed for homogenization. Pans were sealed, and allowed to
stand for 1 hr at room temperature for even distribution of water. The scanning

177

temperature and the heating rates were 30-120 C and 5 C/min respectively. An
empty pan was used as reference for all measurements.

Differential scanning calorimetry of synthesized amylose-lipid complex in native


and acetylated potato starches using -lysophosphatidylcholine
-lysophosphatidylcholine potato starch (native and acetylated) amylose-lipid
complex was synthesized in the DSC pan. Starch (3 mg) was directly measured onto
the DSC pan and -lysophosphatidylcholine (0.6 mg dissolved in water) was added to
make 5:1 starch to -lysophosphatidylcholine ratio. Distilled water (3 L) was added
and mixed for homogenization. Pans were sealed, and heated at 90 C for 1.5 hr in an
incubator. After cooling to room temperature pans were scanned. The scanning
temperature and the heating rates were 30-140 C and 10 C /min respectively.

Pasting properties
Pasting properties of starches were determined using a Rapid Visco-Analyser (RVA)
model 3D (Newport Scientific, Warriewood, Australia). Distilled water (25.5 g) was
added to starch (2.5 g, db) in the RVA canister to obtain a total constant sample
weight of 28 g. The slurry was then manually homogenized using the plastic paddle
to avoid lump formation before the RVA run. A programmed heating and cooling
cycle was set for 22 minutes, where it was first held at 50 C for 1.0 min, heated to 95
C in 7.5 min, further held at 95 C for 5 min, cooled to 50 C within 7.5 min and
held at 50 C for 1 min.

Gel hardness
Gel hardness was determined on the starch gel made in the RVA testing using a TAXT2 Texture Analyzer (Stable Micro Systems, Godalming, Surrey, England). After
RVA testing, the paddle was removed and the starch paste in the canister was covered
178

by Parafilm and stored at 4 C for 4 hr. The gel was compressed at a speed of 0.5
mm/sec to a distance of 10 mm with a 6 mm cylindrical probe. The maximum force
peak in the TPA profile represents the gel hardness.

Retrogradation
After the gelatinization test samples were stored at 4 C for 24 hr to initiate nucleation,
and then were kept at 35 C for 7 days before rescanning. Temperature range and
heating rate were 30-120 C and 10 C/Min respectively.

Enzymatic hydrolysis
Enzymatic hydrolysis was measured using a viscoamylographic (RVA) method as
described by Li et al. (2000) with slight modifications. Development of peak
viscosity was measured before and after adding 100 units of fungal -amylase
(enzyme was directly added to the canister) and percent decrease of peak viscosity
was calculated to detect the extent of enzyme hydrolysis. In this analysis, same
heating cooling program (Std. 2), and starch concentration were used as in the
measurement of pasting properties.

Acetylation
Acetylation was carried out as described by Wang and Wang (2002) with slight
modifications. Starch (100 g db) was dissolved in distilled water (185 mL) to make
35% slurry. The pH of the slurry was adjusted to 8.0-8.5 with 1M NaOH and then
mechanically stirred for 30 min. Acetic anhydride (8 g) was slowly (dropwise) added
while maintaining pH 8.0-8.5. The reaction was continued for 90 min before
acidifying to pH 5.5 with 1M HCl. The slurry was then washed with three fold
distilled water three times and dried at 35 C.

179

Note: Hydroxypropyl -cyclodextrin (MS=0.8) 0.01M was added instead of water in


all above experiments where appropriate.

6.3. Results and Discussion

Swelling factor and amylose leaching


Swelling factor of native starches at 85 C followed the order: potato > waxy maize >
wheat > high-amylose maize starch (Table 1). As commonly observed acetylation
increased swelling ability of all the starches by promoting granule hydration. HPCD increased swelling factor of wheat starch to a greater extent but had less or no
impact on low lipid starches (potato and waxy maize) [Table 1]. Although highamylose maize starch had comparatively high content of lipids, it showed less effect
of HP-CD. Acetylation substantially increased amylose leaching of all the starches
whereas HP-CD increased amylose leaching only in wheat starch (Table 2). Similar
to native starch, swelling factor and amylose leaching of acetylated wheat starch was
increased in the presence of HP-CD. HP-CD had little or no impact on swelling
factor and amylose leaching of other acetylated starches.

Consistent with previous findings of this study, only native and acetylated wheat
starches (higher lipid content starches) showed higher amylose leaching in the
presence of HP-CD consistent with disruption of the amylose-lipid complex. It was
explained that the formation of amylose-lipid complex either naturally or in situ
reduced the swelling ability and amylose leaching of the starch granules (Richardson
et al., 2003; Becker et al., 2001; Morrison et al., 1993; Morrison, 1981; Biliaderis et
al., 1985). Virtually lipid-free potato and waxy maize starches (Table1) showed lesser
interaction with HP-CD. In case of high amylose-maize starch it appears that
180

densely packed amylose chains prevent the penetration of both water and HP-CD to
the starch granules resulting in low swelling and amylose leaching. Despite, fairly
high content of total lipids in high-amylose maize starch (Table 1), DSC curves did
not show the presence of amylose-lipid complex (curves not shown). Increased
amylose leaching resulted in all the acetylated starches indicating that extensively
hydrated starch granules after acetylation promote the mobility of starch chains.
Perhaps by preventing interactions between adjacent amylose chains acetyl groups
facilitate the amylose leaching.

Gelatinization
Gelatinization parameters of native and acetylated starches in the presence of HP(Table 3) and corresponding DSC curves of native and acetylated wheat starches (Fig.
1) are presented. Melting parameters of synthesized amylose-lipid complex using lysophosphatidylcholine in native and acetylated potato starches in the presence of
HP-CD (Table 3) and corresponding DSC curves (Fig. 2) are presented. Decreases
were observed for gelatinization temperature and enthalpy of all the starches after
acetylation. HP-CD decreased gelatinization enthalpy only in wheat starch but it
slightly increased gelatinization temperature of all the starches. Gelatinization
enthalpy of potato, waxy maize and high-amylose maize starch was unchanged in the
presence of HP-CD.

Introduction of acetyl groups could prevent the interchain association loosening the
starch structure. This facilitates more water penetration into the starch granules with a
consequent increase of swelling. Accelerated granular swelling reduces the energy
requirement to disrupt the starch structure because gelatinization is a swelling driven
process. As observed in my previous studies, disruption of amylose-lipid complex by

181

HP-CD also hydrates the wheat starch granules reducing the energy requirement to
reach gelatinization.

Melting enthalpy of amylose-lipid complexes in both native and acetylated wheat


starches was decreased in the presence of HP-CD but melting temperature was
unaffected. Acetylation also decreased the melting enthalpy of amylose-lipid complex
but melting temperature was shifted to a lower temperature point (Table 2 and Fig.1).
Virtually lipid-free potato starch did not show any amylose-lipid complex, but a
formation of amylose-lipid complex was evident when -lysophosphatidylcholine
was added (Fig. 2). Interestingly, transition parameters observed for the synthesized
potato--lysophosphatidylcholine amylose-lipid complexes followed a similar trend
as that exhibited for native and acetylated wheat starch amylose-lipid complexes in
the presence of HP-CD (Table 2 and Fig. 2). Melting enthalpy of amylose-lipid
complex synthesized in acetylated potato starch is lower than that of amylose-lipid
complex synthesized in native potato starch. This indicates that acetyl groups in
amylose chains reduce the complex formation ability of amylose polymer.
Acetylation increases the hydrophobicity of amylose chains. This may change the
molecular characteristics of amylose chains negatively affecting the complex
formation. Perhaps, depending on degree of acetylation and distribution mode
(random or bloked) of acetyl groups on amylose chains, actual helix formation may
be diminished thus making unfeasible the development of well-organized amylose
structures (V-structure). Similar to gelatinization, acetylation broadens the transition
temperature range of amylose-lipid complexes. Reduction of melting temperature and
enthalpy of wheat starch amylose-lipid complex after acetylation is most probably
due to the greater hydration of starch granules resulting from acetylation. Substitution
could also take place in the amylose-lipid complex weakening its bonding forces and

182

structurally weaker complex may require less energy to complete the thermal
transition. In contrast, HP-CD reduces melting enthalpy of amylose-lipid complexes
by complexing with starch lipids and thereby increases the hydration of the starch
granules.

Pasting properties
All starches showed increased peak viscosity and early onset of pasting after
acetylation. Inclusion of HP-CD decreased peak viscosity of wheat and potato
starches but it slightly increased in waxy maize starch and high-amylose maize starch
was unaffected (Table 4 and Fig. 3). The early onset of peak viscosity was noticed for
wheat starch in the presence of HP-CD but it not in other starches (Fig.3). Addition
of HP-CD to acetylated wheat starch increased peak viscosity and further decreased
the onset of pasting but other acetylated starches showed no influence on peak
viscosity or pasting onset temperature in the presence of HP-CD. Viscosity in a
dilute system is governed by the volume fraction of swollen granules but in a
concentrated system it is governed by particle rigidity (Steenekan, 1989). Doublier
(1987) suggested that overall viscosity of starch paste is primarily governed by a
combination of swollen starch granules and composition of the continuous phase thus
swelling, amylose leaching and rupture of starch granules play a key role in
determining the development of viscosity. Early viscosity increase of wheat starch in
the presence of HP-CD could relate to its greater influence on swelling and amylose
leaching. Perhaps addition of HP-CD could promote the close packing of swollen
starch granules resulting in an early onset of viscosity. Evidence has been reported
that polyhydroxy compounds such as sugars accelerate early onset of peak viscosity
by promoting the close packing of swollen starch granules (Richardson et al., 2003;
Doublier et al., 1987). Disruption of amylose-lipid complex followed by extensive

183

swelling in the presence of HP-CD may weaken the structure of swollen wheat
starch granules causing more deformation in swollen granules during stirring. This
can reduce the peak viscosity. However, addition of HP-CD to acetylated wheat
starch increased PV and caused early viscosity development. This indicates that
acetylated wheat starch granules have gained some granular rigidity (resistance to
deformation during stirring) and close packing tendency with the inclusion of HPCD. Several studies have shown that polyhydroxy compounds such as sugars can
bridge starch chains stabilizing amorphous region (Spice and Hoseney, 1982; Baek et
al., 2004; Chiotelli et al., 2000; Hoover and Senanayake, 1996). This kind of
interaction could be promoted by introducing acetyl groups to starch chains in the
amorphous region. However, inclusion of HP-CD had less impact on peak viscosity
of other acetylated starches. This discrepancy may relate to the variation of physical
properties of starch polymers, swelling tendency, and close packing ability in the
presence of HP-CD. Increased cold paste viscosity of acetylated starches is
consistent with higher amylose leaching resulting from acetylation because in cooling,
starch paste gains some viscosity increase primarily due to the realignment of
amylose chains. HP-CD increased setback (retrogradation) of wheat starch. As
shown in my previous study, HP-CD facilitates amylose leaching during pasting by
the action of defatting the amylose chains. Lipid free amylose can retrograde to a
greater extent than lipid complexed amylose.

Gel hardness of the native starches followed the order: wheat > potato > waxy maize
(Table1). High-amylose maize starch produced a very soft gel that was not
measurable under the experimental conditions. HP-CD decreased gel hardness of all
the starches, but most in wheat starch. This is in agreement with previous results of
this study. Acetylation markedly decreased gel hardness of all the starches. Except

184

potato starch, gel hardness was not measurable other acetylated starches under the
conditions used. Although acetylation and HP-CD increased amylose leaching, the
resulting gels were very soft. According to Ring (1985) the starch gel is a composite
in which swollen gelatinized starch granules reinforce an interpenetrating amylose gel
matrix. Mechanical properties of starch gel would depend on the rheological
characteristics of the amylose matrix, the volume fraction and the rigidity
(deformability) of the gelatinized granules, and the interactions between the dispersed
and the continuous phases (Eliasson, 1986). Properties of starch gel should be
interpreted in relation to the properties of the gel matrix amylose, deformable filler
swollen particles, volume fraction of the swollen particles, and filler-matrix
interaction (Morris, 1990). The decreased gel hardness of acetylated starch therefore
should be due to the interplay among the following factors; less rigid swollen
particles resulted from acetylation; substituted acetyl groups in the filler component
(swollen starch granules) could weaken the interaction between filler component and
the gel matrix; acetyl groups could reduce the conformation ordering, and
intermolecular association of starch polymers by inhibiting the formation of junction
zones. Greatly reduced gel hardness of wheat starch in the presence of HP-CD could
be attributed to the weaker swollen gelatinized granules resulting in the disruption of
amylose-lipid complex and the subsequent greater hydration of starch granules.

Retrogradation
Potato and waxy maize starch showed greater amylopectin retrogradation while highamylose maize starch was less prone to it (Table 1 and Fig. 3). The diversity of
amylopectin retrogradation among native starches could be attributed to variation in
molecular characteristics of amylopectin, particularly average unit chain length of
outer branches of amylopectin. Except for high-amylose maize starch, acetylation

185

greatly decreased retrogradation by inhibiting the interchain association of


amylopectin. Densely packed amylose chains in the amorphous region of highamylose maize could prevent the access of reaction reagent to the highly ordered
crystalline region. This could prevent the modification of amylopectin polymers in
high-amylose maize starch. Unchanged amylopectin retrogradation for both native
and acetylated starches in the presence of HP-CD suggests that it has no influence
on the reassociation of amylopectin chains in the starch paste.

Enzymatic hydrolysis
Percent decrease of peak viscosity in viscoamylometry was used to detect the extent
of starch hydrolysis by -amylase (Table 5 and Fig. 4). The extent of enzymatic
hydrolysis in native starches followed the order: waxy maize > potato > wheat. For
high-amylose maize starch this method cannot be successfully applied since it does
not develop a distinguishable peak viscosity. Acetylation increased the enzymatic
digestibility of all the starches. Loosened starch structure resulting after acetylation
due to the disruption of hydrogen bonding between adjacent starch chains could
increase the enzyme access to starch granules. Weselake and Hill (1983) reported that
-CD inhibited cereal -amylase hydrolysis by binding with the`non-catalytic site of
the enzyme. This inhibits the binding of -amylase with the substrate (starch). The
same authors (Weselake and Kim, 1982) reported that -amylase has strong affinity
for -CD. However in this study I did not see any increase of percent peak viscosity
for both native and acetylated starches with enzyme added in the presence of HP-CD
indicating no inhibition of -amylase. In my earlier study I observed that both HPCD and -CD had no inhibiting effect on starch hydrolysis by bacterial -amylase,
fungal -amylase and -amylase using the same method. Similarly, no inhibition was
reported for wheat starch bacterial -amylase hydrolysis in the presence of -CD (Li

186

et al., 2000). This discrepancy however, may be attributed to the variation among
types of -amylases, cyclodextrins and methods used for the analysis.

6.4. Conclusions
Both acetylation and addition of hydroxypropyl -cyclodextrin are capable of
increasing swelling and amylose leaching while decreasing gelatinization enthalpy of
wheat starch. The common cause for these changes in properties relates to the
hydration of the starch granule. Loosened starch structure by acetylation prevents
interchain association facilitating more water penetration into the granules. By
disrupting the amylose-lipid complex by complexing with starch lipids,
hydroxypropyl -cyclodextin extensively hydrates the starch granules. Acetylation
reduces the complex forming ability of amylose and facilitates the dissociation of the
amylose-lipid

complex.

Understanding

the

transformation

and

dissociation

parameters of amylose-lipid complex within the starch granules is of great


fundamental and technological importance, considering the multifunctional role of
lipids in starch-based products. This study showed that HP-CD is capable of
disrupting amylose-lipid complex making significant changes to starch functional
properties. Ability to disrupt amylose-lipid complexes by HP-CD could successfully
be applied to improve processing effectiveness such as in production of glucose syrup
from cereal starches via amylolytic reaction. This is possible because HP-CD has no
inhibiting effect on -amylase hydrolysis and thus HP-CD and -amylase can be
used together. A small quantity of HP-CD is sufficient to greatly destabilize the
amylose-lipid complex.

6.5. References

187

Baek, M.H., Yoo, B., and Lim, S-T. 2004. Effects of sugars and sugar alcohols on the
thermal transition and cold stability of corn starch gel. Food Hydrocolloids 18: 133142.
Becker, A., Hill, S.E., and Mitchell, J.R. 2001. Relevance of amylose-lipid complex
to the behavior of thermally processed starches. Starch/Starke 53: 121-130.
Biliaderis, C.G., and Tonogai, J.R. 1991. Influence of lipids on the thermal and
mechanical properties of concentrated starch gels. J. Agric. Food Chem. 39: 833-840.
Biliaderis, C.G., Page, C.M., Slade, L., and Sirett, R.R. 1985. Thermal behavior of
amylose-lipid complexes. Carbohydr. Polym. 5: 367-389.
Biliaderis, C.G., Tonogai, J.R., Pereze, C.M., and Juliano, B.O. 1993.
Thermophysical properties of milled rice starches as influenced by variety and
parboiling method. Cereal Chem. 70: 512-516.
Chiotelli, E., Rolee, A., Meste, M.L. 2000. Effect of sucrose on the thermomechanical
behavior of concentrated wheat and waxy corn starch-water preparations. J. Agric.
Food Chem. 48: 1327- 1339.
Chrastil, J. 1987. Improved colorimetric determination of amylose in starches or
flours. Carbohydr. Res. 159: 367-389.
Doublier, J.L., Llamas, G., and Le Meur, M. 1987. A rheological investigation of
cereal starch pastes and gels. Effects of pasting procedures. Carbohydr. Polym. 7:
251-275.
Eliasson, A.C. 1986. Viscoelastic behavior during the gelatinization of starch. I.
Comparison of wheat, maize, potato, and waxy-barley starches. J. Text. Stud. 17: 253265.
Galliard, T., and Bowler, P. 1987. Morphology and composition of starch. In: Starch:
Properties and Potential. T.Galliard (Ed.,), pp. 55-79. Wiley, Chichester, England.
Hoover, R., and Hadziyev, D. 1981. Characterization of potato starch and its
monoglyceride complexes. Starch 33: 290-300.
Hoover, R., and Senanayake, N. 1996. Effect of sugars on the thermal and
retrogradation properties of oat starches. J. Food Biochem. 20: 65-83.
Krog, N. 1971. Amylose complexing effect of food grade emulsifiers. Starch 23: 206210.
Li, W.D., Huang, J.C., and Corke, H. 2000. Effect of -cyclodextrin on pasting
properties of wheat starch. Nahrung -Food 44: 164-167.
Liu, H.J., Ramsden, L., and Corke, H. 1997. Physical properties and enzymatic
digestibility of acetylated ae, wx and normal maize starch. Carbohydr. Polym. 34:
283-289.

188

Liu, H.J., Ramsden, L., and Corke, H. 1999. Physical properties of cross-linked and
acetylated normal and waxy rice starch. Starch/Starke 51: 249-252.
Master, M., and Steeneken, P.A.M. 1998a. Filtration characteristics of maize and
wheat starch hydrolysates. Cereal Chem. 75: 289-293.
Master, M., and Steeneken, P.A.M. 1998b. Origins of the poor filtration
characteristics of wheat starch hydrolysates. Cereal Chem. 75: 289-293.
Morris, M.J. 1990. Starch gelation and retrogradation. Trends in Food Science &
Technology 1: 2-6.
Morrison, W.R. 1981. Starch lipids: A reappraisal. Starch/Starke 33: 408-410.
Morrison, W.R., Tester, R.F., Snape, C.E., Law, R., and Gidley, M. 1993. Swelling
and gelatinization of cereal starches. IV. Some effects of lipid-complexed amylose
and free amylose in waxy and normal barley starches. Cereal Chem. 70: 385-391.
Nebesnky, E., Rosicka, J., and Tkaczyk, M. 2002. Effect of enzymatic hydrolysis of
wheat starch on amylose-lipid complexes stability. Starch 54: 603-608.
Nebesnky, E., Rosicka, J., and Tkaczyk, M. 2004. Influence of conditions of maize
starch enzymatic hydrolysis on physicochemical properties of glucose syrup. Starch
56: 132-137
Richardson, G., Langton, M., Bark, A., and Hermansson, A-M. 2003. Wheat starch
gelatinization-the effect of sucrose, emulsifiers and the physical state of the emulsifier.
Starch 55: 150-161.
Ring, S.G. 1985. Some studies on starch gelation. Starch 37: 80-83.
Spies, R.D., and Hoseney, R.C. 1982. Effect of sugar on starch gelatinization. Cereal
Chem. 59: 128-131.
Steeneken, P.A.M. 1989. Rheological properties of aqueous suspensions of swollen
starch granules. Carbohydr. Polym. 11: 23-42.
Tester, R., and Morrison, W.R. 1990a. Swelling and gelatinization of cereal starches.
. Effects of amylopectin, amylose, and lipids. Cereal Chem. 67: 551-557.
Wang, Y.J., and Wang, L., 2002. Characterization of acetylated waxy maize starches
prepared under catalysis by different alkali and alkaline-earth hydroxides. Starch 54:
25-30.
Weselake, R.J., and Hill, R.D. 1982. Cycloheptaamylose as an affinity ligand of
cereal -amylase. Characteristics and a possible mechanism of the interaction.
Carbohydr. Res. 108: 153 -161.
Weselake, R.J., and Hill, R.D. 1983. Inhibition of alpha-amylase-catalyzed starch
granule hydrolysis by cycloheptaamylose. Cereal Chem. 60: 98-101.

189

Table 1. Swelling factor (SF) at 85 C, gel hardness (GH), retrogradation


enthalpy (HR), total lipid, and total amylose content of starches
Starch
Treatmenta
SF
GH
HR
Total lipid
(g)
(J/g)
(%)
Wheat
Native
13.20.2 741.7 4.20.1 1.10.2
Native+CD 48.10.4 420.6 4.40.2
AC
21.10.3 1.80.1
AC+CD
35.20.2 1.80.4

Total
amylose (%)
24.30.3

Potato

Native
Native+CD
AC
AC+CD

49.60.5
49.90.1
58.10.3
57.90.2

310.8
230.7
150.2
140.6

8.20.3
8.10.2
5.7.0.3
5.50.5

0.10.2

24.50.2

Waxy maize

Native
Native+CD
AC
AC+CD

38.70.1
40.10.6
44.10.4
44.60.2

70.9
-

9.10.2
9.00.3
3.30.1
3.40.1

0.150.1

3.80.1

High-amylose
maize

Native

3.20.3

2.30.2

0.70.1

65.40.2

Native+CD 3.50.1 2.50.2


AC
9.10.2 2.50.1
AC+CD
8.90.1 2.60.2
a
AC=acetylated, CD=HP-CD.
Values are means of triplicate determinations standard deviation.

190

Table 2. Amylose leaching of native and acetylated starches at different


temperature in the presence of HP-CD
Starch
Treatmenta 60 (C )
70 (C )
80 (C )
90 (C )
Wheat
Native
1.20.1
4.70.3
10.30.2
23.10.1
Native+CD 5.20.2
20.10.4
22.30.1
23.80.2
AC
1.20.1
7.10.1
13.80.4
23.20.3
AC+CD
1.60.1
19.20.5
21.20.2
22.60.1
Potato

Native
Native+CD
AC
AC+CD

5.10.3
5.30.1
7.40.1
7.50.2

10.20.3
10.80.2
11.90.4
11.30.4

14.10.5
14.80.2
16.10.1
15.10.3

19.20.5
19.30.3
20.10.1
19.70.3

Waxy maize

Native
Native+CD
AC
AC+CD

0.20.0
0.30.0
0.20.0
0.10.0

0.60..03
0.50.01
0.60.02
0.60.01

1.20.0
1.40.1
1.30.2
1.00.0

2.10.2
2.60.1
2.40.1
2.50.0

High-amylose
maize

Native

0.0

2.10.1

3.40.2

8.20.1

Native+CD 0.0
2.00.1
3.80.4
AC
1.20.1
4.80.3
6.80.2
AC+CD
1.20.0
4.70.4
6.30.0
a
AC=Acetylated, CD=HP-CD.
Values are means of triplicate determinations standard deviation.

7.90.3
15.80.2
13.10.2

191

Table 3. Gelatinization parameters and dissociation parameters of amylose-lipid complexes (native wheat starch amylose-lipid complex and synthesized
amylose-lipid complex in native and acetylated potato starches) in the presence of HP-CD
Gelatinization parametersa
Transition parameters of amylose-lipid complexesb
Starch
Treatmentc
To (C)
Tp (C)
Tc (C)
Tc-To
HG
To (C)
Tp (C)
Tc (C)
Tc-To
HA(J/g)
(J/g)
Wheat
Native
55.30.2 59.40.3 83.70.1
28.40.1
11.10.2 103.10.1 106.30.1 112.80.4 9.70.5 1.70.2
Native+CD 55.90.1 59.80.5 84.80.3
28.90.3
9.70.1 101.30.4 105.60.2 113.10.6 11.80.3 0.80.2
AC
52.20.3 56.10.2 82.40.2
30.20.1
9.60.2 96.40.5
103.70.3 109.50.2 13.10.2 1.20.1
AC+CD
52.20.1 56.40.1 81.50.1
29.30.2
8.40.1 94.20.7
103.10.1 109.20.3 15.00.1 0.70.1
Potato

Native
Native+CD
AC
AC+CD

59.80.4
60.70.1
55.30.2
56.40.1

64.20.3
65.50.1
60.50.2
61.60.3

82.60.4
81.80.5
80.80.1
80.10.3

22.80.3
21.10.3
25.50.2
23.70.1

15.50.2
15.90.2
14.20.3
14.30.1

Waxy maize

Native
Native+CD
AC
AC+CD

62.80.3
64.20.1
59.20.3
60.30.4

68.40.1
70.00.3
65.20.3
66.70.4

88.30.1
90.20.2
85.90.1
86.90.3

25.50.2
26.00.2
26.70.3
26.60.1

15.10.3
14.90.4
12.30.4
12.10.2

High-amylose
maize

Native

69.90.1 78.30.1 111.70.4

41.80.2

10.50.1

100.20.3
99.50.2
95.40.4
95.20.2

111.10.1
110.20.3
107.10.2
106.40.4

120.20.1
119.60.2
118.20.1
118.10.3

20.00.2
20.30.4
22.80.3
23.20.5

5.80.2
4.90.1
3.50.1
2.80.1

Native+CD 71.20.2 80.40.2 111.50.3 40.30.1


10.40.3
AC
62.20.1 74.60.1 105.30.2 43.10.1
9.50.1
AC+CD
62.90.3 74.80.1 105.60.2 42.70.2
9.80.2
a
To=onset, Tp=peak, Tc=conclusion, HG = gelatinization enthalpy.
b
To=onset, Tp=peak, Tc=conclusion, HA= melting enthalpy of amylose-lipid complexes.
c
AC=acetylated, CD=HP-CD. Values are means of triplicate determinations standard deviation.

192

Table 4. Pasting properties of native and acetylated starches in the presence of HP-CD
Starch
Treatmentb
PVa
HPV
BD
Wheat
Native
1530.4
1141.3
390.4
Native+ CD
1280.7
981.2
290.6
AC
2440.5
991.5
1440.7
AC+CD
2701.2
1031.1
1681.3

CPV
2121.2
2160.8
2290.7
2001.4

SB
1010.3
1221.3
1301.5
970.8

Potato

Native
Native+CD
AC
AC+CD

6551.1
6400.9
7021.3
6971.5

1740.5
1791.1
1420.8
1110.4

4811.2
4591.8
5601.3
5860.5

2541.8
2581.5
2740.9
3001.2

811.3
761.4
1320.9
1891.3

Waxy maize

Native
Native+CD
AC
AC+CD

1750.4
2820.8
3531.3
3481.6

1150.9
1040.4
1360.8
1301.3

1780.7
1780.3
2170.9
2180.7

1471.3
1470.8
1741.5
1720.9

321.3
430.8
400.5
410.6

High-amylose maize

Native
30.1
30.1
10.1
30.2
Native+CD
30.2
30.2
10.1
30.3
AC
140.4
140.3
0.0
180.5
AC+CD
150.3
150.7
0.0
210.6
a
PV=peak viscosity, HPV=hot paste viscosity, BD=breakdown, CPV=cold paste viscosity, SB=setback,
b
AC=acetylated, CD= HP-CD.
Values are means of triplicate determinations standard deviation.

10.0
10.0
50.7
60.7

193

Table 5. Percent decrease of peak viscosity (PV) of native and acetylated


starches after addition of -amylase in the presence of HP-CD
Starch
Treatmenta
PV (control)
PV (with % Decrease
amylase)
Wheat
Native
1530.4
301.2
800.7
Native+CD
1280.7
170.8
860.8
AC
2440.5
190.5
920.6
AC+CD
2701.2
200.4
920.4
Potato

Native
Native+CD
AC
AC+CD

6551.1
6400.9
7021.3
6971.5

730.9
710.7
360.8
410.6

881.1
891.2
950.6
950.7

Waxy

Native
Native+CD
AC
AC+CD

1750.4
2820.8
3531.3
3481.6

100.3
90.5
60.8
60.5

940.8
940.6
980.9
981.2

High-amylose Native
30.1
2.50.4
160.3
maize
Native+CD
30.2
2.50.2
160.2
AC
140.4
100.1
290.3
AC+CD
150.3
100.5
330.6
a
AC=acetylated, CD= HP-CD, En=-amylase enzyme (100 mg).
Values are means of triplicate determinations standard deviation.

194

-0.2

Amylose-lipid complex
N+W

Heat flow (W/g)

N+CD

-0.3

AC+W

AC+CD

-0.4

40

60

80

100

Temperature (C)
Fig. 1. DSC curves of native and acetylated wheat starches in the presence
of HP-CD; N=native, W=water, AC=acetylated, CD=HP-CD.

195

-2.6

N+W

-2.8

N+CD

Heat flow (W/g)

-3.0

AC+W

-3.2

AC+CD
-3.4

-3.6

60

70

80

90

100

110

120

130

Temperature (C)
Fig. 2. DSC curves of synthesized amylose-lipid complexes using lysophosphatidylcholine in native and acetylated potato starches in the
presence of HP-CD. N+W=native potato starch amylose-lipid complex
with water, AC+CD=native potato starch amylose-lipid complex with
HP-CD, AC+W=acetylated potato starch amylose-lipid complex with
water, AC+CD=acetylated potato starch amylose-lipid complex with
HP-CD.

196

100

300

100

800

A
80

240

ac
ac+cd

180

80
600

60
400

60

n+ac

40

120

40

200
20

60

20

ac+cd

Viscosity (RVU)

ac
0

10

15

20

25

10

15

20

25

100

100

400

24
80

80

ac
300
60

200

ac

40

100

20

16

60

ac+cd
n+cd

20

ac+cd

40

n+ac
0

10

15

20

25

-8

10

15

20

Time (Mins)
Fig. 3. RVA curves of native and acetylated wheat (A), potato (B), waxy
maize (C), and high-amylose maize (D) in the presence of HP-CD. n=native,
ac=acetylated, cd= HP-CD.

197

25

Temperature (C)

n+cd

-0.6

-0.6

-0.8

-0.8

-1.0

-1.0

Heat flow (W/g)

-1.2

-1.4

40

50

60

70

80

90

100

-1.2

0.0

40

50

60

70

80

90

100

110

-0.4

-0.2

g)

-0.4

g)

-0.6

-0.8
-0.6

-1.0
-0.8

-1.0

40

50

60

70

80

90

-1.2

40

50

60

70

80

Temperature (C)
Fig. 3. DSC curves of retrograded native and acetylated wheat (A), potato
(B), waxy maize (C), and high-amylose maize (C) in the presence of HPCD. In each figure, curves (from top to bottom) are native with water,
native with HP-CD, acetylated with water, and acetylated with HP-CD.

198

90

100

30

10

ac+cd

ac

18

12

cd

ac+
cd+en

ac+en

n+cd+en

n+en

Time (Mins)
Fig. 4. RVA curves of enzyme added native and acetylated wheat starches
in the presence of HP-CD. n=native, ac=acetylated, cd= HP-CD,
en=enzyme (100 units).

199

Temperature (C)

Viscosity (RVU)

24

Chapter 7
Influence of Prior Acid Treatment on Acetylation of Wheat, Potato
and Maize Starches
Abstract
Functional properties of acid-thinned, acetylated, and acid-thinned acetylated wheat,
potato, and maize starches were investigated. Total amylose content and the extent of
amylose leaching were increased after acid-thinning in all the starches indicating
creation of more linear segments. Acid treatment decreased the swelling factor of
potato starch at all tested acid concentrations but slightly increased it in wheat and
maize starches in low acid treatment. Acid-thinning increased gelatinization
temperatures and enthalpy but greatly decreased peak viscosity of all the starches.
Wheat and maize starches showed early viscosity increase after acid-thinning but it
was delayed in potato starch. Potato starch produced firmer gels after acid-thinning
but wheat starch gave weaker gels. For maize starch, gel hardness was increased with
low concentration acid treatment. Unchanged melting enthalpy of amylose-lipid
complex in acid-thinned starch reflects its resistance to acid hydrolysis. Amylopectin
retrogradation of potato and maize starches was not affected by acid modification but
it decreased in wheat starch. Acetylation decreased gelatinization temperature and
enthalpy, amylopectin retrogradation and gel hardness, but increased swelling factor,
amylose leaching and peak viscosity. Acid-thinning decreased the degree of
substitution. Introduction of acetyl groups to acid-thinned starches decreased
gelatinization and retrogradation transition parameters and produced very soft gels.

200

7.1. Introduction
Acid-thinning and acetylation are two common techniques for making modified
starches. Acid-thinning plays a key role in manufacture of gum and candies.
Acetylation alters a wide range of functional characteristics of native starches such as
conferring higher peak viscosity and paste clarity and increasing freeze-thawed
stability. Acid-modified starch is usually prepared by acidifying aqueous starch slurry
with dilute acid (HCl, H2SO4, or H3PO3) at a temperature below the gelatinization
point. Acid treatment can modify the starch granules without substantial changes in
the granular form of starch. While acid molecules preferentially attack less compact
amorphous regions, acid hydrolysis can take place at branch points as well as in the
linear segments. It has been reported that at the early stages, the extent of acid
hydrolysis is much greater in amylopectin than in the amylose fraction of maize starch
(Rohwer and Klem, 1984). According to Wang and Wang (2001) acid molecules
primarily attack the amorphous region and both amylose and amylopectin are
simultaneously hydrolyzed. Bertoft (2004) reported that the long chains and shortest
chains of amylopectin molecules are sensitive to lintnerisation, probably because they
are located outside the crystallites.

Acid treatment can markedly alter physicochemical properties of native starches.


Increases have been reported for both transition temperatures and the breadth of the
transition endotherm, however the effect of acid treatment on the gelatinization
enthalpy varied with the starch source, type and concentration of acid and the
hydrolyzing time (Atichokudomchai et al., 2002; Jayakody and Hoover, 2000;
Thirathumthavorn and Charoenrein, 2005; Tang et al., 2001; Jacobs et al., 1998; Muhr
et al., 1984). Low peak viscosity development and accelerated shear-thinning and
higher gelling power were common for acid- treated starch. The gelling power of

201

maize starch was reported to be inversely correlated with the acid concentration
(Wang et al., 2003). The extent of retrogradation as measured by DSC was reported to
increase with progressive acid hydrolysis of tapioca starch (Atichokudomchai et al.,
2002), but amylopectin retrogradation of rice starch stored at 4 C for 7days was
unaffected (Thirathumthavorn and Charoenrein, 2005).

By introducing acetyl groups to native starches, useful functional properties such as


higher peak viscosity, higher paste clarity and improved cold storage can be achieved
(Liu et al., 1997, 1999; Gonzalez and Perez, 2002; Hoover and Sosulski, 1985). The
prime action of substituted acetyl groups is the reduction of starch bond strength
between molecules preventing the inter-chain association (Liu et al., 1997).

Although many investigations have been carried out on the influence of acid-thinning
on starch structural and functional properties very little information is available on the
use of acid-thinning as a pre-or post-treatment for other modifications. Reordering
and self-association of macromolecules resulting after acid-thinning may change the
location of reaction site and thus acid treatment before acetylation may increase the
utility of starch resulting in novel starch properties. One avenue for the production of
novel starch products is the control of the reaction site on the starch macromolecules
(BeMiller, 1997).

7.2. Materials and Methods

7.2.1. Materials
Potato, wheat, and maize starches and fungal -amylase were from Sigma Chemical
Co., (St. Louis, MO, USA). Acetic anhydride was from Merck Co., Germany.
202

7.2.2. Methods

Total amylose content


An amylose/amylopectin assay kit from Megazyme International Ireland Co
Wicklow, Ireland, was used to estimate the total amylose content of all the starches.
This assay is based on the principle of specific formation of amylopectin complex
with Concanavalin (Con A), after a pre-treatment to remove lipids.

Swelling factor
Swelling factor, the ratio of the volume of swollen starch granules to the volume of
dry starch was determined by the method of Tester and Morrison (1990a), when
starch (50 mg, db) was heated at 85 C.

Amylose leaching
Distilled water or solution (10 mL) was added to starch (20 mg, db) in a screw cap
tube. Tubes were then heated at 85 C for 30 min. After cooling to ambient
temperature, samples were centrifuged at 2000 g for 10 min. Amylose content of
supernatant (0.1 mL) was estimated as described by Chrastil (1987).

Differential scanning calorimetry


Gelatinization and dissociation parameters of amylose-lipid complex were measured
using a TA 2920 Modulated DSC Thermal Analyzer differential scanning calorimeter
equipped with a thermal analysis data station (TA Instruments, Newcastle, DE).
Starch (3 mg) was directly measured onto the aluminum DSC pan and distilled water
(9 L) was added with a microsyringe and mixed for homogenization. Pans were
sealed, and allowed to stand for 1 hr at room temperature for even distribution of

203

water. The scanning temperature and the heating rates were 30-120 C and 10 C/min
respectively. An empty pan was used as reference for all measurements.

Pasting properties
Pasting properties of starches were determined using a Rapid Visco-Analyser (RVA)
model 3D (Newport Scientific, Warriewood, Australia). Distilled water (25.5 g) was
added to starch (2.5 g, db) in the RVA canister to obtain a total constant sample
weight of 28 g. The slurry was then manually homogenized using the plastic paddle to
avoid lump formation before the RVA run. A programmed heating and cooling cycle
was set for 22 minutes, where it was first held at 50 C for 1.0 min, heated to 95 C in
7.5 min, further held at 95 C for 5 min, cooled to 50 C within 7.5 min and held at 50
C for 1 min.

Gel hardness
Gel hardness was determined on the starch gel made in the RVA testing using a TAXT2 Texture Analyzer (Stable Micro Systems, Godalming, Surrey, England). After
RVA testing, the paddle was removed and the starch paste in the canister was covered
by Parafilm and stored at 4 C for 24 hr. The gel was compressed at a speed of 0.5
mm/sec to a distance of 10 mm with a 6 mm cylindrical probe. The maximum force
peak in the TPA profile represents the gel hardness.

Retrogradation
After gelatinization test the sample was stored at 4 C for 24 hr to initiate nucleation.
After that samples were kept at 40 C for 10 days before rescanning. Temperature
range and heating rate were 30-120 C and 10 C/min respectively.

204

Acid modification
A 40% (d.b.) starch slurry was acid-modified as described by Singh and Ali (2000)
with slights modifications using 0.1M, 0.5M and 1M hydrochloric acid at 50 C for
1.5 hr. The slurry was stirred frequently during the treatment period, neutralized with
1M NaOH at the end and washed with distilled water repeatedly until the filtrate was
free from acid molecules.

Acetylation
Acetylation was carried out as described by Wang and Wang (2002) with some slight
modifications. Starch (100 g db) was dissolved in distilled water (185 mL) to make
35% slurry. The pH of the slurry was adjusted with 1M NaOH to 8.0-8.5 and then
mechanically stirred for 30 min. Acetic anhydride (8 g) was slowly (dropwise) added
to the slurry while maintaining at pH 8.0-8.5. The reaction was continued for 60 min
before acidifying to pH 5.5 with 1M HCl. The slurry was then washed with three fold
distilled water three times and dried at 35 C.

Determination of degree of substitution


The degree of substitution for the acetylated starches was determined according to
Wurzburg (1964). Blanks with unmodified starches were analyzed concurrently.
% Acetyl = [ml. (Blank) ml. (Sample) normality of acid 0.043 100] / Wt. of
starch sample (g db)
Degree of substitution (D.S.) = 162 % acetyl / 4300 (42 % acetyl)

7.3. Results and Discussion

Swelling factor and amylose leaching


205

Swelling factor and amylose leaching of native, acid-thinned, and acid-thinned


acetylated starches are presented (Table 1). A slight increase of swelling factor was
found for acid-thinned wheat and maize starches from 0.1M HCl treatment but it
decreased with increased acid concentration. For potato starch, swelling factor
decreased in all acid-thinned samples. Potato starch was the most affected starch by
acid treatment. Swelling factor of maize starch was little affected by acid treatment
compared with wheat and potato starches. The disruption of hydrogen bonding
between adjacent starch polymers by acid could increase the swelling ability of starch
granules at low concentration but with greater erosion of the amorphous region at
higher concentration of acid it would decrease the swelling power. Jayakody and
Hoover (2000) observed increased swelling factor for cereal starches at the first stage
of acid hydrolysis. They suggested that interaction between hydrolyzed amylose and
water could increase the swelling ability at the early stage of the treatment. Jane et al.
(1997) postulated that the branch points of B-type amylopectin like in potato starch
are mainly located in the amorphous region making them susceptible to acid
hydrolysis, but branch points of A-type starches amylopectin like those of cereals are
scattered in both amorphous and crystalline regions and thus are less susceptible to
acid hydrolysis. This might be one reason for the observed greater acid hydrolysis in
potato starch.

Apparent amylose content and the extent of amylose leaching were increased in all
the starches after acid-thinning indicating that acid hydrolysis creates more linear
segments that can behave as amylose. The higher the concentration of acid used, the
more the apparent amylose content and the degree of amylose leaching. The highest
content of total amylose resulting in acid-thinned potato starch is expected because of
its greater susceptibility to acid degradation. Perhaps longer average unit chain length

206

of potato starch amylopectin could create more linear segments than those of shorter
length.

As commonly observed, acetylation increased swelling factor of all the starches. By


weakening the bonding forces between starch chains, larger bulky acetyl groups
prevent the inter-chain association and thereby increase the hydration and swelling
power of starch granules. Increased amylose leaching after acetylation may be
attributed to increased starch chain mobility due to the extensive hydration of
acetylated starch granules. Inhibition of inter-chain association could also facilitate
more amylose chains to come out from the starch granule on heating. Introduction of
acetyl groups to acid-thinned starches caused further reduction of swelling factor.
Weaker starch granules due to acid degradation could be further weakened by
acetylation as a result of decreasing bonding forces between adjacent starch chains
and the structurally weaker granules would permit less swelling. After acid
modification substitution was decreased in all the starches. Degradation of amylose
and amylopectin and realignment and self-association of starch macromolecules
during acid treatment could reduce the reaction sites for acetylation.

Gelatinization
DSC curves for gelatinization of native, acid-thinned and acid-thinned acetylated
starches are presented (Fig. 1) and summarized (Table 2). Increases were found for
gelatinization temperature and enthalpy of all the starches after acid-modification.
The effect was more pronounced as the acid concentration was increased. These
increases were greater in potato starch while maize starch was less impact. Acid can
easily hydrolyze more accessible amorphous regions than highly ordered crystalline
regions, consequently eroded amorphous regions would have less impact on

207

destabilizing the crystalline region in the gelatinization process, requiring more


energy to pull the crystals apart (Donovan,1979). Higher transition temperatures of
acid hydrolyzed starch could be due to longer amylopectin double helices resulting
from acid hydrolysis than in unhydrolyzed amylopectin helices, because branch points
in unhydrolyzed amylopectin may reduce the length of helix-forming side chains
Morrison et al. (1993). Jayakody and Hoover (2000) suggested that the higher H
resulting from acid hydrolysis reflects the formation of more double helical starch due
to the interaction between amylose-amylose and amylose-amylopectin chains during
acid hydrolysis. This is quite possible because cleavage of macromolecules by acid
could lead to a greater realignment and self-association of starch macromolecules
forming more double helical-like structures.

Consistent with previous studies, acetylation decreased gelatinization temperature and


enthalpy of all the starches. Substituted acetyl groups inhibit the inter-chain
association of starch chains promoting granule hydration. Extensively hydrated
granules required less energy to reach gelatinization. Acetylation of acid treated
starches reduced the endothermic transition temperature and enthalpy compared to
corresponding acid treated native starches. But the transition temperature and
enthalpy were increased with the increase of acid concentration in acid-thinned
acetylated starches (Table 2). This indicates that the action of acetyl groups has been
weakened in acid treated starches probably due the eroded amorphous region and
greater realignment and self- association of macromolecules resulting from acid
degradation. The reduction of substitution ability after acid treatment could also
contribute to this development. Wheat starch amylose-lipid complex showed
apparently unchanged H for all acid treatments indicating its resistance to acid

208

hydrolysis. The melting temperature of amylose-lipid complex slightly shifted to a


higher temperature with increase of acid concentration (Table 2 and Fig. 2).

Pasting properties
Pasting profiles of native, acid-modified, acetylated, and acid-modified acetylated
starches are presented (Fig. 3) and summarized (Table 3). After acid treatment rapid
reduction of peak viscosity (PV) resulted in all the starches, while onset of viscosity
was earlier in acid-thinned wheat and maize starches but delayed in potato starch (Fig.
3). In dilute starch paste, viscosity is primarily governed by swelling characteristics
and leaching of soluble carbohydrates (mainly amylose). Early onset of PV is
indicative of early and rapid swelling of starch granules. This is consistent with the
observed swelling factor. Thus delayed PV onset of acid-modified potato starch is
consistent with the observed decreased swelling factor after acid-thinning. Erosion of
amorphous region by acid molecules could result in weaker starch granules. Weaker
granules are more deformable at stirring and thus reduce the viscosity. During the
cooling phase, regaining viscosity (CPV) is primarily due to realignment of amylose
chains into a certain level of order (retrogradation). Acid-thinning reduced the cold
paste viscosity. However, by creation of more linear segments as reflected in total
amylose content and greater amylose leaching, the cold paste viscosity should be
increased in acid-thinned starches. However, the decreased CPV may be due to
inadequate time for the cooling phase in RVA programme used. Similar low amylose
retrogradation tendency was reported for acid-modified rice starches which were
subject to insufficient time in the RVA cooling phase (Thirathumthavorn and
Charoenrein, 2005).

209

Acetylation increased PV of all the native starches and caused early onset of viscosity
development. This anticipated consequence is in agreement with the increased
swelling power after acetylation. Extensively hydrated weaker acetylated starch
granules disintegrate more easily at higher temperature resulting a greater shearthinning. However, slightly increased hot paste viscosity (HPV) in acetylated potato
starch may reflect some physical interaction occurring in the collapsed fragile starch
particles. Acetylation increased the CPV of maize and wheat starch but decreased it
potato starch. Substituted acetyl groups could promote creation of junction zones in
amylose molecules facilitating realignment of amylose in the cold starch paste
depending on the molecular characteristics of the amylose.

Introduction of acetyl groups to acid-thinned starch, particularly in higher


concentration acid treated potato and maize starches reduced PV than their
corresponding acid-thinned starches. The action of acetyl groups may further weaken
bonding forces between starch chains in acid-modified starch that have already been
weakened by acid degradation, and thus more weaker granules could undergo rapid
breakdown under shear force limiting viscosity development.

Gel hardness
Gel hardness of native starch followed the order; wheat > maize > potato. Acid
treatment increased the gel hardness of potato starch in all cases. The hardest potato
gel was produced at 1M HCl (Table 1). Gel hardness of wheat starch greatly
decreased after acid-thinning. For maize starch, the hardest gel was produced at 0.1M
HCl but hardness was greatly reduced when acid concentration was increased.
Acetylation greatly reduced the gel hardness of all native starches and the introduction
of acetyl groups to acid-thinned starches produced very soft gels. Gel hardness of all

210

acid-thinned acetylated wheat starches was not measurable under the test conditions.
A similar trend was observed for acetylated 0.5M, and 1M acid-thinned maize and
potato starches. A starch gel can readily form when gelatinized starch or starch paste
undergoes coiling as a result of amylose aggregation. According to Ring (1985) a
starch gel is a composite in which swollen gelatinized starch granules reinforce an
interpenetrating amylose gel matrix. Doublier et al. (1987) suggested that the main
structural parameters involved in starch gelation are the deformability of swollen
starch particle and the amylose concentration of the continuous network. Morris (1990)
explained that starch gel properties relate to the characteristics of the gel matrix, the
amylose, the deformable fillers (swollen granules) that are embedded in the
continuous amylose matrix, the volume fraction of the filler, and the filler-matrix
interaction. Mechanical properties of a starch gel would depend on the rheological
characteristics of the amylose matrix, the volume fraction and the rigidity
(deformability) of the gelatinized granules, and the interactions between the dispersed
and the continuous phases (Eliasson, 1986). Amylose retrogradation depends on the
molecular characteristics and the chain length of the amylose (Gidley, 1990). The
harder acid-modified starch gel therefore could be due to its optimum chain length
with higher concentration of linear segments in the continuous network as evidenced
by greater amylose leaching (Table 1). Softer gel resulting from acid-thinned starches
may be due to the greater loss of granular rigidity (deformability) of swollen starch
particles in the discontinuous media. Perhaps the chain length of amylose segments
after acid degradation may also negatively affect the effectiveness of the aggregation,
although amylose concentration increased after acid-thinning of those starches.

Acetylation decreased the gel hardness of all the starches and caused further
weakening of the gel in acid-modified starches. By inhibiting the intermolecular

211

association of starch chains acetyl groups prevent the close proximity of amylose
chains and thereby hinder amylose aggregation via hydrogen bonding. Softer granules
resulting in acid-thinned acetylated starch could cause a greater reduction of granular
rigidity (deformability) of swollen particles in the discontinuous phase leading to
formation of a very soft starch gel.

Retrogradation
The extent of amylopectin retrogradation in native starches followed the order; potato
> maize > wheat. Reordering of amylopectin in potato and maize starch gel was not
apparently affected after acid modification but it decreased in wheat starch (Table 4
and Fig. 4). It has been shown that chain length of amylopectin with DP 12-22 are
more prone to reassociate during the retrogradation, whereas greater relative
proportion of amylopectin with DP 6-9 and DP > 25 decreased amylopectin
retrogradation enthalpy (Vandeputte et al., 2003). Acid can hydrolyze branch points
of amylopectin and thereby affect the average chain length distribution of
amylopectin whereas the cleavage of macromolecules by acid molecules could
facilitate realignment and self-association of macromolecules forming more double
helical-like structures in starch gel. Thus the melting enthalpy of retrograded acidmodified starch could be attributed to the interplay of the above two factors.
Acetylation decreased amylopectin retrogradation to a greater extent in all the native
starches. No retrogradation was evident for all the wheat starches after acetylation.
This is expected as the bulky acetyl groups attached to amylopectin prevent the
reordering of amylopectin chains in the starch paste. Acetylation also decreased
retrogradation of acid-modified starches but to a lesser extent compared to
corresponding acetylated native starches (Table 4). This tendency was more

212

pronounced when acid concentration is increased for acid-thinning. This is consistent


with the decreased degree of modification after acid treatment.

7.4. Conclusions
Increased total amylose content and degree of amylose leaching reflect creation of
more linear segments during acid modification probably due to the cleavage of
amylopectin branch points. Such cleavage, including amylose chains, facilitates more
realignment and self-association of macromolecules within the starch granules
affecting gelatinization, pasting, and retrogradation properties. Gelling power after
acid-thinning depends on the botanical source of starch and the acid concentration
used for modification. Amylose-lipid complex has greater resistance to acid
degradation. As commonly exhibited, acetylation decreased gelatinization parameters
and retrogradation and increased swelling power, amylose leaching, and peak
viscosity. Decreased substitution in acid-thinned starches shows that acid degradation
reduces the number of reaction sites probably due to the realignment and selfassociation of amylose and amylopectin during acid treatment.

7.5. References
Atichokudomchai N., Varavinit, S., and Chinachoti, P. 2002. A study of annealing
and freeze-thaw stability of acid-modified tapioca starches by differential scanning
calorimetry (DSC). Starch/Starke 54: 343-349.
BeMiller, J.N. 1997. Starch modification: challenges and prospects. Starch/Starke 49:
127-131.
Bertoft, E. 2004. Lintnerization of two amylose-free starches of A-and B-crystalline
types, respectively. Starch/Starke 56: 167-180.
Chrastil, J. 1987. Improved colorimetric determination of amylose in starches or
flours. Carbohydr. Res. 159: 154-158.

213

Donovan, A.M. 1979. Phase transitions of the starch-water system. Biopolymers 18:
263-275.
Doublier, J.L., Llamas, G., and Le Meur, M. 1987. A rheological investigation of
cereal starch pastes and gels. Effects of pasting procedures. Carbohydr. Polym. 7:
251-275.
Eliasson, A.C. 1986. Viscoelastic behavior during the gelatinization of starch. I.
Comparison of wheat, maize, potato, and waxy-barley starches. J. Text. Stud. 17: 253265.
Gidley, M. 1990. Molecular structures and chain length effects in amylose gelation. In:
Gums and Stabilizers for Food the Industry. G.O. Philips, P.A. Williams, D.J.
Wedlock (Eds.), pp. 89-101. Oxford University Press, Oxford.
Gonzalez, Z., and Perez, E. 2002. Effect of acetylation on some properties of rice
starch. Starch/Starke 54: 148-154.
Hoover, R., and Sosulski, F. 1985. A comparative study of the effect of acetylation on
starches of Phaseolus vulgaris biotypes. Starch/Starke 37: 397-404.
Jacobs, H., Eerlingen, R.C., Rouseu, N., Colonna, P., and Delcour, J.A. 1998. Acid
hydrolysis of native and annealed wheat, potato, and pea starches-DSC melting
features and chain length distributions of lintnerised starches. Carbohydr. Res. 308:
359-371.
Jane, J.L., Wong, K.S., and McPherson. A.E. 1997. Branch-structure difference in
starches of A- and B-type X-ray patterns revealed by their Naegeli dextrins.
Carbohydr. Res. 300; 219-227.
Jayakody, L., and Hoover, R. 2000. The effect of lintnerisation on cereal starch
granule. Food Res. Int. 38: 665-680.
Liu, H.J., Ramsden, L., and Corke, H. 1997. Physical properties and enzymatic
digestibility of acetylated ae, wx and normal maize starch. Carbohydr. Polym. 34:
283-289.
Liu, H.J., Ramsden, L., and Corke, H. 1999. Physical properties of cross-linked and
acetylated normal and waxy rice starch. Starch/Starke 51: 249-252.
Morris, M.J. 1990. Starch gelation and retrogradation. Trends in Food Sci. & Technol.
1: 2-6.
Morrison, R., Tester, R.F., Gidley, M.J., and Karkalas, J. 1993. Resistance to acid
hydrolysis of lipid-complexed amylose and lipid-free amylose in lintnerised waxy and
non-waxy barley starches. Carbohydr. Res. 245: 289-302.
Muhr, A.H., Blanshard, J.M.V., and Bates, D.R. 1984. The effect of lintnerisation on
wheat and potato starch granules. Carbohydr. Polym. 4: 399-425.
Ring, S.G. 1985. Some studies on starch gelation. Starch/Starke 37: 80-83.

214

Rohwer, R.G., and Klem, R.E. 1984. Acid-modified starch: production and uses. In:
Starch Chemistry and Technology. R.L. Whistler, J.N. BeMiller, E.F. Paschall (Eds.),
pp. 529-541. Academic Press, Orlando, FL.
Singh, V., and Ali, S.Z. 2000. Acid degradation of starch. The effect of acid and
starch type. Carbohydr. Polym. 41: 191-195.
Tang, H.R., Burn, A., and Hills, B. 2001. A proton NMR relaxation study of the
gelatinisation and acid hydrolysis of native potato starch. Carbohydr. Polym. 46: 7-18.
Tester, R.F., and Morrison, W.R. 1990a. Swelling and gelatinization of cereal starches.
1. Effects of amylopectin, amylose, and lipids. Cereal Chem. 67: 551-557.
Thirathumthavorn, D., and Charoenrein, S. 2005. Thermal and pasting properties of
acid-treated rice starches. Starch/Starke 57: 217-222.
Vandeputte, G.E., Vermeylen, R., Geeroms, J., and Delcour, J.A. 2003. Rice starches.
III. Structural aspects provide insight in amylopectin retrogradation properties and gel
texture. J. Cereal Sci. 38: 61-68.
Wang, L., and Wang, Y.J. 2001. Structures and physicochemical properties of acidthinned corn, potato and rice starches. Starch/Starke 53: 570-576.
Wang, Y.J., and Wang, L., 2002. Characterization of acetylated waxy maize starches
prepared under catalysis by different alkali and alkaline-earth hydroxides. Starch 54:
25-30.
Wang, Y.J., Truong, V.D., and Wang, L. 2003. Structure and rheological properties of
corn starch as affected by acid hydrolysis. Carbohydr. Polym. 52: 327-333.

215

Table 1. Swelling factor (SF), amylose leaching (AL), total amylose content (AMC),
gel hardness (GH), and degree of modification (DS), of native, acid-thinned,
acetylated, and acid-thinned acetylated wheat, potato, and maize starches
Starch
Treatment SF at 85 C AL at 85 C AMC (%) GH (g)
DS
Wheat

Potato

Maize

Native
0.1M
0.5M
1M
AC
0.1M-AC
0.5M-AC
1M-AC

14.80.2
16.90.3
13.80.6
10.30.2
20.80.2
20.20.1
15.40.1
15.20.4

9.20.2
13.30.4
22.50.6
25.80.3
14.40.1
12.90.5
20.10.7
22.90.8

24.30.4
26.10.1
28.20.2
30.10.4

Native
0.1M
0.5M
1M
AC
0.1M-AC
0.5M-AC
1M-AC

50.50.7
42.20.9
28.90.5
21.20.3
55.30.1
38.00.2
26.60.2
16.80.1

15.50.3
18.80.3
25.10.4
30.10.1
16.90.4
17.70.4
23.90.6
28.10.4

24.70.2
25.80.2
28.10.2
34.30.5

Native
0.1M
0.5M
1M
AC
0.1M-AC
0.5M-AC
1M-AC

20.40.4
21.70.4
20.20.8
18.50.3
22.50.5
20.80.1
20.30.3
17.50.5

14.10.4
15.90.5
23.50.3
25.80.7
15.10.3
15.30.8
21.20.2
23.10.1

23.40.3
25.10.4
27.20.7
29.10.3

801.4
300.3
280.9
131.1
270.8
-

0.143
0.136
0.130
0.123

360.7
380.7
781.2
561.3
180.9
140.8
-

0.138
0.127
0.120
0.114

581.5
631.4
270.8
240.5
231.1
170.8
-

0.132
0.128
0.124
0.119

AC=acetylated.
1M, 0.5M and 1M=concentrations of HCl.
M-AC=acid-thinned acetylated.
Values are mean of triplicate determination standard deviation. For gel hardness
values are at least means of duplicate determinations standard deviation.

216

Table 2. Gelatinization parameters of native, acid-thinned, acetylated, and acidthinned acetylated wheat, potato, and maize starches
Starch
Treatment
To (C)
Tp (C)
Tc (C)
H (J/g)
Wheat

Native
0.1M
0.5M
1M
AC
0.1M-AC
0.5M-AC
1M-AC

58.00.6
62.00.4
62.90.9
64.20.2
55.20.6
57.80.6
59.30.4
60.00.3

63.00.4
65.30.3
66.2o.2
68.00.3
60.10.4
62.00.8
63.50.9
64.51.1

74.50.1
75.00.2
76.20.5
78.50.2
70.30.2
70.60.5
73.50.2
74.60.1

9.60.3
9.90.2
9.90.2
11.10.1
8.50.4
8.30.2
9.20.4
9.40.1

Potato

Native
0.1M
0.5M
1M
AC
0.1M-AC
0.5M-AC
1M-AC

60.00.2
63.70.8
65.10.7
67.00.3
56.20.5
59.20.1
60.50.4
62.00.6

65.00.4
67.40.5
69.00.3
71.10.4
61.80.2
64.00.3
65.40.7
67.10.4

76.10.1
78.00.4
80.00.3
82.00.2
73.10.7
76.10.8
76.50.6
80.10.2

15.20.2
15.60.3
15.90.3
16.90.4
14.50.5
15.20.1
15.30.3
15.70.1

Maize

Native
0.1M
0.5M
1M
AC
0.1M-AC
0.5M-AC
1M-AC

66.10.2
67.20.4
67.50.2
67.80.2
62.50.4
62.90.2
63.20.4
63.40.1

70.70.1
71.00.2
71.40.3
71.90.2
67.80.6
68.60.2
68.30.5
68.70.1

80.20.3
80.20.2
82.60.5
82.80.3
77.10.8
78.80.1
78.50.2
78.50.2

11.50.3
11.70.3
11.90.4
12.70.5
10.10.3
10.20.4
10.50.1
10.60.2

Wheat
(AMLC)

Native
93.40.3
98.50.4
104.10.4
0.780.09
0.1M
93.70.4
98.70.2
104.10.5
0.750.06
0.5M
95.20.2
100.30.1
105.60.3
0.800.01
1M
95.30.3
101.20.1
105.90.1
0.750.07
Gelatinization parameters; To=onset, Tp=peak, and Tc=conclusion;
H=gelatinization enthalpy; AC=acetylated; AMLC=amylose-lipid complex.
1M, 0.5M, and 1M=concentrations of HCl.
M-AC=acid-thinned acetylated.
All values are means of triplicate determinations standard deviation.

217

Table 3. Pasting properties of native, acid-thinned, acetylated, and acid-thinned


acetylated wheat, potato, and maize starches
Starch
Treatment PV
HPV
BD
CPV
SB
Wheat

Native
0.1M
0.5M
1M
AC
0.1M-AC
0.5M-AC
1M-AC

1531.2
1161.5
120.8
60.6
2070.9
1271.2
150.6
50.4

1141.4
320.8
5.00.7
30.6
1021.5
250.8
40.4
20.7

390.8
841.1
7.00.8
30.4
1071.3
1011.6
101.7
30.6

2161.6
691.7
141.2
80.8
2181.1
701.4
80.3
50.2

1011.4
360.7
90.4
50.4
1171.3
441.4
40.8
20.4

Potato

Native
0.1M
0.5M
1M
AC
0.1M-AC
0.5M-AC
1M-AC

6551.4
4161.3
920.9
300.7
7241.6
4761.9
811.4
240.9

1741.7
1141.3
220.7
60.5
630.8
1300.4
140.9
40.5

4811.9
3021.7
701.8
240.7
6610.4
3460.7
671.1
200.3

2541.8
1851.7
351.2
80.8
2451.2
1881.4
201.6
50.8

811.1
710.7
131.2
20.6
1821.5
581.1
60.4
20.2

Maize

Native
1941.1
1631.4
900.8
2101.8
1071.6
0.1M
1700.7
401.1
1291.5
100.6
581.1
0.5M
440.6
60.8
371.1
140.7
80.5
1M
110.8
50.7
60.7
80.2
30.2
AC
1990.9
920.6
1070.8
2521.8
1601.4
0.1M-AC 1310.5
470.4
840.5
1152.2
680.9
0.5M-AC 300.8
600.7
2406
12.7
50.4
1M-AC
100.6
50.7
50.5
70.4
20.5
PV=peak viscosity; HPV=hot paste viscosity; BD=breakdown; SB setback.
AC=acetylated.
0.1M, 0.5M and 1M=concentrations of HCl.
M-AC=acid-thinned acetylated.
Values are at least means of duplicate determinations standard deviation.

218

Table 4. Melting enthalpy (HR) of retrograded amylopectin starch crystals of


native, acid-thinned, acetylated, and acid-thinned acetylated starches
Starch
Treatment
HR (J/g)
wheat

Native
0.1M
0.5M
1M
AC
0.1M-AC
0.5M-AC
1M-AC

0.60.1
0.30.05
0.20.03
-

Potato

Native
0.1M
0.5M
1M
AC
0.1M-AC
0.5M-AC
1M-AC

4.60.1
4.40.2
4.50.3
4.60.01
1.80.02
2.20.04
2.40.07
3.00.02

Maize

Native
0.1M
0.5M
1M
AC
0.1M-AC
0.5M-AC
1M-AC

1.20.02
1.30.01
1.20.1
0.80.02
0.20.01
0.40.05
0.60.02
0.70.04

AC=acetylated.
0.1M, 0.5M, and 1M=concentrations of HCl.
M-AC=acid-thinned acetylated.
Values are means of triplicate determinations standard deviation.

219

-1.8

-2.1

-2.1
-2.3

-2.0

-2.5

-2.4

g)

-2.7
-2.9

Heat flow (W/g)

-2.2

-2.3

-3.1

-2.5

-2.7

-3.3

-2.6

-3.5

-2.9

-2.8

-3.7
-3.9

-3.0

-4.1

Wheat
-3.2

45

55

-3.1

65

75

85

-4.3

Potato
40

45

50

Maize
55

60

65

70

75

80

85

90

-3.3

40

45

50

55

60

65

70

75

80

85

90

95

100

Temperature (C)

Fig. 1. DSC curves of native, acid-thinned, acetylated, and acid-thinned acetylated wheat, potato and maize starches. N=native; ac=acetylated;
M-ac=acid-thinned acetylated.

220

-1.9

Native

0.1M acid

Heat flow (W/g)

-2.1

0.5M acid

-2.3

1M acid

-2.5
Amylose-lipid complex
Gelatinization

-2.7
40

60

80
Temperature (C)

100

120

Fig. 2. DSC curves of native and acid-thinned wheat starch. Large and small
endotherms represent the gelatinization endotherm and amylose-lipid
complex endotherm respectively.

221

100

100

100

ac

240

240
80
ac

80
n

600
180

0.1M-ac
60

60

60

0.1M

0.1M

400
120

120
40

0.1Mac

40

0.1M

0.5
Mac

60

40

Temperature (C)

180

Viscosity (RVU)

ac

80

200
20

0.5M

0.1M-ac

60

0.5
M

20

20
0. 5M

0.5M-ac
0

10

15

20

0
25

10

15

0.5M-ac

20

0
25

10

15

20

0
25

Time (Mins)
Fig. 3. RVA curves of native, acid-thinned, acetylated, and acid-thinned acetylated wheat (A), potato (B), and maize (C) starches. n=native,
ac=acetylated, 0.1M, 0.5M, and 1M= acid concentration used for the modification, 0.1M-ac and 0.5M-ac=acid-thinned acetylated.

222

-1.7

n
0.1M

-1.9

0.5M

Heat flow (W/g)

1M

ac

-2.1

0.1M-ac

0.5M-ac

-2.3

1M-ac

-2.5

50

60

70

80

90

100

Temperature (C)
Fig. 4. DSC curves of retrograded native, acid-tinned, acetylated, and acidthinned acetylated potato starches. n=native, ac=acetylated, 0.1M, 0.5M, and
1M =acid concentrations used for the modification.

223

Chapter 8
Effect of Hydroxypropylation on Some Structural and
Physicochemical Properties of Heat-Moisture Treated Wheat, Potato
and Waxy Maize Starches
Abstract
Hydroxypropylation was carried out on heat-moisture treated wheat, potato and waxy
maize starches. Hydroxypropylation increased swelling factor and amylose leaching
of both native and heat-moisture treated starches. Hydroxypropylation of heatmoisture treated starches reduced enthalpies of gelatinization and amylopectin
retrogradation more than for the corresponding hydroxypropylated native starches.
This indicates that heat-moisture treatment increases the derivatization of
amylopectin. Disruption and reorientation of amylopectin double helices during heatmoisture treatment could facilitate the access of reaction reagent to the highly ordered
crystalline regions resulting in greater derivatization. Cold paste viscosity was greatly
increased with high pasting stability when hydroxypropyl groups were introduced to
heat-moisture treated wheat and potato starches. Alkaline treatment increased
gelatinization temperature of all the starches but enthalpy was unaffected. Amylose
leaching and swelling factor greatly increased in wheat but decreased in potato and
waxy maize starches by alkaline treatment. This increased amylose leaching and
swelling factor along with greater reduction of amylose-lipid complex endotherm of
wheat starch by alkaline treatment is consistent with the disruption of amylose-lipid
complex. The amylose-lipid complex is susceptible to hydrolysis in alkaline
conditions. Heat-moisture treatment had no influence on transition parameters of
amylose-lipid complex.

224

8.1. Introduction
Modified starches are important functional ingredients in processed foods.
Hydroxypropylation is commonly used for making modified starch in food industry.
Hydroxypropyl groups are hydrophilic in nature and when introduced into starch
granules weaken the internal bond structure holding granules together. The
substituent disturbs the association of the polysaccharide chains preventing
retrogradation due to the hydrogen bonds. Improved functional properties of
hydroxypropylated starches such as extended shelf life of cold storage products
(freeze-thaw stability), higher peak viscosity and paste clarity, and decreased
gelatinization temperatures are well documented (Hoover et al., 1988; Kim and
Eliasson, 1993; Perera et al., 1997; Liu et al., 1999; Pal et al., 2002). By light
microscopic studies Kim et al. (1992) claimed that hydroxypropylation mainly takes
place in the central region of potato starch granule. Biliaderis (1982) reported that
modification occurs throughout the granules. Gray and BeMiller (2005) believed that
derivatization first occurs in the most accessible amorphous region and proceeds
through various regions until it reaches the highly organized crystalline region. Shi
and BeMiller (2000, 2002) for normal maize starch, and Kavitha and BeMiller (1998)
for potato starch showed that derivatization occurs more in the amylose chains than in
amylopectin. According to Shi and BeMiller (2002) derivatization facilitates amylose
leaching but preferential leaching of derivatized amylose decreased as the MS of the
whole granule increased.

Perera et al. (1997) showed that alkaline treatment (NaOH, and Na2SO4) during
hydroxypropylation disrupts the double helices within the amorphous region altering
crystalline orientation of the starch granule. It has been reported that granule swelling
is essential in order for the substitution reaction to take place in granular starch

225

(Hauber et al., 1992). Alkaline treatment could induce swelling by ionizing starch
hydroxyl groups at higher pH (Gray and BeMiller, 2005).

In a study on hydroxypropylation of heat-moisture treated potato starch Perera et al.


(1997) reported that heat-moisture treatment enhanced the access of hydroxyl groups
to the amorphous region. However, systematic studies to characterize the influence of
physical modification (eg., heat-moisture treatment) on the derivatization process
have not been reported. Comparative study of the derivatization of native starch with
heat-moisture treated starch may reveal more information on the effect of
derivatization on the structural and functional properties. Structural changes within
amorphous region and crystalline region of starch granules such as starch chain
interaction within the amorphous region and disruption and reorientation of starch
crystallites caused by heat-moisture treatment has been discussed by many
researchers (Hoover et al., 1994; Gunaratne and Hoover, 2002; Hoover and
Vasanthan, 1994). Derivatization of structurally altered starch granules may require
less chemical to achieve desirable chemical bonding and target functionalities
probably by affecting the accessibility to reaction reagents and the reaction site. For
example, destabilization and reorientation of amylopectin double helices could allow
more derivatization to take place within the crystalline region by facilitating access of
reaction reagents to the crystalline region. BeMiller (1997) reported that one avenue
for the development of novel starch products is the control of the location of the
reaction within the starch granules. The objective of this study was therefore to
characterize some structural and functional properties when hydroxypropyl groups
are introduced to heat-moisture treated starches and to investigate the influence of the
alkaline treatment (reaction conditions used during hydroxypropylation) on some
structural and functional properties.

226

8.2. Materials and Methods

8.2.1. Materials
Potato starch, wheat starch, waxy maize starch and propylene oxide were from Sigma
Chemical Co., (St. Louis, MO, USA).

8.2.2. Methods

Swelling factor
Swelling factor, the ratio of the volume of swollen starch granules to the volume of
dry starch was determined by the method of Tester and Morrison (1990a), where
starch (50 mg, db) was heated at 85 C for 30 min in 5 mL water.

Amylose leaching
Distilled water or solution (10 mL) was added to starch (20 mg, db) in a screw cap
tube. Tubes were then heated at 85 C for 30 min. After cooling to ambient
temperature, samples were centrifuged at 2000 g for 10 min. Amylose content of
supernatant (0.1 mL) was estimated as described by Chrastil (1987).

Differential scanning calorimetry


Gelatinization and dissociation parameters were measured using a TA 2920
Modulated DSC Thermal Analyzer differential scanning calorimeter equipped with a
thermal analysis data station (TA Instruments, Newcastle, DE). Starch (3 mg) was
directly measured on to the aluminum DSC pan and distilled water (9 L) was added
with a microsyringe and mixed for homogenization. Pans were sealed, and allowed to

227

stand for 1 hr at room temperature for even distribution of water. The scanning
temperature and the heating rates were 30-120 C and 10 C/min respectively. An
empty pan was used as reference for all measurements.

Pasting properties
Pasting properties of starches were determined using a Rapid Visco-Analyzer (RVA)
model 3D (Newport Scientific, Warriewood, Australia). Distilled water or
hydrochloric acid solution (25.5 g) was added to starch (2.3 g, db) in the RVA
canister to obtain a total constant sample weight of 27.8 g (8.2% starch
concentration). The slurry was then manually homogenized using the plastic paddle to
avoid lump formation before the RVA run. A programmed heating and cooling cycle
was set for 22 minutes, where it was first held at 50 C for 1.0 min, heated to 95 C in
7.5 min, further held at 95 C for 5 min, cooled to 50 C within 7.5 min and held at
50 C for 1 min.

Gel hardness
Gel hardness was determined on the starch gel made in the RVA testing using a TAXT2 Texture Analyzer (Stable Micro Systems, Godalming, England). After RVA
testing, the paddle was removed and the starch paste in the canister was covered by
Parafilm and stored at 4 C for 7 hrs. The gel was compressed at a speed of 0.5
mm/sec to a distance of 10 mm with a 6 mm cylindrical probe. The maximum force
peak in the TPA profile represents the gel hardness.

Hydroxypropylation
Hydroxypropylation of starches was performed according to the method of Choi and
Kerr (2004) with some slight modifications. Starch sample (50 g db) was suspended

228

in distilled water (110 mL) containing 10 g Na2SO4 in a centrifuge bottle. After


adjusting pH to 11.3 with 1M NaOH, 1.5, 3.0, 4.5 mL of propylene oxide was added
and the bottle was immediately capped and shaken vigorously. Sample was then
placed at 35 C in shaking water bath with continuous shaking for 24 hr. The reaction
was terminated by adjusting pH to 5.3 with 1M HCl. Slurry was then centrifuged at
3000 g for 10 min and recovered starch cake was washed with distilled water and
dried at 35 C. Degree of modification was determined by the method of Johnson
(1969).

Retrogradation
Starch gel was prepared in the DSC pan using 1:1 starch to water ratio scanning the
sample to 120 C. Samples were then stored at 4 C for 24 hr to initiate nucleation.
After that samples were kept at 40 C for 7 days before rescanning by DSC.
Temperature range and heating rate were 30-120 C and 10 C/min respectively.

8.3. Results and Discussion

Swelling factor and Amylose leaching


Swelling factor and amylose leaching of native and modified starches are presented in
Table 1. As commonly observed, heat-moisture treatment decreased swelling ability
of all tested starches while hydroxypropylation increased it. Additional chain
interaction and loss of some of the double helices during heat-moisture treatment
could restrict the swelling capacity (Hoover and Vasanthan, 1994; Gunaratne and
Hoover, 2002). Increased granular swelling after hydroxypropylation is expected
result because increased hydrophilicity and loosened starch structure resulting from

229

hydroxypropylation could increase the swelling factor by promoting water


penetration into the granule.

Alkaline treatment (under the conditions prevailing during hydroxypropylation, but in


the absence of propylene oxide) [NaOH, and Na2SO4, heating at 40 C for 24 h]
increased the swelling factor of wheat starch but slightly decreased it in potato and
waxy maize starches. In the hydroxypropylation reaction medium, many factors could
affect granular swelling such as pH, swelling inhibiting salt types and their
concentration, and starch concentration. By ionizing hydroxyl groups of starch chains,
increased pH could disrupt the hydrogen bonds and thereby increase the granular
swelling (Gray and BeMiller, 2005). It has been reported that swelling must first
occur in the granular starch in order for the derivatization process to proceed (Hauber
et al., 1992). Confirming this claim, Shi and BeMiller (2000) reported that replacing
of Na2SO4 by the more powerful swelling inhibitor potassium citrate requires 1.8
times more propylene oxide to achieve the same degree of modification. Alkaline
treatment greatly increased the amylose leaching of wheat starch but slightly
decreased it in potato and waxy maize starches. DSC curves of alkaline-treated wheat
starch showed loss of amylose-lipid complex to a greater extent (Fig. 1). It is well
known that the existence of amylose-lipid complex reduces the amylose leaching and
swelling ability of starch granule thus the cleavage of amylose-lipid complex at
higher alkalinity could be the reason for resulting higher amylose leaching and
swelling ability of alkaline-treated wheat starch. For potato and waxy maize starches
amylose leaching was decreased by alkaline treatment. Hydroxypropylation increased
the amylose leaching of all the starches. Hydroxypropyl groups disrupt the hydrogen
bonding of adjacent starch chains promoting their flexibility. This could facilitate the
mobility of amylose chains.

230

Gelatinization
Decreases were observed for gelatinization temperature and enthalpy of all native and
heat-moisture treated starches after hydroxypropylation. As commonly observed,
heat-moisture treatment increased gelatinization temperature while decreasing the
enthalpy. Alkaline treatment increased gelatinization temperature of all the starches
but enthalpy was unaffected (Table 2). Delayed gelatinization after alkaline treatment
could be due to the influence of a swelling inhibiting salt (Na2SO4). Salts like Na2SO4
have been shown to increase the gelatinization temperature and to decrease the
leaching of starch polysaccharide chains (Oosten, 1990). Consistent with previous
studies, increased granular swelling after hydroxypropylation reduces energy
requirement for gelatinization.

Introduction of hydroxypropyl groups to heat-moisture treated starches decreased


gelatinization enthalpy to a higher extent compared with corresponding
hydroxypropylated native starches. Gray and BeMiller (2005) reported that in the
derivatization process, reagent solution enters most readily to the less ordered
amorphous region derivatizing it first, and that derivatization could further swell
granules by opening the crystalline region permitting access of the reagent solution. It
seems that disruption of starch crystallites and reorientation of amylopectin double
helices by heat-moisture treatment allowed more reaction reagents to access the
crystalline region increasing derivatization in that region. More derivatization could
disrupt more hydrogen bonds in the crystalline region, subsequently it would require
less energy to disrupt the crystalline region during gelatinization. Heat-moisture
treatment increased the degree of substitution (Table 1). Similar results were reported
for the hydroxypropylation of heat-moisture treated potato starch (Perera et al., 1997).
According to them, heat-moisture treatment increased the amount of amorphous

231

region within the granular interior facilitating the derivatization. In this study, as
described above heat-moisture treatment facilitates more derivatization to take place
in the crystalline region. This may also contribute to increase the degree of
modification.

Pasting properties and gel hardness


Pasting curves of native, heat-moisture treated, hydroxypropylated native and
hydroxypropylated heat-moisture treated starches are presented (Fig. 2 a, b & c) and
results summarized (Table 3). As commonly observed, hydroxypropylation decreased
pasting temperature but increased peak viscosity in all the native starches. Heatmoisture treatment increased pasting temperature and pasting stability to a greater
extent whereas hydroxypropylation decreased pasting stability of all native starches.
These are expected pasting characteristics of hydroxypropylated starches because
higher swelling and amylose leaching resulting from the loosened hydroxypropylated
granule will allow granules to reach higher peak viscosity, but structurally weaker
granules will disintegrate rapidly at higher temperature resulting in rapid loss of
viscosity. By restricting swelling and amylose leaching, heat-moisture treatment
could increase the pasting temperature while decreasing peak viscosity. Higher
pasting stability of heat-moisture treated starches is consistent with the greatly
reduced granule breakdown during the pasting process. Similar to native starches,
introduction of hydroxypropyl groups decreased the pasting onset temperature of
heat-moisture treated starches.

The most interesting pasting characteristic observed for the hydroxypropylated heatmoisture treated wheat and potato starches is the increase of cold paste viscosity
(CPV) and setback (SB) to a greater extent while maintaining higher pasting stability

232

(Table 3 and Fig 2a & b). Thus the introduction of hydroxypropyl groups to heatmoisture treated starches could be applied for products which require higher end
viscosity. It is well known that when gelatinized starch paste is subjected to cooling
the extent of viscosity increase is mainly governed by the rapid reassociation of linear
amylose chains via formation of gel matrix. In case of potato starch, heat-moisture
treatment decreased amylose leaching to a greater extent. Alkaline treatment or
hydroxypropylation also did not much affect the magnitude of the amylose leaching
of heat-moisture treated potato starch (Table I) but still cold paste viscosity and
setback were greatly increased in hydroxypropylated heat-moisture treated potato
starch. This indicates that the concentration of amylose in the continuous media does
not play a key role in the increase of final viscosity. Therefore, it appears that greater
extent of derivatization in amylopectin polymers in heat-moisture treated starch as
concluded before could promote more interactions between amylopectin rich filler
component and continuous gel matrix. This could increase the CPV. Control samples
of wheat and potato starches also showed a substantial increase of cold paste viscosity.
This was probably due to the influence of swelling inhibiting salt (Na2SO4) in the
formation of gel matrix. Apparently unchanged CPV and SB of hydroxypropylated
heat-moisture treated waxy maize starch are expected results as it is virtually amylose
free.

While hydroxypropylation produced soft gel, heat-moisture treatment increased the


gel hardness. Alkaline treatment slightly increased the gel hardness of potato and
waxy maize starches but it decreased in wheat starch proving that weaker wheat
starch granules resulting due to the disruption of amylose-lipid complex during
alkaline treatment and subsequent hydration reduce the stiffness of the gel.
Hydroxypropylation caused decreased gel hardness of heat-moisture treated wheat

233

and potato starch. At higher degree of modification both native and heat-moisture
treated waxy maize and potato native starch produced very soft gel which was not
measurable under the given condition (Table 1). In the starch gel, interchain
association of exuded amylose chains from the swollen granules surround the
gelatinized starch granules, and thus starch gel can be regarded as a hydrated polymer
composites where swollen amylopectin-rich granules are embedded in and reinforce a
continuous matrix of entangled amylose molecules (Ring, 1985). The textural and
mechanical properties of starch gel where swollen gelatinized amylopectin-rich
granules are embedded in the continuous amylose gel matrix would depend on the
amylose concentration and rheological characteristics of the amylose gel matrix,
rigidity (deformability) of the swollen starch particles, volume fraction of the swollen
granules, and interaction between swollen particles and amylose matrix (Morris, 1990;
Doublier et al., 1987; Eliasson, 1986). Weaker, fragile less rigid swollen granules
resulting from hydroxypropylation could produce soft gel. Eerlingen et al. (1997)
found that increased or decreased gel storage moduli could result based on the heatmoisture treatment conditions and the concentration of the starch gel, as these affect
the extent of swelling, solubility and close packing concentration. Hoover et al. (1994)
reported that restriction of swelling by HMT favors amylose aggregation in the
formation of the gel matrix leading to firmer gels after HMT.

Retrogradation
Hydroxypropylation decreased amylopectin retrogradation as expected in all the
starches. The greater the degree of modification, the more the retrogradation
decreased. By inhibiting interchain association, hydroxypropyl groups substituted to
amylopectin chains could decrease the extent of reassociation or retrogradation.
Inclusion of same propylene oxide volume in the derivatization reaction resulted in

234

more inhibition of amylopectin retrogradation in heat-moisture treated starch


compared to their corresponding native starches (Table I and Fig. 3). As I concluded
before, disruption and reorientation of amyloepectin double helices during heatmoisture treatment facilitates more reaction reagent to access the crystalline region
causing greater modification of amylopectin. More derivatized amylopectin could
reduce the interchain association of amylopectin to a greater extent compared with
native starch decreasing amylopectin retrogradation. Alkaline treatment showed no
influence on amylopectin retrogradation in all the starches, but heat-moisture
treatment resulted in amylopectin retrogradation of potato starch while other starches
were unaffected. As I previously reported longer amylopectin chains of potato starch
(B-type starch) could be more susceptible to the degradation during heat-moisture
treatment reducing the average unit chain length. This may weaken the lateral
association of amylopectin in the retrogradation process.

8.4. Conclusions
In the derivatization reaction, alkaline treatment had a greater influence on properties
of wheat starch such as higher amylose leaching, swelling, and early onset of peak
viscosity. This seems highly related to the disruption of amylose-lipid complex at
higher pH (11.4) provided in the reaction condition. Destabilization and reorientation
of amylopectin double helices during heat-moisture treatment facilitates more
modification in the crystalline region. Thus heat-moisture treatment could be
employed to alter the location of the derivatization reaction while achieving novel
starch properties such as higher final paste viscosity with greater pasting stability.
Increased degree of modification after heat-moisture treatment indicates the increased
reaction efficiency. Thus it would require less concentration of reaction reagents to
achieve the desired degree of chemical bonding and improved functional
235

characteristics. Characterization of hydroxypropylation of heat-moisture treated


starches which are treated with different heat and moisture combinations, and
determination of the best combination that best mimics the desired behavior with
minimum chemical bonding, is important.

8.5. References
BeMiller, J.N. 1997. Starch modification: challenges and prospects. Starch/Starke 49:
127-131.
Biliaderis, C.G. 1982. Physical characteristics, enzymatic digestibility, and structure
of chemically modified smooth pea and waxy maize starches. J. Agric. Food Chem.
30: 925-930.
Choi, S.G. and Kerr, W.L. 2004. Swelling characteristics of native and chemically
modified wheat starches as a function of heating temperature and time. Starch 56:
181-189.
Chrastil, J. 1987. Improved colorimetric determination of amylose in starches or
flours. Carbohydr. Res. 159: 154-158.
Doublier, J.L., Llamas, G., and Le Meur, M. 1987. A rheological investigation of
cereal starch pastes and gels. Effects of pasting procedures. Carbohydr. Polym. 7:
251-275.
Eerlingen R.C., Jacobs, H., Block, K., and Delcour, J.A. 1997. Effects of
hydrothermal treatments on the rheological properties of potato starch. Carbohydr.
Res. 297: 247-356.
Eliasson, A.C. 1986. Viscoelastic behavior during the gelatinization of starch. .
Comparison of wheat, maize, potato, and waxy barley starches. J. Text. Stud. 17: 253265.
Gray, J.A., and BeMiller, J.N. 2005. Influence of reaction conditions on the location
of reactions in waxy maize starch granules reacted with a propylene oxide analog at
low substitution levels. Carbohydr. Polym. 60: 147-162.
Gunaratne A. and Hoover, R. 2002. Effect of heat-moisture treatment on the structure
and physicochemical properties of tuber and root starches. Carbohydr. Polym. 49:
425-437.
Hauber, R., BeMiller, J.N., and Fannon, J.E. 1992. Swelling and reactivity of maize
starch granules. Starch/Starke 44: 323-327.

236

Hoover, R., Hannouz, D., and Sosulski, F.W. 1988. Effect of hydroxypropylation on
thermal properties, starch digestibility and freeze-thaw stability of field pea (Pisum
sativum cv Trapper) starch. Starke 40: 383-387.
Hoover, R. Vasanthan T., Senanayake, N.J., and Martin, A.M. 1994. The effect of
defatting and heat-moisture treatment on the retrogradation of starch gels from wheat,
oat, potato, and lentil. Carbohydr. Res. 261: 13-24.
Hoover, R., and Vasanthan, T. 1994. Effect of heat-moisture treatment on the
structure and physicochemical properties of cereal, legume, and tuber starches.
Carbohydr. Res. 252: 33-53.
Johnson, D.P. 1969. Spectrophotometric determination of the hydroxypropyl groups
in the starch ethers. Analytical Chemistry 41: 859-860.
Kavitha, R., and BeMiller, J.N. 1998. Characterization of hydroxypropylated potato
starch. Carbohydr. Polym. 37: 115-121.
Kim, H.R., and Eliasson, A.C. 1993. The influence of molar substitution on thermal
transition properties of hydroxypropyl potato starches. Carbohydr. Polym. 22:331335.
Kim, H.R., Hermansson, A.M., and Eriksson, C.E. 1992. Structural characteristics of
hydroxypropyl potato starch granules depending on their molar substitution.
Starch/Starke 44: 111-116.
Liu, H.J., Ramsden, L., and Corke, H. 1999. Physical properties and enzymatic
digestibility of hydroxypropylated ae, wx, and normal maize starches. Carbohydr.
Polym. 40: 175-182.
Morris, M.J. 1990. Starch gelation and retrogradation. Trends in Food Science &
Technology 1: 2-6.
Oosten, B.J. 1990. Interactions between starch and electrolytes. Starch/Starke 42:
327-330.
Pal, J., Singhal, P.S., and Kulkarni, P.R., 2002. Physicochemical properties of
hydroxypropyl derivative from corn and amaranth starch. Carbohydr. Polym. 48: 4953.
Perera, C., Hoover, R., and Martin, A.M. 1997. The effect of hydroxypropylation on
the structure and physicochemical properties of native, defatted and heat-moisture
treated potato starches. Food Research International 30: 235-247.
Shi, X., and BeMiller, J.N. 2000. Effect of sulfate and citrate salts on derivatization of
amylose and amylopectin during hydroxypropylation of corn starch. Carbohydr.
Polym. 43: 333-336.
Ring, S.G. 1985. Some studies on starch gelation. Starch 37: 80-83.
Shi, X., and BeMiller, J.N. 2002. Aqueous leaching of derivatized amylose from
hydroxypropylated common corn starch granules. Starch/Starke 54: 16-19.
237

Tester, R.F., and Morrison, W.R. 1990a. Swelling and gelatinization of cereal
starches. 1. Effects of amylopectin, amylose, and lipids. Cereal Chem. 67: 551-557.

238

Table 1. Swelling factor (SF), amylose leaching (AMYL), gel hardness (GH),
retrogradation enthalpy (HR), and degree of modification (DS) of
hydroxypropylated native and heat-moisture treated hydroxypropylated wheat,
potato and waxy maize starches
Starch Treatmenta
SF at
AMYL (%) GH (g) HR
DS
(J/g)
85 C
at 85 C
Wheat Native
15.00.2b 9.50.1
760.4 4.40.1 Control
19.20.1 16.70.2
730.6 4.50.2 HP1.5
21.10.1 19.30.3
380.2 2.80.1 0.023
HP3.0
27.50.2 20.60.1
280.4 1.90.2 0.042
HP4.5
29.00.3 21.80.2
200.1 1.50.1 0.127
HMT
Control
HMT-HP1.5
HMT-HP3.0
HMT-HP4.5

10.80.1
12.10.1
17.80.3
19.40.2
22.10.4

6.40.1
13.10.4
14.70.3
16.50.1
18.20.1

1000.4
1040.3
550.5
230.1
230.2

4.20.3
4.20.1
2.20.1
1.10.2
0.80.1

0.024
0.049
0.139

49.10.2
47.20.1
56.60.4
59.30.1
64.00.2

15.00.3
12.20.1
12.80.2
13.70.3
13.80.1

400.1
450.3
240.4
-

8.50.0
8.40.1
6.70.3
5.00.1
4.20.2

0.027
0.047
0.139

HMT
Control
HMT-HP1.5
HMT-HP3.0
HMT-HP4.5

11.30.2
8.40.1
14.60.1
17.00.3
18.10.1

3.80.1
3.40.0
3.70.1
3.70.1
3.60.1

450.6
520.2
160.2
120.5
100.1

7.20.1
7.30.2
5.00.2
4.20.1
3.00.2

0.029
0.055
0.154

Native
Control
HP1.5
HP3.0
HP4.5

38.00.3
34.10.1
37.70.3
39.10.2
39.80.5

2.10.0
2.00.2
1.90.1
1.40.1
1.20.2

80.4
90.3
-

8.70.3
8.60.1
7.60.2
6.90.1
3.10.2

0.019
0.038
0.098

Potato Native
Control
HP1.5
HP3.0
HP4.5

Waxy
maize

HMT
34.00.1 1.70.1
90.3
8.80.3 Control
28.30.1 1.80.0
90.7
8.70.3 HMT-HP1.5 34.10.3 1.20.1
7.20.2 0.024
HMT-HP3.0 34.40.1 1.30.1
5.80.1 0.043
HMT-HP4.5 36.70.2 1.20.0
2.20.1 0.109
a
HMT=heat-moisture treated, HP=hydroxypropylated, HMT-HP=heat-moisture
treated hydroxypropylated, 1.5, 3.0, and 4.5=volume (mL) of propylene oxide.
b
Values are means of triplicate determinations standard deviations.

239

Table 2. Gelatinization parametersa of hydroxypropylated native and heatmoisture treated hydroxypropylated wheat, potato, and waxy maize starches
Starch
Treatment b
To (C)
Tp (C)
Tc (C)
H (J/g)
c
Wheat
Native
58.90.1
64.30.1
76.80.5
10.40.1
Control
61.50.2
65.70.1
79.00.1
10.30.3
HP1.5
58.00.2
62.60.3
75.50.3
9.70.2
HP3.0
56.70.3
60.60.2
71.20.1
8.50.2
HP4.5
53.60.1
58.50.4
69.80.3
7.40.3

Amylose-lipid
complex
Potato

Waxy maize

HMT
Control
HMT-HP1.5
HMT-HP3.0
HMT-HP4.5

71.10.1
71.80.4
70.20.3
66.50.2
64.60.2

76.80.5
77.90.2
75.70.1
74.90.2
73.10.1

86.00.1
89.20.3
85.10.4
84.10.1
84.80.1

5.80.2
5.40.1
4.10.5
3.80.1
3.30.4

Native

92.90.3

100.10.4

107.40.2

1.60.1

HMT

93.10.6

100.30.2

107.20.5

1.70.3

Native
Control
HP1.5
HP3.0
HP4.5

61.30.1
62.30.4
60.10.1
59.60.3
55.90.1

67.00.3
68.20.1
65.20.3
64.50.2
61.10.1

82.60.5
81.30.1
78.10.1
76.00.3
73.00.2

15.60.3
15.40.2
14.30.2
13.80.1
13.20.4

HMT
Control
HMT-HP1.5
HMT-HP3.0
HMT-HP4.5

73.50.2
74.20.4
72.20.2
70.70.1
68.50.1

81.60.3
81.90.5
80.60.3
79.80.2
77.90.1

93.10.4
93.40.3
91.90.1
89.40.3
88.10.4

6.80.3
6.50.1
5.60.2
4.60.1
4.50.1

Native
Control
HP1.5
HP3.0
HP4.5

68.40.3
68.60.1
67.40.1
66.10.3
64.80.2

74.10.4
74.90.2
72.20.1
71.40.1
70.30.2

85.30.2
86.90.5
84.80.1
82.30.3
82.10.2

14.50.3
14.20.3
13.90.4
13.50.1
13.10.2

HMT
77.60.3
83.20.3
95.80.3
11.80.3
Control
77.90.4
83.80.2
95.90.1
11.70.1
HMT-HP1.5
77.20.1
83.10.2
93.80.6
9.70.2
HMT-HP3.0
75.80.2
81.20.1
91.20.3
9.40.2
HMT-HP4.5
74.10.1
80.00.5
90.30.4
8.80.1
a
To=onset, Tp=peak, Tc=conclusion.
b
HP=hydroxypropylated, HMT=heat-moisture treated, HMT-HP=heat-moisture
treated hydroxypropylated, 1.5, 3.0, 4.5=volume (mL) of propylene oxide.
c
Values are means of triplicate determinations standard deviation.

240

Table 3. Pasting propertiesa of hydroxypropylated native and heat-moisture


treated hydroxypropylated wheat, potato, and waxy maize starches
Starch
Treatment b
PV
HPV
BD
CPV
SB
c
Wheat
Native
1500.8 1120.6 380.7
2120.9 1000.5
Control
1641.3 930.9 710.4
2410.7 1450.8
HP1.5
2232.1 741.8 1480.9 1960.6 1221.3
HP3.0
2400.9 691.1 1710.8 1820.9 1121.1
HP4.5
2580.8 680.8 1900.6 1781.2 1090.6

Potato

Waxy
maize

HMT
Control
HMT-HP1.5
HMT-HP3.0
HMT-HP4.5

1001.6
741.5
831.1
880.3
1080.6

850.7
731.5
831.1
880.3
1021.8

150.4
0.0
0.0
0.0
6.00.4

1301.5
1730.9
2400.8
2880.5
3331.3

460.5
990.7
1570.9
2001.2
2320.7

Native
Control
HP1.5
HP3.0
HP4.5

6650.7
7051.6
7261.8
7451.2
7752.1

1740.7
1971.2
1741.5
1750.9
1651.2

4911.5
5121.6
5521.2
5702.1
6101.6

2561.6
2890.9
2281.1
2340.8
2341.3

810.8
920.6
541.3
580.6
681.4

HMT
Control
HMT-HP1.5
HMT-HP3.0
HMT-HP4.5

1621.3
1801.8
1740.9
2081.1
2170.7

1621.3
1801.8
1740.9
2081.1
2170.7

0.0
0.0
0.0
0.0
0.0

2530.9
2901.2
2981.4
3900.9
4451.3

911.2
1090.9
1231.3
1810.6
2300.5

Native
Control
HP1.5
HP3.0
HP4.5

2782.2
2711.5
2801.6
2840.9
3041.9

1171.5
1041.4
980.5
1040.7
1110.4

1601.2
1680.9
1820.8
1801.2
1931.1

1470.5
1380.8
1380.7
1410.5
1481.4

430.6
351.1
430.7
370.7
371.1

HMT
1440.8 691.2 751.5
820.7
130.5
Control
1460.7 621.3 830.5
770.6
140.8
HMT-HP1.5
1621.6 771.1 850.6
940.4
170.5
HMT-HP3.0
1692.9 782.1 900.8
960.7
180.8
HMT-HP4.5
1671.7 800.8 871.5
981.1
171.2
a
PV=peak viscosity, HPV=hot paste viscosity, BD=breakdown, CPV=cold paste
viscosity, SB=setback.
b
HMT=heat-moisture treatment, HP=hydroxypropylation, HMT-HP=heatmoisture treated hydroxypropylated, 1.5,3.0,4.5=volume (mL) of propylene
oxide.
c
Values are means of triplicate determinations standard deviation.

241

-3.2

-3.2

Amylose-lipid
complex

(A)
Amylose-lipid
complex
N

Heat flow (W/g)

-3.4

HMT

-3.4

-3.6

(B)

1.5

-3.6

1.5
3.0

3.0
-3.8

-3.8

4.5

-4.0

40

60

80

100

4.5

120

-4.0

50

60

70

80

90

100

110

120

Temperature (C)
Fig. 1. DSC curves obtained from gelatinization of hydroxypropylated native wheat starch (A); heat-moisture treated hydroxypropylated
wheat starch (B). N=native, HMT=heat-moisture treated, C=control, 1.5, 3.0, 4.5= volume (mL) of propylene oxide used for
hydroxypropylation.

242

100

100

240

B
320
80

80

Viscosity (RVU)

240
60

120

60

160
40

60
20

40

80

20

hmt

0
0

10

Temperature (C)

180

15

20

25

0
0

10

15

20

25

Time (Mins)
Fig. 2a. RVA curves of hydroxypropylated wheat native (A) and heat-moisture treated hydroxypropylated wheat (B) starches;
n=native; hmt=heat-moisture treated, c=control, other curves represented in both A and B from right to left are native or HMT
hydroxypropylated with 1.5, 3.0, and 4.5 mL propylene oxide.

243

100

800

100

400

80

80

300

60

60

400
200
40

40

Temperature (C)

Viscosity (RVU)

600

hmt

200
100
20

20

0
0

10

15

20

25

0
0

10

15

20

25

Time (Mins)
Fig. 2b. RVA curves of hydroxypropylated potato native (A), and heat-moisture treated hydroxypropylated potato (B) starches;
n=native; hmt=heat-moisture treated, c=control, other curves represented in both A and B from right to left are native or HMT
hydroxypropylated with 1.5, 3.0, and 4.5 mL of propylene oxide.

244

100

320

200

120

B
80

100
150

60

80
100

160

40

60

50

80

20

Temperature (C)

Viscosity (RVU)

240

40

0
0

10

15

20

25

10

15

20

25

Time (Mins)
Fig. 2c. RVA curves of hydroxypropylated waxy maize native (A) and heat-moisture treated hydroxypropylated waxy maize (B)
starches; C=control, other curves represented in both A and B from right to left are native or HMT, hydroxypropylated with 1.5, 3.0,
and 4.5 mL of propylene oxide.

245

-1.6

-1.6

HMT

Heat flow (W/g)

-1.8

-1.8

C
1.5

1.5
3.0

-2.0

-2.0

3.0

4.5

-2.2

40

50

60

70

80

4.5

90

-2.2

40

50

60

70

80

90

Temperature (C)
Fig. 3. DSC curves obtained from retrogradation of hydroxypropylated wheat native (A) and heat-moisture treated
hydroxypropylated wheat (B) starches. N=native, HMT=heat-moisture treated, C=control, 1.5, 3.0, 4.5 =volume (mL) of propylene
oxide used for hydroxypropylation.

246

Chapter 9
Gelatinizing, Pasting and Gelling Properties of Potato and Amaranth
Starch Mixtures
Abstract
Physicochemical properties of mixtures of native potato and native amaranth, heatmoisture treated potato and heat-moisture treated amaranth, cross-linked potato and
cross-linked amaranth, native potato and heat-moisture treated amaranth, and heatmoisture treated potato and native amaranth were tested at different ratios. Two peaks
were noticed in the pasting curves when large differences of swelling factor and
amylose leaching existed between individual components in the mixture. It seems that
amylose leached out from greater amylose leaching component in the mixture may
affect the swelling and major granular break down of the other component. The
mixtures showed higher hot paste stability, at least more than the less stable
component in the mixture. Some mixtures such as HMT potato and native amaranth
showed very specific non-additive pasting behavior. Mixing 10% of native amaranth
to HMT potato starch caused large reduction of peak viscosity and cold paste
viscosity resulting in a very soft gel. Each component of a mixture gelatinized
independently showing two peaks corresponding to the individual components,
indicating gelatinization in the mixture behaved as the sum of the individual
components. When transition temperatures of both components were similar in the
DSC, a single endotherm resulted. Dramatic changes of pasting and subsequent gel
properties resulted when thermal transition of two components occurs in the same
temperature range. Retrogradation enthalpies as measured by DSC were between the
two individual components in all tested mixtures.

247

9.1. Introduction
Functional properties of starch play a key role in its diverse applications in food and
non-food industries. Native starches generally do not posses desirable functional
properties for a wide range of utilization, so are frequently modified to improve these
properties. Chemical modification is widely used to attain this objective, however
with growing market demand for natural food there is greater necessity to search for
alternatives to chemical modification. One possibility may be the use of blends of
different starches, although this is not commonly practiced. Not much work has done
in this area despite its high potential. Obanni and BeMiller (1997) found that pasting
properties of some starch blends behave similarly to cross-linked starches and the
tendency to retrogradation decreased after blending. Karam et al. (2005) found that
blending native starches from maize, cassava, and yam improves some specific
sensory properties. In a study of starch gelatinization in a rice and wheat starch blend
Liu and Lelievre (1992) found at low starch concentration (<30%) DSC thermograms
are the sum of each individual components in the mixture, but non-additive behavior
was found at higher starch concentration due to the competition for water. The peak
for the higher gelatinization starch (rice) increases as the low gelatinization starch
(wheat) can access more water. Ortega-Ojeda and Eliasson (2001) reported that at
low starch concentration (20%) each individual component in the mixture
independently gelatinized whereas at higher starch concentration (50%) components
they did not.

From the survey of literature it seems that most studies on starch blending have been
limited to the use of native starches, thus it is worth studying the physical properties
of modified-unmodified starch blends, especially using physically modified starch.
The size of the starch granule in individual starch components in the mixture could

248

also affect the properties such as starch pasting, thus in this study we chose blends of
native, heat-moisture treated, and cross-linked potato and amaranth starches.
Amaranth is not a cereal, but its starch properties are similar to those of cereals. Its
extremely small starch granules may contribute specific functional properties to the
starch blends. In addition to this, differences in the physical sphere between small
granules (amaranth) and large granules (potato) could cause specific interaction
between them in thermal transition of their mixtures.

9.2. Materials and Methods

9.2.1. Materials
Potato starch and phosphoryl chloride were from Sigma Chemical Co., (St. Louis,
MO, USA). Amaranth starch was extracted from variety K 350 (Wu et al.1995).

9.2.2. Methods

Isolation of amaranth Starch


Amaranth starch was isolated by the method of Wu et al. (1995). Grain was soaked in
0.25% NaOH solution at 4 C for 24 hr then ground for 6 min in a Waring blender
with 6 volumes of 0.25 NaOH. The slurry was filtered through filtering cloth using a
small amount of water and the filtrate was centrifuged at 3000 g for 10 min. The
sediment was washed with distilled water several times centrifuging at 3000 x g for
15 min each time. The starch cake was then dried at 35 C.

Swelling factor

249

Swelling factor, the ratio of the volume of swollen starch granules to the volume of
dry starch was determined by the method of Tester and Morrison (1990a), when
starch (50 mg, db) was heated from 60 90 C in 5 mL water.

Amylose leaching
Distilled water or solution (10 mL) was added to starch (20 mg, db) in a screw cap
tube. Tubes were then heated at 60 - 90 C for 30 min with frequent agitation. After
cooling to ambient temperature, samples were centrifuged at 2000 g for 10 min.
Amylose content of supernatant (0.1 mL) was estimated as described by Chrastil
(1987).

Differential scanning calorimetry


Gelatinization and dissociation parameters of individual starch and the starch
mixtures were measured using a TA 2920 Modulated DSC Thermal Analyzer
differential scanning calorimeter equipped with a thermal analysis data station (TA
Instruments, Newcastle, DE). Starch (3 mg) was directly measured on to the
aluminum DSC pan (for the mixtures, relevant proportions of individual starch were
directly measured on to the pan separately) and distilled water (9 L) was added with
a microsyringe and mixed for homogenization. Pans were sealed, and allowed to
stand for 1 hr at room temperature for even distribution of water. The scanning
temperature and the heating rates were 30-120 C and 10 C/min respectively. An
empty pan was used as reference for all measurements.

Pasting properties
Pasting properties of starches were determined using a Rapid Visco-Analyzer (RVA)
model 3D (Newport Scientific, Warriewood, Australia). Distilled water (25.5 g) was

250

added to starch (2.5 g, db) in the RVA canister to obtain a total constant sample
weight of 28 g (8.9 % of starch concentration). Since the cross-linked potato starch
showed very high peak viscosity and cold paste viscosity development the amount of
dry starch content for cross-linked starches was decreased to 2.3 g (8.2 % of starch
concentration). Required proportions of individual components were directly
measured to RVA canister in preparing starch mixtures. The slurry was then manually
homogenized using the plastic paddle to avoid lump formation before the RVA run.
A programmed heating and cooling cycle was set for 22 minutes, where it was first
held at 50 C for 1.0 min, heated to 95 C in 7.5 min, further held at 95 C for 5 min,
cooled to 50 C within 7.5 min and held at 50 C for 1 min. The pasting stability ratio
was calculated by dividing hot paste viscosity value by the value of peak viscosity
(HPV/PV 100). Pasting stability ratio measures the degree of resistant to granular
breakdown.

Gel hardness
Gel hardness was determined on the starch gel made in the RVA testing using a TAXT2 Texture Analyzer (Stable Micro Systems, Godalming, Surrey, England). After
RVA testing, the paddle was removed and the starch paste in the canister was covered
by Parafilm and stored at 4 C for 24 hr. The gel was compressed at a speed of 0.5
mm/sec to a distance of 10 mm with a 6 mm cylindrical probe. The maximum force
peak (g) in the TPA profile represents the gel hardness.

Retrogradation
Starch gel was prepared in the DSC pan using 1:3 starch to water ratio and scanning
the sample from 30 C to 120 C. Sample was then stored at 4 C for 24 hr to initiate

251

nucleation, and then kept at 40 C for 10 days before rescanning. Temperature range
and heating rate were 30-120 C and 10 C/min respectively.

Heat-moisture treatment
Moisture content of starch samples (20 g) was brought up to 30 % in tightly capped
bottles. Sample was then kept at room temperature for 24 hr for equilibration of
moisture content before heating at 100 C for 8 hr. After cooling sample was air dried
to a moisture content of ~ 10 %.

Cross-linking
Cross-linking of starch with POCl3 was done as described by Woo and Seib (1997)
with some slight modifications. To make starch slurry, starch (50 g, dry basis) was
mixed with water (70 mL) containing sodium sulfate 1 g and stirred mechanically at
25 C for 1 h. The slurry was adjusted to pH 11.0 by slowly adding 1 M NaOH while
maintaining temperature at 25 C. Phosphoryl chloride (0.01% based on dry weight
starch) was injected with a microsyringe into starch slurry. After 1 hr, the slurry was
adjusted to pH 5.5 with 1 M HCl and starch was recovered by centrifuging (3000 g
for 10 min). Sedimented starch cake was washed three times with distilled water and
dried at 35 C.

9.3. Results and Discussion

Gelatinization
DSC results of individual components and all tested mixtures are presented in Table 2
and some representative DSC curves are shown in Figure 1. Gelatinization
temperature and the enthalpy of native amaranth starch were higher than native potato

252

starch. Consistent with previous research (Hoover and Vasnthan, 1994; Hoover et al.,
1994; Gunaratne and Hoover, 2002; Eerlingen et al., 1996) heat-moisture treatment
decreased gelatinization enthalpy while increasing the gelatinization temperature. The
effect was greater in potato starch. Cross-linking had little effect on gelatinization
parameters of both starches.

Except for the mixture of HMT potato and native amaranth starch, all other tested
mixtures showed two peaks during gelatinization. To of mixtures was similar to that
of the onset of low Tp component (potato) and the Tc was similar to that of the high
Tp component (amaranth) in the mixture (data not shown). Corresponding peak
temperatures observed for two peaks in the mixture was similar to the Tp of individual
components. Since the DSC characteristics of each component appeared in the
gelatinization of their mixtures, it appears that each components gelatinized
independently. This is an agreement with Ortega-Ojeda and Eliasson (2001) and Liu
and Lelievre (1992). The later authors found that at low starch concentration (<30%)
DSC thermograms are the sum of each individual component in the starch mixture.
However, Obanni and BeMiller (1997) noticed that at 33% starch solid none of the
blends of maize:potato, tapioca:wheat, and rice:potato showed two distinct peaks, and
the resulting Tp of the mixtures was different from either component. In our DSC
study at 25% starch concentration, except for HMT potato and native amaranth
mixture, all tested mixtures showed two peaks during gelatinization. The appearance
of a single peak in HMT potato and native amaranth mixture could be due to
overlapping of the two endotherms in the mixture because of reduction of
gelatinization temperature difference between potato and amaranth starches by heatmoisture treatment of potato starch so some thermal transition could take place over
the same temperature range.

253

Since cross-linking did not much affect gelatinization temperatures of both starches
the two peaks in the cross-linked mixture at a given ratio were very similar to the
peaks appearing for the unmixed starches (Fig. 1C). Among all tested mixtures,
native potato and HMT amaranth had the biggest difference in gelatinization
temperature and thus showed two peaks very clearly in gelatinization (Fig. 1D). Thus
it seems if the transition temperature of two components was close, the endotherms of
individual components in the mixture can overlap, indicating that starch crystals
which have similar thermal stabilities can gelatinize at same temperature range
regardless of botanical origin. What is important here is that gelatinization of each
component of starch in the mixture at the same temperature may cause some
interaction between two components in the mixture. That may be the reason for
significant changes of pasting and textural properties resulting from blends of HMT
potato and HMT amaranth, and HMT potato and native amaranth starch.

Pasting properties
Results for pasting properties of individual starches (8 9 %) and their blends are
presented in (Table 3 a, b, c, and d). As commonly observed potato starch showed
high peak viscosity development followed by rapid granular breakdown leading to a
progressive decrease of viscosity. Pasting properties of this specific amaranth starch
were somewhat similar to the behavior of normal cereal starches with low peak
viscosity and low granular breakdown. Wu and Corke (1999) found wide variation of
pasting properties among amaranth species and among genotypes within species
mainly due to differences in amylose content. Generally, amaranth starches are
characterized by low amylose content or are waxy. With blending of native potato
and native amaranth starches, peak viscosity increased as the proportion of potato
starch content increased, whereas a more stable paste was produced with increased

254

amaranth starch content. Thus mixing potato starch with amaranth starch produced a
more shear-resistant product than potato starch alone.

In the pasting profiles of native potato and of native amaranth, and native potato and
HMT amaranth starch mixtures, development of two peaks was observed. In both
cases the first peak corresponded to the peak viscosity of native potato starch. The
second peak position of the native potato and native amaranth mixture was slightly
before the peak viscosity of native amaranth starch. For native potato and HMT
amaranth mixture the position of the second corresponded peak was slightly before
the peak of HMT amaranth starch (Fig. 2). Obanni and BeMiller (1997) also noticed
two peaks when normal amylose maize and waxy maize starch were blended at 1:1
ratio. They speculated that the two peak characteristics could attribute to differences
in the granular breakdown between two components, i.e., at the time waxy maize
granules tend to break down, the normal amylose maize starch granules were
producing the viscosity, or the amylose leaching from normal amylose maize starch
granules may prevent the breakdown of waxy maize starch granules. We also
observed that the appearance of two peaks was more pronounced at 1:1 blend.
However, when cross-linking potato and cross-linking amaranth mixture was pasted
the two peaks characteristic disappeared (Fig. 2C).

Starch granular swelling and leaching of soluble carbohydrate (mainly amylose) are
the main factors that determine the viscosity development during the pasting process.
What we observed after HMT and cross-linking is a reduction of swelling factor and
amylose leaching and reduction in the differences in these two factors between
components of the blend (Table 1). Consequently it could decrease the differences in
swelling behavior and amylose leaching in the two components in the above three

255

mixtures allowing starches in the mixtures to swell more equally producing one peak
as usual.

we also found that the greater the differences in extent of swelling, amylose leaching,
and granular breakdown of the two components in the mixture the more the two
peaks in the pasting profile increased, which

is why native potato and HMT

amaranth mixture produced more distinctly two major peaks. In all cases when two
peaks showed, a greater difference of amylose leaching was noticed between
components. Thus it can be speculated that leached amylose from higher amylose
leaching component (potato) would prevent the breakdown of amaranth starch
granule to some extent. Obanni and BeMiller (1997) also claimed that greater
amylose leaching from one component (normal amylose maize) may prevent the
major granular breakdown in the other component (waxy maize) resulting two peaks
in the pasting process.

Heat-moisture treatment decreased the PV of potato starch but increased it in


amaranth starch. Additional starch chain interaction within the amorphous region,
destabilization and reordering of amylopectin double helices are adequately studied in
heat-moisture treated starches (Hoover and Vasanthan, 1994; Hoover et al., 1994). As
potato starch is more responsive to HMT more starch chain interaction can take place
within the amorphous and crystalline region leading to a larger reduction of swelling
and amylose leaching. As a consequence of this decreased peak viscosity is
anticipated. But in case of amaranth starch more granular rigidity is attained after
HMT due to starch chain interactions within the amorphous region which may
dominate the negative effect of decreased swelling. Hoover and Vasanthan (1994)
also reported similar finding for heat-moisture treated wheat starch. They suggested

256

that increase in the Brabender 95 C viscosity on heat-moisture treatment could be


attributed to an increase in granular rigidity resulting from an increase in crystalline
order and starch chain interactions within the amorphous region. Rigid granules
create more friction between granules during stirring because they deform less than
more swelling granules resulting higher PV. A similar explanation can be given for
the elevation of PV after cross-linking. Table 1 shows cross-linking reduces swelling
and thus probably increase the stiffness of granules. Less effectiveness of crosslinking in amaranth starch could be attributed to its low amylose content. Higher
pasting stability after HMT and cross-linking could be related to lower granular
breakdown because both HMT and cross-linking strengthen the starch granules
enabling them to maintain structure at higher temperature.

All blends of HMT potato and HMT amaranth, and HMT potato and native amaranth
starches showed lower PV and SB than the heat-moisture treated individual
components. Interestingly when a small quantity (10%) of HMT amaranth was mixed
with HMT potato a larger decrease in PV resulted without affecting the shape of the
pasting profile and the pasting stability ratio (Table 3b). A similar trend occurred
when small quantity of HMT potato was mixed with native amaranth starch (Table
3d). This kind of pasting behavior may be important in some applications and in an
area worthy of further research. In these two mixtures the dramatic reduction in PV
indicates decreased swelling characteristics in the mixed system, compared to the
individual components alone. When cross-linked potato and amaranth starches was
mixed PV increased as the proportion of potato starch increased. Significantly, no
breakdown was noticed for all blends indicating increased pasting stability of all
cross-linked mixtures compared to individual cross-linked components (Table 2c).

257

Cold paste viscosity of two blends (90:10 and 80:20) native potato and native
amaranth was more than for individual components. Since the prime factor for cold
paste viscosity is the aggregation of amylose in the starch paste, inclusion of a low
proportion of amaranth starch promotes the amylose association in the continuous gel
matrix.

Gel hardness
Amaranth starch gel was softer than potato starch gel. After HMT and cross-linking
the gels produced by the two starches were firmer than their native counterparts
(Table 3a, b, c, and d). Starch gels are metastable and non-equilibrium systems and
therefore undergo structural changes during storage (Ferrero et al., 1994). Miles et al.
(1985) and Ring et al. (1987) attributed the initial gel firmness during retrogradation
to the formation of an amylose matrix and the subsequent slow increase in gel
firmness to reversible crystallization of amylopectin. Morris (1990) explained that
starch gel properties relate to the characteristics of gel matrix, the amylose, the
deformable fillers (swollen granules) that are embedded in the continuous amylose
matrix, volume fraction of the filler, and the filler-matrix interaction. Doublier et al.
(1987) suggested the main structural parameters of a starch gel involved are the
deformability of swollen starch particles and the amylose concentration in the
continuous network. The weaker gel of amaranth starch therefore could be attributed
to its low concentration of amylose in the continuous network and also to deformable
and relatively soft starch granules. Eerlingen et al. (1997) found that the increased or
decreased gel storage moduli could result based on the heat-moisture treatment
conditions and the concentration of the starch gel, as these affect the extent of
swelling, solubility and close packing concentration. Hoover et al. (1994) reported

258

that restriction of swelling by HMT favors the amylose aggregation in the formation
of gel matrix leading to firmer gel after HMT.

Similarly to HMT a harder gel was formed after cross-linking but to a greater extent.
Cross-linking reinforces the starch granule improving its rigidity. This could affect
the deformability of the starch granules in the gel matrix, and perhaps increasing
bonding of amylose after cross-linking may create more junction zones promoting
amylose aggregation. The ability of cross-linking to increase gel hardness in both
starches is in agreement with the increase in cold paste viscosity. Gel hardness for the
all mixtures of native potato and native amaranth were in between individual starch
components. A similar trend was seen when cross-linked potato starch was mixed
with cross-linked amaranth starch, but a dramatic reduction of gel hardness resulted
from the HMT potato and HMT amaranth, and HMT potato and native amaranth
mixtures (Table 3c and 3d). This is in agreement with what is seen for the large
reduction of peak viscosity and cold paste viscosity when these two mixtures were
pasted.

In the starch gel, it is worth testing whether the two types of starch co-exist in a single
phase in the absence of significant steric exclusion phenomena, or do they form a
phase separated network with distinct polymeric domains. If the later is the case, it is
worth testing whether it is a bicontinous blend, or a system with a continous phase
supporting the discotinous filler inclusions of the second polymer. Work on this issue
could be improved with an element of modeling/quantative analysis of the hardness
values and optical microscopy images offering tangible evidence of the topology of
the binary starch gel.

259

Retrogradation
Differential scanning calorimetry was used to measured the melting parameters of reorganized or retrograded amylopectin for individual components and their mixtures
stored after nucleation at 4 C for 24 h and growth at 40 C for 7 days (Table 4 with
some representative DSC curves in Fig. 3). DSC curves of all the individual starches
and their blends showed a single monomodal endotherm. With the exception of HMT
potato H, all the other melting properties of retrograded native, HMT, and crosslinked potato were generally similar. In case of amaranth starches both HMT and
cross-linking had no influence on melting properties of retrograded amylopectin. The
gels of all potato starches showed more retrogradation than the amaranth starches.
This may relate to the structural properties of amylopectin, specifically the side chain
length and the ratio of the long chains. Amylopectin with a long side chain length in a
high ratio has been shown to retrograde more than that with a short chain length
(Kohyama et al., 2004). Perhaps the low amylose content of amaranth starch may
reduce amylopectin retrogradation. Bulkin et al. (1987) have shown that amylose has
a template effect which accelerates amylopectin retrogradation. According to Hoover
and Gunaratne (2002) heat-moisture treatment decreased the retrogradation of B-type
starch such as potato, but caused no significant changes to the retrogradation of Atype starch such as wheat. This is an agreement with the observed results in this
study.

Cross-linking had no influence on amylopectin retrogradation indicating it largely


occurred in the amylose region, or that covalant bonds formed by cross-linking
cannot interfere with the associations of amylopectin chains. The melting
temperatures (Tp) of retrograded amaranth starch samples were less than for potato
starch in all cases and in the mixtures Tp decreased as the proportion of low

260

retrograded amaranth starch increased. Generally Tp was in between the value of


individual components based on their proportion in the mixture. A similar trend was
observed for the melting enthalpies, thus it seems each component in a mixture
contributes to the retrogradation according to its proportion.

9.4. Conclusions
This study shows specific non-additive pasting behavior for some of the potato and
amaranth starch mixtures. Such pasting behavior may be useful for some application
in food products and may be an alternative to the current chemical modification.
Interaction between two components in a blend can occur when swelling ability and
amylose leaching differed greatly for starch of different granule size. In many cases
more stable paste can be produced when potato starch is mixed with amaranth starch,
depending on the ratio of the mixture. Except HMT potato and HMT amaranth, and
HMT potato and native amaranth pasting properties and gel hardness were
intermediate between the two individual components, and in retrogradation properties
as measured by DSC all tested mixtures were intermediate between the values of
individual components. Individual components in the mixture appeared to have
gelatinized independently. When the two components in the mixture gelatinized at a
similar temperature range more interaction between two components could take place
causing more changes in pasting properties. It is worth testing physical properties of
more mixtures including modified and unmodified starches, and to determine their
optimum ratio for the desired physical properties.

9.5. References
Bulkin, B.J., Kwak, Y., and Dea, I.C.M. 1987. Raman spectroscopic study of starch
retrogradation. Carbohydr. Res. 160: 95-112.

261

Chrastil, J. 1987. Improved colorimetric determination of amylose in starches or


flours. Carbohydr. Res. 159: 154-158.
Doublier, J.L., Llamas, G., and Le Meur, M. 1987. A rheological investigation of
cereal starch pastes and gels. Effects of pasting procedures. Carbohydr. Polym. 7:
251-275.
Eerlingen R.C., Jacobs, H., Block, K., and Delcour, J.A. 1997. Effects of
hydrothermal treatments on the rheological properties of potato starch. Carbohydr.
Res. 297: 247-356.
Eerlingen, R.C., Jacobs, H., Van Win, H., Delcour, J.A., 1996. Effect of hydrothermal
treatment on the gelatinization properties of potato starch as measured by differential
scanning calorimetry. J. Therm. Anal. 47: 1229-1246.
Ferrero, C., Martin, M.N., and Zantzky, N.E. 1994. Corn-starch-xanthan gum
interaction and its effect on the stability during storage of frozen gelatinized
suspension. Starch 46: 300-305.
Gunaratne A. and Hoover, R. (2002) Effect of heat-moisture treatment on the
structure and physicochemical properties of tuber and root starches. Carbohydr.
Polym. 49: 425-437.
Hoover, R., and Vasanthan, T. 1994. Effect of heat-moisture treatment on the
structure and physicochemical properties of cereal, legume, and tuber starches.
Carbohydr. Res. 252: 33-53.
Hoover, R. Vasnathan T., Senanayake, N.J., and Martin, A.M. 1994. The effect of
defatting and het-moisture treatment on the retrogradation of starch gels from wheat,
oat, potato, and lentil. Carbohydr. Res. 261: 13-24.
Karam, L., Grossmann, M.V.E., Silva, R.S.F., Ferrero, C., and Zaritzky, N.E. 2005.
Gel textural characteristics of corn, cassava and yam starches blends: A mixture
surface response methodology approach. Starch 57: 62-72.
Kohyama, K., Matsuki, J., Yasui, T., Sasaki, T. 2004. A differential thermal analysis
of the gelatinization and retrogradation of wheat starches with different amylopectin
chain lengths. Carbohydr. Polym. 58: 71-77.
Liu, H., and Lelievre, J. 1991. Differential scanning calorimetry study of melting
transitions in aqueous suspensions containing blends of wheat and rice starch.
Carbohydr. Polym. 17: 145-149.
Morris, M.J. 1990. Starch gelation and retrogradation. Trends in Food Sci. & Technol.
1: 2-6.
Obanni, M., and BeMiller, J.N. 1997. Properties of some starch blends. Cereal
Chem. 74: 431-436.
Ortega-Ojeda, F.E., and Eliasson, A.C. 2001. Gelatinization and retrogradation
behavior of some starch mixtures. Starch 53: 520-529.

262

Ring, S.G., Colonna, P., IAnson, K.J., Kalicheversk, M.T., Miles, M.J., Morris, V.J.,
and Orford, P.D. 1987. The gelation and crystallization of amylopectin. Carbohydr.
Res. 162: 277-293.
Tester, R.F., and Morrison, W.R. 1990a. Swelling and gelatinization of cereal
starches. 1. effects of amylopectin, amylose, and lipids. Cereal Chem. 67: 551-557.
Woo, K. and Seib, P.A. 1997. Cross-linking of wheat starch and hydroxypropylated
wheat starch in alkaline slurry with sodium trimetaphosphate. Carbohydr. Polym. 33:
263-271.
Wu, H. X., and Corke, H. 1999. Genetic diversity in physical properties of starch
from a world collection of Amaranthus. Cereal Chem. 76: 877: 883.
Wu, H. X., Yue, S., Sun, H., and Corke, H. 1995. Physical properties of starch from
two genotypes of Amaranthus cruentus of agricultural significance in China. Starch
47: 295-297.

263

Table 1. Swelling factora and amylose leachingb of native, HMT, and cross-linked
potato and amaranth starches at different temperatures
Swelling factor
Starch
Treatment
60 C
70 C
80 C
90 C
Potato

Native
HMT
Cross-linked

21.00.1c
3.00.2
6.30.3

38.40.2
10.50.1
17.20.1

56.00.3
14.20.4
25.60.1

48.10.2
18.30.2
32.10.2

Amaranth

Native
HMT
Cross-linked

6.70.4
2.60.2
4.00.1

18.20.2
5.40.1
12.20.1

27.60.1
10.40.1
16.80.2

30.00.1
20.00.2
24.60.1

Amylose leaching (%)


Potato

Amaranth

Native
HMT
Cross-linked

8.20.2
0.20.1
2.80.1

13.30.1
1.80.2
3.60.1

14.70.1
3.60.1
4.10.2

Native
0.80.1
1.90.3
3.50.1
HMT
0.10.3
0.30.1
2.40.2
Cross-linked 0.40.2
1.20.2
3.30.3
a
Starch concentration (1%).
b
Starch concentration (0.2%).
c
Values are means of triplicate determinations standard deviation.

23.40.2
6.30.1
6.10.3
3.80.2
3.20.1
3.50.4

264

Table 2. Gelatinization parametersa for mixtures of potato and amaranth starchess


Starch
Native potato:Native
HMTpotato:HMT
Cross-linked
ratio
amaranth
amaranth
potato:Cross-linked
(Potato:
amaranth
Amaranth)
Tp (C)
H
Tp (C)
H
Tp (C)
H (J/g)
100:0

64.80.1c

14.20.2

77.00.3

9.10.2

64.80.2

14.30.2

0:100

78.50.1

15.00.1

83.70.1

14.10.2

78.10.1

14.80.3

65.00.3
14.30.2
76.90.2
9.50.4
65.40.4
78.80.2
84.10.1
78.00.3
70:30
65. 40.2 13.90.1
77.20.2
10.50.1
65.10.3
79.00.1
84.30.1
78.10.2
50:50
65.70.2
14.10.2
77.40.2
10.20.1
65.10.2
78.90.6
84.10.2
77.80.1
30:70
65.30.4
14.30.1
77.90.3
11.50.4
65.30.1
78.60.1
82.90.3
78.00.3
65.30.2
14.90.3
77.60.2
13.40.1
65.40.2
10:90
78.80.1
84.10.1
77.90.3
Potato:
Native potato:HMT
HMT potato:Native
amaranth
Amaranth amaranth
90:10
65.30.2
14.10.2
77.40.1
8.50.2
83.90.1
70:30
64.80.4
14.30.2
77.80.3
10.5.2
84.00.1
50:50
65.00.2
14.20.1
77.90.2
13.80.1
84.10.2
30:70
65.20.2
14.60.2
78.00.1
11.60.2
84.20.3
10:90
65.10.2
14.20.2
78.30.2
14.10.1
83.90.1
a
Tp=gelatinization temperature; H=enthalpy.
b
Starch concentration (25%).
c
Values are means of triplicate determinations standard deviation.

14.10.4

90:10

14.00.3
14.30.1
14.00.4
15.10.3

265

Table 3a. Pasting propertiesa and gel hardness (g) of mixtures of native potato:native amaranth starchesb
Starch
Ratio
PV
HPV
BD
CPV
SB
GH

SR

Potato

100:0

6552.2c

1740.9

4811.7

2542.9

812.2

370.5

0.260.01

Amaranth

100:0

1650.8

1270.7

380.4

1831.4

570.4

90.4

0.760.03

Potato:Amaranth

90:10

5011.4

1760.7

3240.9

2561.5

802.7

340.7

0.350.04

70:30

3762.3

1640.9

2121.4

2410.8

781.8

270.3

0.440.08

50:50

2730.8

1443.6

1290.9

2140.7

700.2

200.3

0.530.00

30:70

2071.2

1250.8

810.6

1951.3

691.3

160.4

0.600.03

1171.2
1770.7 600.8
120.3
0.720.00
10:90
1621.5
450.9
PV=peak viscosity; HPV=hot paste viscosity; BD=breakdown; CPV=cold paste viscosity, SB=setback; GH=gel hardness; SR=stability
ratio.
b
Starch concentration (8.9%).
c
Values are means of triplicate determinations standard deviation
a

266

Table 3b. Pasting propertiesa and gel hardness (g) of mixtures of HMT potato:HMT amaranth starchesb
Starch
Ratio
PV
HPV
BD
CPV
SB
GH

SR

HMT potato

100:0

1441.2c

1441.1

0.0

2181.2

743.5

400.6

1.00.00

HMT amaranth

100:0

2061.3

1661.4

390.7

2271.9

602.8

120.6

0.80.00

HMT potato:HMT
amaranth

90:10

480.4

480.4

0.0

780.6

300.3

60.3

1.00.00

70:30

461.6

460.6

0.0

710.3

232.4

20.1

1.00.00

50:50

941.7

940.7

0.0

1231.2

291.5

30.2

1.00.00

30:70

1302.3

1211.0

8.60.3

1720.7

510.8

60.2

0.930.02

10:90 1811.9
1541.9
270.8
2100.8
551.6
50.1
0.850.00
PV=peak viscosity; HPV=hot paste viscosity; BD=breakdown; CPV=cold paste viscosity; SB=setback; GH=gel hardness; SR=stability
ratio.
b
Starch concentration (8.9).
c
Values are means of triplicate determinations standard deviations.
a

267

Table 3c. Pasting propertiesa and gel hardness (g) of mixtures of cross-linked potato and cross-linked amaranth starchesb
Starch
Ratio
PV
HPV
BD
CPV
SB
GH
SR
Potato native

100:0

5513.2c

1522.1

4001.4

2263.3

750.9

330.7

0.830.00

Amaranth native

100:0

1251.1

1142.2

121.3

1550.9

410.8

90.4

0.930.03

Potato cross-linked

100:0

6752.1

5632.4

1121.2

7173.5

1531.4

900.5

0.830.00

Amaranth cross-linked

100:0

1280.9

1191.3

90.2

1581.2

380.8

150.2

0.930.03

Potato cross-linked:
Amaranth cross-linked

90:10

5493.1

5493.1

0.00

7364.5

1881.6

710.6

1.00.00

70:30

3912.4

3913.4

0.00

6862.3

2945.2

420.1

1.00.00

50:50

2843.7

2843.7

0.00

4942.2

1092.3

300.3

1.00.00

30:70

1751.2

1751.2

0.00

2750.7

1000.3

160.5

1.00.00

10:90 1222.0
1222.0
0.00
1680.4
460.4
110.4
1.00.00
PV=peak viscosity; HPV=hot paste viscosity; BD=breakdown; CPV=cold paste viscosity; SB=setback; GH=gel hardness; SR=stability
ratio.
b
Starch concentration (8.2%).
c
Values are means of triplicate determinations standard deviation.
a

268

Table 3d. Pasting propertiesa


starchesb
Ratio
Starch
Native Potato
100:0
HMT amaranth
100:0
Native potato:HMT
90:10
amaranth
70:30
50:50
30:70
10:90
HMT potato
Native amaranth
HMT potato:Native
amaranth

100:0
100:0
90:10

and gel hardness (g) of mixtures of native potato:HMT potato and HMT potato and native amaranth
PV
6553.2c
2061.4
5303.1

HPV
1742.1
1661.9
1742.1

BD
4811.4
392.3
3552.5

CPV
2540.8
2272.8
2501.2

SB
810.6
601.9
752.1

GH
370.3
120.2
340.7

SR
0.260.02
0.800.01
0.320.02

3830.9
3261.5
2683.1
2273.4

1823.2
1811.8
1821.2
1662.2

2011.4
1442.3
851.5
613.5

2550.7
2503.4
2492.2
2372.3

730.7
680.6
662.1
702.1

270.4
230.5
190.7
160.8

0.470.01
0.550.03
0.670.02
0.730.03

1443.7
1652.1
332.1

1443.7
1270.9
332.1

0.0
381.2
0.0

2181.9
1831.8
631.3

741.2
571.4
301.9

401.3
90.8
20.2

1.00.0
0.760.01
1.00.00

70:30 420.9
420.9
0.0
670.8
341.3
30.6
1.00.00
851.4
0.0
1222.1
373.3
30.2
50:50 851.3
1.00.00
30:70 1192.8
1192.8
0.0
1612.4
421.9
30.8
1.00.00
10:90 1543.1
1381.5
160.9
30.4
0.890.01
2012.3
633.3
a
PV=peak viscosity; HPV=hot paste viscosity; BD=breakdown; CPV=cold paste viscosity; SB=setback; GH=gel hardness; SR=stability
ratio.
b
Starch concentration (8.9%).
c
Values are means of triplicate determinations standard deviation.

269

Table 4. Melting parametersa obtained from DSC for retrograded starch of potato
and amaranth starch mixturesb
Starch ratio
Native potato:Native
HMT potato:HMT
Cross-linked
(Potato:Amaranth) amaranth
amaranth
potato:Crosslinked
amaranth
Tp (C)

H (J/g)

Tp (C)

H (J/g)

Tp (C)

100:0

70.10.7 c

4.50.4

69.70.5

3.40.2

70.20.4

4.30.8

0:100

66.10.3

0.70.1

65.90.3

0.70.1

66.10.2

0.80.1

90:10

70.10.4

4.10.1

70.00.6

2.80.5

70.70.3

3.40.4

70:30

70.30.9

3.50.4

70.00.7

2.80.5

70.50.4

2.90.5

50:50

70.20.5

2.80.1

70.10.4

2.20.3

70.70.3

2.70.2

30:70

69.70.7

1.50.3

70.40.4

2.00.1

71.40.8

2.20.4

67.60.6
0.80.1
HMT potato:Native
amaranth

66.90.8

1.10.1

10:90
Potato:Amaranth

67.10.3
1.30.1
Native potato:HMT
amaranth

90:10

70.10.6

3.90.3

70.10.9

2.80.3

70:30

70.20.4

3.70.2

70.10.5

2.10.1

50:50

70.10.7

2.40.1

69.80.2

2.00.1

30:70

70.00.2

1.90.4

69.10.5

1.60.2

10:90
67.50.5
1.10.2
67.20.3
0.90.1
Tp=peak; H=enthalpy.
b
Starch concentration (50%).
c
Values are means of triplicate determinations standard deviation.
a

270

H (J/g)

-2.0

-2.2

-2.1

pn

phmt

-2.3

-2.4

an

ahmt

-2.5

-2.6

-2.7

50:50

-2.8

50:50
-2.9

-3.0

-3.2

-3.1

40

50

60

70

80

90

100

50

60

70

80

90

100

110

Heat flow (W/g)

-2.0

-2.1

-2.2

pc

pn

-2.3
-2.4

ac

-2.6

-2.5

ahmt
-2.7

-2.8

50:50

-3.2

40

50

60

-2.1

70

80

50:50

-2.9

-3.0

90

100

-3.1

40

50

60

70

80

90

an

-2.3

-2.5

phmt

-2.7

50:50

-2.9

-3.1

50

60

70

80

90

100

110

Temperature (C)

Fig. 1. DCS curves of 50:50 mixture of native potato:native amaranth (A), HMT
potato:HMT amaranth (B), cross-linked potato:cross-linked amaranth (C), native
potato:HMT amaranth (D), HMT potato:native amaranth (E). pn=potato native;
an=amaranth native; phmt=potato heat-moisture treated; ahmt=amaranth heatmoisture treated; pc=potato cross-linked; ac=amaranth cross-linked.

271

100

110

60

10
0

300

240

180

100

pn

45
0

50:50 (pn:an)

80

60

ahmt
phmt

30
4

15

120

40

60

20

Viscosity RVU

an
0

0
0

15

0
0

10

15

20

25

100

100

600

80

pn

80

80

450

60

50:50 (pn:ahmt)

60

Temperature (C )

50:50 (phmt:ahmt)

60

pcl
ahmt

300

40

phmt

an
40

40

50:50 (pcl:acl)

150

20

20

20

50:50 (phmt:an)
acl
0

0
0

10

15

20

25

0
0

10

15

20

25

Time Mins

Fig. 2. Pasting curves at 8.9% or 8.2% starch solids for 50:50 mixture of pn:an (A), phmt:ahmt (B), pcl:acl (C), pn:ahmt, and
phmt:an (D). pn=potato native; an=amaranth native; phmt=heat-moisture treated potato; ahmt=heat-moisture treated amaranth;
pcl=cross-linked potato; acl=cross-linked amaranth.

272

-1.6

-2.0

-2.1

pn

ph

-1.8

Heat fow (W/g)

-2.2

-2.3

an

-2.0

ah
-2.4

50:50

-2.2

-2.5

50:50
-2.6

-2.4

40

50

60

70

80

90

100

110

-2.7

40

50

60

70

80

90

100

Temperature (C)

Fig. 3. DSC curves obtained for retrogradation of 50:50 mixture of pn:an (A), ph:ah (B). pn=native potato, an=native amaranth; ph=heatmoisture treated potato; ah=heat-moisture treated amaranth.

273

Chapter 10
Functional Properties of Hydroxypropylated, Cross-Linked and
Hydroxypropylated Cross-Linked Tuber and Root Starches
Abstract
Functional properties of some under exploited tuber and root starches (true yam,
gourd yam, taro, lotus, and sweet potato) were investigated before and after
hydroxypropylation, cross-linking, and hydroxypropylated cross-linking using potato
starch as the reference. Low swelling ability, poor viscosity development but greater
shear stability, gel hardness, and resistant to enzyme hydrolysis was observed in true
yam and gourd yam. The extent of retrogradation was also greater in these two
starches. Most of the functional properties of lotus starch were similar to those of
potato starch. Hydroxypropylation increased the swelling factor, susceptibility to
enzyme hydrolysis but decreased acid tolerance, retrogradation, gelatinization
parameters, gel hardness, and shear stability. Decreased or increased amylose
leaching

and

peak

viscosity

depending

on

the

starch

resulted

from

hydroxypropylation. Cross-linking decreased swelling factor, and amylose leaching,


and increased shear stability and resistance to enzyme and acid hydrolysis. Crosslinking had very little influence on gelatinization and retrogradation properties but a
larger effect on pasting properties. Increased or decreased peak viscosity and gel
hardness were noted for different cross-linked starches. Introduction of cross-linking
to hydroxypropylated starches increased commonly desirable functional properties
providing a wider range of potential applications.

274

10.1. Introduction
Starch is the major reserve polysaccharide in higher plants and can be found in stems,
tubers, roots, fruits, and grains, etc. Tropical tuber and root starches are the main
staple food for many people in the worlds hot and humid region. These crops are
highly adapted to the tropical agro climatic environment and can grow in large
abundance with little or no artificial inputs. Although their agronomic and phenotypic
properties are well documented little effort has been made to exploit the functional
properties of these starches aiming to bring them to commercial level applications. A
detailed knowledge of starch characteristics of underexploited starches may help to
select suitable genotypes that behave like potato and cereal starches. Functionality is
the key to many applications and for those starches that do not have desirable
functional properties modification techniques can be used to improve the
functionalities.

Hydroxypropylation and cross-linking are two widely used methods for making
modified starch. Hydroxypropylation has been shown to increase freeze thaw stability,
decrease gelatinization and pasting temperatures, and increased paste clarity (Hoover
et al., 1988; Kim and Eliasson, 1993; Perera et al., 1997; Liu et al., 1999; Pal et al.,
2002) while cross-linking provides more stable paste at higher temperature and at low
pH (Woo and Seib, 1997; Liu et al., 1999; Wurzburg, and Szymanski, 1970). It can
therefore be expected that cross-linking of hydroxypropylated starch may provide a
wider range of functional properties permitting more applications. To date, most
study and production of chemically modified starches has been limited to widely
available starches such as wheat, maize and potato, thus studying chemical
modifications with starches from other botanical source may provide deeper insight
on the structure and physicochemical properties of starch. My objective was to

275

investigate the functional properties of some under-exploited tuber and root starches
and how the properties of these starches could be improved by means of
hydroxypropylation, cross-linking, and hydroxypropylated cross-linking.

10.2. Materials and Methods

10.2.1. Materials
Potato starch, fungal -amylase (EC 3.2.1.1), phosphoryl chloride and propylene
oxide were from Sigma Chemical Co., (St. Louis, MO, USA). True yam (Dioscorea
alata), gourd yam (Cucurbita foetidissima), taro (Alocassia esculenta), lotus
(Nelumbo nucifera), and sweet potato (Ipomea batatas) were purchased in a local
market.

10.2.2. Methods

Starch isolation
Starches were isolated as described by Hoover and Hadziyev (1981) with some slight
modifications. The tubers and roots were peeled, washed, diced, dipped in ice-cold
water containing 100 ppm NaHSO3 and homogenized at low speed in a Warning
blender. The slurry was filtered through filtering cloth using small amount of water
and the filtrate was centrifuged at 3000 g for 10 min. The supernatant and the
amber brown layer of protein on the surface of the sediment were removed and
further purification was achieved by repeated suspension in water and centrifugation.
The purified starch was dried at 35 C.

276

Swelling factor
Swelling factor, the ratio of the volume of swollen starch granules to the volume of
dry starch was determined by the method of Tester and Morrison (1990a), where
starch (50 mg, db) was heated at 85 C for 30 min in 5 mL water.

Amylose leaching
Distilled water or solution (10 mL) was added to starch (20 mg, db) in a screw cap
tube. Tubes were then heated at 85 C for 30 min. After cooling to ambient
temperature, samples were centrifuged at 2000 g for 10 min. Amylose content of
supernatant (0.1 mL) was estimated as described by Chrastil (1987).

Differential scanning calorimetry


Gelatinization and dissociation parameters were measured using a TA 2920
Modulated DSC Thermal Analyzer differential scanning calorimeter equipped with a
thermal analysis data station (TA Instruments, Newcastle, DE). Starch (3 mg) was
directly measured on to the aluminum DSC pan and distilled water (9 L) was added
with a microsyringe and mixed for homogenization. Pans were sealed, and allowed to
stand for 1 hr at room temperature for even distribution of water. The scanning
temperature and the heating rates were 30-120 C and 10 C/min respectively. An
empty pan was used as reference for all measurements.

Pasting properties
Pasting properties of starches were determined using a Rapid Visco-Analyzer (RVA)
model 3D (Newport Scientific, Warriewood, Australia). Distilled water or
hydrochloric acid solution (pH=3) (25.5 g) was added to starch (2.3 g, db) in the
RVA canister to obtain a total constant sample weight of 27.8 g (8.2% starch

277

concentration). The slurry was then manually homogenized using the plastic paddle to
avoid lump formation before the RVA run. A programmed heating and cooling cycle
was set for 22 minutes, where it was first held at 50 C for 1.0 min, heated to 95 C in
7.5 min, further held at 95 C for 5 min, cooled to 50 C within 7.5 min and held at
50 C for 1 min.

Gel hardness
Gel hardness was determined on the starch gel made in the RVA testing using a TAXT2 Texture Analyzer (Stable Micro Systems, Godalming, Surrey, England). After
RVA testing, the paddle was removed and the starch paste in the canister was covered
by Parafilm and stored at 4 C for 7 hrs. The gel was compressed at a speed of 0.5
mm/sec to a distance of 10 mm with a 6 mm cylindrical probe. The maximum force
peak in the TPA profile represents the gel hardness.

Enzymatic hydrolysis
Enzymatic hydrolysis was measured using a viscoamylographic (RVA) method as
described by Li et al. (2000). Development of peak viscosity was measured before
and after adding 100 units of fungal -amylase and then percent decrease of peak
viscosity was calculated to detect the extent of enzyme hydrolysis. In this analysis,
same heating cooling program (Std. 2), and same starch concentration, given above,
was used as in the measurement of pasting properties.

Retrogradation
After gelatinization test (DSC) samples were stored at 4 C for 24 hr to initiate
nucleation. Then samples were kept at 40 C for 10 days before rescanning by DSC.
Temperature range and heating rate were 30-120 C and 10 C/min respectively.

278

Hydroxypropylation
Hydroxypropylation of tuber and root starches was performed according to the
method of Choi and Kerr (2004) with some slight modifications. Starch sample (50 g
db) was suspended in distilled water (110 mL) containing 10 g Na2SO4 in a centrifuge
bottle. After adjusting pH to 11.3 with 1M NaOH, 5 mL of propylene oxide was
added and the bottle was immediately capped and shaken vigorously. Sample was
then placed at 35 C in shaking water bath with continuous shaking for 24 hr. The
reaction was terminated by adjusting pH to 5.3 with 1M HCl. Slurry was then
centrifuged at 3000 g for 10 min and recovered starch cake was washed with three
times with distilled water and dried at 35 C.

Cross-linking
Cross-linking of starch with POCl3 was done as described by Woo and Seib (1997)
with some slight modifications. To make starch slurry, starch (50 g, db) was mixed
with water (70 mL) containing sodium sulfate (1 g) and stirred mechanically at 25 C
for 1 hr. The slurry was adjusted to pH 11 by slowly adding 1 M sodium hydroxide
while maintaining temperature at 25 C. Phosphoryl chloride (0.01% based on dry
weight starch) was injected with a microsyringe into starch slurry. After 1 hr, the
slurry was adjusted to pH 5.5 with 1 M HCl and starch was recovered by
centrifugation (3000 g for 10 min). Sediment starch cake was washed three times
with distilled water and dried at 35 C.

Cross-linking of hydroxypropylated starch


Hydroxypropylation was accomplished according as above and samples were then
cooled to room temperature (25 C). After adjusting pH to 11.0 phosphoryl chloride
(0.01% based on dry weight starch) was injected with a microsyringe into starch

279

slurry. After 1 hr, the slurry was adjusted to pH 5.5 with 1 M HCl and starch was
recovered by centrifuging (3000 g for 10 min). Sediment starch cake was washed
three times with distilled water and dried at 35 C.

10.3. Results and Discussion

Swelling factor and amylose leaching


A wide variation in swelling ability was observed for native tuber and root starches at
85 C (Table 1). These differences could be attributed to variation in granule size, the
amount and structural characteristics of amylose and amylopectin, crystalinity, and
phosphate content in the starch granules.

Swelling ability reflects the magnitude of starch chain interaction within the
amorphous and crystalline regions (Hoover, 2001). Potato had the highest swelling
starch followed by lotus, with lowest value in true yam. It is believed the presence
high level of phosphate groups in potato starch impart its greater swelling ability
because of the repulsion between negatively charged phosphate groups in neighboring
amylopectin chains could weaken the hydrogen bonding in the crystalline region
leading to rapid hydration and swelling (Galliard and Bowler 1987).

Hydroxypropylation increased the swelling factor of all the starches compared to their
corresponding native starches, while cross-linking decreased the swelling factor.
Etherified hydroxypropyl groups in the neighboring starch chains prevent inter-chain
association facilitating water molecules to penetrate into the granules and thereby
increase the swelling power, but cross-linking that reinforced the starch granules will
limit the water absorption restricting the mobility of starch chains in the amorphous
280

region. As expected, in all starches cross-linking decreased the swelling factor of


hydroxypropylated starches. Except for potato and lotus starch increased amylose
leaching was observed after hydroxypropylation. Amylose is derivatized to a greater
extent than amylopectin in hydroxypropylation (Shi and BeMiller 2000 and Kavitha
and BeMiller, 1998) facilitating
preference

leaching

from the starch granules, but the

for leaching of derivatized amylose decreases as the degree of

modification of the whole starch granule increases (Shi and BeMiller, 2002). Thus the
decreased amylose leaching observed for potato and lotus starches may be attributed
to a greater degree of modification occurring in these two starches under the given
conditions in hydroxypropylation. As expected, cross-linking decreased the amylose
leaching to a greater extent and reduced the increase in amylose leaching due to
hydroxypropylation.

Gelatinization
DSC curves recorded for the gelatinization of native and modified tuber and root
starches are presented in Figure1 and gelatinization enthalpy is presented in Table 2.
The highest H was noticed for true yam followed by gourd yam and lowest in lotus
starch. Both potato and lotus starch gelatinized nearly at a same temperature whereas
other starches comparatively gelatinized at higher temperature.

After hydroxypropylation decreased gelatinization temperatures and the enthalpy in


all the starches are anticipated consequences because extensively hydrated granules
required less energy to reach the gelatinization. Seow and Thevamalar (1993)
reported that hydroxypropylation facilitates water penetration and absorption into the
starch granules and increases the initial rate of plasticization of amorphous region
promoting the gelatinization tendency. Perera et al. (1997) suggested that a possible

281

disruption of double helices due to the rotation of the flexible hydroxypropyl groups
within the amorphous region would leave fewer helices to melt in gelatinization.
Theoretically the greater swelling reduction resulting from cross-linking should delay
gelatinization, however cross-linking had very little effect on the gelatinization
parameters in contrast to the greater effect on pasting properties. Introduction of
cross-linking to hydroxypropylated potato starch increased gelatinization temperature
but other starches were unaffected.

Pasting properties
Pasting curves of native, hydroxypropylated, cross-linked, and cross-linked
hydroxypropylated tuber and root starches are presented in Fig. 2. and some results
summarized in Table 3. Pasting curves of potato, lotus, and sweet potato starch
showed

pasting characteristics typical to tuber and root starches with high peak

viscosity development and rapid shear thinning at higher temperature, whereas other
starches showed greater resistance to shear thinning and true yam starch behaved like
chemically cross-linked starches. Strong internal bonding forces in the structure of
these starches as evidenced by low swelling, delayed pasting onset, higher pasting
temperature, and higher H value in gelatinization could provide greater resistance
to break down of starch granules at higher temperature under shear stress, creating
more shear resistant starch paste. The highest peak viscosity was observed for lotus
starch followed by potato starch and the lowest in taro starch. Greater peak viscosity
development of lotus and potato starches is in agreement with observed higher
swelling factor for those starches. Generally potato and lotus starches showed similar
pasting characteristics such as early onset of pasting, higher peak viscosity, low
pasting temperature, and rapid loss of viscosity at higher temperature including
greater paste clarity (data not shown).

282

It has been reported that the location of phosphate groups mainly in the amylopectin
chains of potato starch were highly responsible for most of its functional properties
such as higher swelling and peak viscosity, rapid shear thinning, paste clarity, and
low retrogradation rate (Jane et al., 1996).Thus it can be speculated that the similar
functional properties in lotus starch could also be related to the presence of high level
phosphate groups in the starch granules. Lim et al. (1994) showed by 13 P-nuclear
magnetic resonance studies tuber and root starches contain high level of organic
phosphorus as phosphate monoesters in contrast to low content in cereal starches
where the most of phosphorus was found in the form of phospholipids.

Hydroxypropylation lead to an increase in peak viscosity of tuber and root starches


except for taro and gourd yam starches. Loosening of starch structure after
hydroxypropylation could permit granules to swell to a greater extent creating higher
peak viscosity. In case of taro and gourd yam starches it seems hydroxypropylation
resulted weaker granules than other starches by disrupting hydrogen bonds between
starch chains. Structurally weaker granules can undergo greater breakdown causing
rapid loss of peak viscosity resulting increased shear thinning. Slightly higher hot
paste viscosity shown in potato and true yam starches after hydroxypropylation could
be due to some physical interaction occurring between swollen fragile granules.

Cross-linking increased the peak viscosity of true yam starch but decreased the peak
viscosity of other starches. The properties of cross-linked starch vary widely
depending on the level of cross-linking. Mild cross-linking has been reported to
increase the peak viscosity but higher degree decreases it due to greater restriction of
granular swelling (Wuzburg and Szymanski 1970; Howling 1980). This implies that
true yam was cross-linked to a lesser extent than other starches under the given

283

conditions. A greater thermal pasting stability resulted in all the starches after crosslinking. Strengthened starch granule structure after cross-linking can resist rupture at
higher temperature to form thermally stable starch paste. Cross-linking of the
hydroxypropylated potato and taro starches increased peak viscosity more than in the
other corresponding starches. Cross-linking reinforces the starch granules while
hydroxypropylation increases the hydration thus cross-linking of hydroxypropylated
starch will allow more granular swelling to take place for a longer time before some
break down occurs for potato and taro starches. In all cases introduction of crosslinking to hydroxypropylated starches resulted in more thermally stable starch paste.

Cross-linking renders starch granules more resistant to low pH and thereby reduces
the shear thinning under acid conditions (Table 3). With the exception of taro and
lotus, decreased peak viscosity was observed at low pH. Although there was a peak
viscosity increase for taro and lotus starches at low pH, the pasting stability was
decreased (Table 3). Acidity could break the hydrogen bonds to bring more rapid
swelling creating higher peak viscosity but weaker granules due to the disruption of
hydrogen bonds could be more disintegrative at higher temperature. In case of potato
starch, low pH accelerated the reduction of peak viscosity but its pasting stability was
higher than the corresponding control starch (Table 3). Some physical interaction in
the digested starch particles may cause slight increase in hot paste viscosity and slight
increase of pasting stability. Similar to enzymic hydrolysis, increased accessibility to
starch granules acid in hydroxypropylated starches decreased the peak viscosity and
shear resistance of all the starches. As expected at low pH, cross-linking of
hydroxypropylated starches increased in both peak viscosity and paste thermal
stability more than corresponding hydroxypropylated starches.

284

Gel hardness
Gel hardness of native starches (Table 4) follow the order; true yam > gourd yam >
sweet potato > potato > taro > lotus. Stronger true yam and gourd yam gels are
consistent with the increased cold paste viscosity of these starches during the pasting
process. Initial gel firmness of freshly cooked starch is attributed to the short term
rapid reassociation of amylose chains within the cold starch paste. The textural and
mechanical properties of starch gel where swollen gelatinized amylopectin rich
granules are embedded in the continuous amylose gel matrix would depend on the
amylose concentration and rheological characteristics of the amylose gel matrix,
rigidity (deformability) of the swollen starch particles, volume fraction of the swollen
granules, and interaction between swollen particles and amylose matrix (Morris 1990;
Doublier et al., 1987; Eliasson, 1986).

Hydroxypropylation decreased the gel hardness of all the starches to a greater extent
and led to very weak gels in potato and taro starches where gel hardness was not
measurable under the conditions used. The action of hydroxypropylation in loosening
the starch granule structure would permit more granules to breakdown leaving weaker
cohesive swollen starch particles in the gel matrix leading to weaker starch gel.
Hydroxypropyl groups on amylose chains may prevent the amylose aggregation
interrupting the formation of junction zones.

Cross-linking increased the gel hardness of potato starch and decreased it in other
starches. Cross-linking that strengthened the starch granules will allow stronger rigid
granules to remain in the gel matrix. Also with the possible creation of more junction
zones with the introduced covelant bonds cross-linking could increase the gel
hardness. However, significant reduction of amylose leaching (Table 1) upon cross-

285

linking indicates decreased amylose concentration in the continuous network. This


latter effect seems more influential in determining gel hardness of other starches those
resulted weaker gel structure after cross-linking.

Cross-linking of hydroxypropylated potato, true yam and taro starches increased the
gel hardness but it decreased in gourd yam, lotus, and sweet potato starches. Although
cross-linking decreased the amylose leaching in hydroxypropylated potato, true yam
and taro starches it can reinforce the starch granules affecting the deformability of
gelatinized swollen particles in the gel matrix and also could introduce more junction
zones facilitating amylose aggregation. For other hydroxypropylated cross-linked
starches, the former effect may be more significant in determining gel hardness.

Enzymatic hydrolysis
The percent decrease of peak viscosity was used to measure the extent of enzymic
hydrolysis (Table 5 and Fig. 3).

For native starches the extent of enzymatic

hydrolysis follow the order; lotus > sweet potato > potato > taro > gourd yam > true
yam. Generally all the yam starches showed greater resistance to enzymic hydrolysis
than potato starch. Stronger starch structure of the above starches may restrict the
accessibility of enzyme into the starch granules. Hydroxypropylation increased the
digestibility of all the starches by facilitating enzyme accessibility as the
hydroxypropylation loosen the starch structure. The results showed that there was a
strong positive correlation between swelling ability and enzymic digestibility. A
larger reduction of enzymatic hydrolysis for cross-linked starches was expected
because strongly cross-bonded starch granules will resist the enzyme attack. This is
further evidenced by the reduction of enzymatic hydrolysis after introduction of
cross-linking to hydroxypropylated starches.

286

Retrogradation
Amylopectin retrogradation values as measured by DSC for native and modified
starches are presented (Table 6). Re-ordering of branched amylopectin is a longer
term slow process which occurs over a few days or several weeks. The extent of the
retrogradation of native starches followed the order; true yam > gourd yam > potato >
sweet potato > lotus > taro. The highest H value recorded for true yam was more
than twice that of potato starch. Taro starch retrograded to a much lesser extent. This
diversity in the retrogradation behavior should relate to the structural properties of
amylopectin that determine the quality and the quantity of the retrograded starch
crystals. Kohyama et al. (2004) reported amylopectin with higher ratio of longer side
chain retrograded more than amylopectin in shorter side chains. It has been reported
that cereal starches which have shorter average chain lengths retrograded to a lesser
extent compared with longer in potato and pea starches (Kalichevsky et al., 1990).
Miyazaki et al. (2000) demonstrated that amylose and much longer amylopectin in
sweet potato starch promotes the retrogradation while shorter amylopectin (DP 10)
inhibits retrogradation. However, Vandeputte et al. (2003) showed in a study of five
waxy rice starches that shorter length (DP 6-9) as well as longer chain length (DP >
25) inhibits retrogradation but amylopectin with DP 12-22 favors retrogradation.
After hydroxypropylation, retrogradation decreased in all the starches whereas crosslinking had no effect on retrogradation. After 10 days no retrogradation was shown
for the hydroypropylated tuber and root starches except in true yam starch (Table 6 &
Fig. 4). By inhibiting formation of hydrogen bonding, larger bulky hydroxylpropyl
groups attached to the amylopectin chains could prevent the re-ordering of
amylopectin chains however, it seems the covalent bonds that replaced the hydrogen
bonds in cross-linking had no effect on keeping amylopectin from reassociation, and
also cross-linking had no influence on the retrogradation of hydroxypropylated

287

starches. The greater reduction of amylopctin retrogradation observed after


hydroxypropylation suggests that although amylose has been shown to derivatize to a
greater extent (Shi and BeMiller, 2000 and Kavitha and BeMiller, 1998) the
derivatization could also take place in the amylopectin region to some greater extent.

10.4. Conclusions
This investigation of the properties of some underexploited tuber and root starches
has shown some useful functionalities, such as greater shear stability at higher
temperature and at low pH, and higher resistance to enzymic hydrolysis existed in
some starches. Properties that are generally undesirable such as higher retrogradation,
low swelling, low peak viscosity development, and low shear resistance can be
improved by hydroxylpropylation and cross-linking. In this context, cross-linking of
hydroxypropylated starches provides more desirable functional properties permitting
for a wider range of applications.

10.5. References
Choi, S.G. and Kerr, W.L. 2004. Swelling characteristics of native and chemically
modified wheat starches as a function of heating temperature and time. Starch 56:
181-189.
Chrastil, J. 1987. Improved colorimetric determination of amylose in starches or
flours. Carbohydr. Res. 159: 154-158.
Doublier, J.L., Llamas, G., and Le Meur, M. 1987. A rheological investigation of
cereal starch pastes and gels. Effects of pasting procedures. Carbohydr. Polym. 7:
251-275.
Galliard, T., and Bowler, P. 1987. Morphology and composition of starch. In: Starch:
Properties and Potential, T.Galliard (Ed.,), pp 55-79. Wiley, Chichester. England.
Eliasson, A.C. 1986. Viscoelastic behavior during the gelatinization of starch. .
Comparison of wheat, maize, potato, and waxy-barley starches. J. Text. Stud. 17: 253265.

288

Hoover, R. 2001. Composition, molecular structure, and physicochemical properties


of tuber and root starches: a review Carbohydr. Polym. 45: 253-267.
Hoover, R., Hannouz, D., and Sosulski, F.W. 1988. Effect of hydroxypropylation on
thermal properties, starch digestibility and freeze-thaw stability of field pea (Pisum
sativum cv Trapper) starch. Starke 40: 383-387.
Hoover, R., and Hadziyev, D. 1981. Characterization of potato starch and its
monoglyceride complexes. Starke 33: 290-300.
Howling, D. 1980. The influence of the structure of starch on its rheological
properties. Food Chem. 6: 51-61.
Jane, J., Kasemsuwan, T., and Chen, J.F. 1996. Phosphorus in rice and other starches.
Cereal Foods World 41: 827-832.
Kalichevsky, M.T., Orford, P.D., and Ring, S.G. 1990. The retrogradation and
gelation of amylopectins from various botanical sources. Carbohydr. Res. 198: 49-55.
Kavitha, R. and BeMiller, J.N. 1998. Characterization of hydroxypropylated potato
starch. Carbohydr. Polym. 37: 115-121.
Kim, H.R., and Eliasson, A.C. 1993. The influence of molar substitution on thermal
transition properties of hydroxypropyl potato starches. Carbohydr. Polym. 22:331335.
Kohyama, K., Matsuki, J., Yasui, T., Sasaki, T. 2004. A differential thermal analysis
of the gelatinization and retrogradation of wheat starches with different amylopectin
chain lengths. Carbohydr. Polym. 58: 71-77.
Li, W.D., Huang, J.C., and Corke, H. (2000). Effect of -cyclodextrin on pasting
properties of wheat starch. Nahrung-Food 44: 164-167.
Lim, S.T., Kasemsuwan, T., and Jane, J. 1994. Characterization of phosphorus in
starches using 31P-NMR spectroscopy. Cereal Chem. 71: 488-493.
Liu, H.J., Ramsden, L., and Corke, H. 1999. Physical properties and enzymatic
digestibility of hydroxypropylated ae, wx, and normal maize starches. Carbohydr.
Polym. 40: 175-182.
Liu, H.J., Ramsden, L., and Corke, H. 1999. Physical properties of cross-linked and
acetylated normal and waxy rice starch. Starch/Starke 51: 249-252.
Miyazaki, K.I., Kumamoto, T.K., Kagoshima, K.K., and Komamoto, O.Y. 2000.
Retrogradation of sweet potato starch. Starch/ Starke 52: 13-17.
Morris, M.J. 1990. Starch gelation and retrogradation. Trends in Food Science &
Technology 1: 2-6.
Pal, J., Singhal, P.S., and Kulkarni, P.R., 2002. Physicochemical properties of
hydroxypropyl derivative from corn and amaranth starch. Carbohydr. Polym .48: 4953.
289

Perera, C., Hoover, R., and Martin, A.M. 1997. The effect of hydroxypropylation on
the structure and physicochemical properties of native, defatted and heat-moisture
treated potato starches. Food Research International 30: 235-247.
Seow, C.G., and Thevamalar, K. 1993. Internal plasticization of granular rice starch
by hydroxypropylation: effect on phase transition associated with gelatinization.
Starke 45: 85-88.
Shi, X., and BeMiller, J.N. 2002. Aqueous leaching of derivatized amylose from
hydroxypropylated common corn starch granules. Starch/Starke 54: 16-19.
Shi, X., and BeMiller, J.N. 2000. Effect of sulfate and citrate salts on derivatization of
amylose and amylopectin during hydroxypropylation of corn starch. Carbohydr.
Polym. 43: 333-336.
Tester, R.F., and Morrison, W.R. 1990a. Swelling and gelatinization of cereal
starches. 1. effects of amylopectin, amylose, and lipids. Cereal Chem. 67: 551-557.
Vandeputte, G.E., Vermeylen, R., Geeroms, J., and Delcour, J.A. 2003. Rice starches.
. Structural aspects provide insight in amylopectin retrogradation properties and gel
texture. J. Cereal Sci. 38: 61-68.
Woo, K. Seib, P.A. 1997. Cross-linking of wheat starch and hydroxypropylated wheat
starch in alkaline slurry with sodium trimetaphosphate. Carbohydr. Polym. 33: 263271.
Wurzburg, O.B., and Szymanski, C.D. 1970. Modified starches for the food industry.
J. Agric. Food Chem. 18: 997-1001.

290

Table 1. Swelling factor and amylose leaching at 85 C of native,


hydroxypropylated (HP), cross-linked (CL), and hydroxypropylated cross-linked
(HP-CL) tuber and root starches
Starch
Treatment
Swelling factor
Amylose leaching (%)
Potato
Native
50.20.3a
16.40.2
HP
60.10.2
15.10.2
CL
20.20.1
4.00.2
HP-CL
24.50.2
5.00.2
True yam

Native
HP
CL
HP-CL

18.10.4
40.10.1
16.70.2
26.40.1

17.20.2
19.20.5
9.10.1
9.70.1

Gourd yam

Native
HP
CL
HP-CL

20.20.4
23.20.1
15.00.1
15.40.0

12.10.3
14.10.2
6.80.3
9.40.4

Taro

Native
HP
CL
HP-CL

35.00.2
40.10.1
19.00.4
23.00.1

13.40.2
14.30.5
6.40.4
6.30.2

Lotus

Native
HP
CL
HP-CL

45.00.5
47.10.6
12.70.2
15.40.1

12.50.3
9.60.1
1.00.1
1.00.1

Sweet potato

Native
39.00.4
12.30.4
HP
43.20.1
13.10..3
CL
18.80.3
4.10.2
HP-CL
21.00.1
5.30.1
a
Values are means of triplicate determinations standard deviation.

291

Table 2. Gelatinization enthalpy (HG) of native, hydroxypropylated (HP), crosslinked (CL), and hydroxypropylated cross-linked (HP-CL) tuber and root starches
HG (J/g)
Starch
Treatment
Potato
Native
15.10.1a
HP
13.60.2
15.30.2
CL
HP-CL
13.70.1
True yam

Native
HP
CL
HP-CL

17.50.3
14.10.1
16.80.1
14.40.1

Gourd yam

Native
HP
CL
HP-CL

16.10.2
13.80.1
16.40.3
13.60.2

Taro

Native
HP
CL
HP-CL

15.70.3
14.00.2
16.00.1
13.80.1

Lotus

Native
HP
CL
HP-CL

14.20.2
12.80.1
14.00.1
12.70.2

Sweet potato

Native
15.10.3
12.30.1
HP
14.90.2
CL
12.20.1
HP-CL
a
Values are means of triplicate determinations standard deviations.

292

Table 3. Pasting propertiesa of hydroxypropylated (HP), cross-linked (CL), and


hydroxypropylated cross-linked (HP-CL) tuber and root starches in distilled
water and hydrochloric acid solution (pH=3)
Distilled water
Acid solution (pH=3)
Starch
Treatment PV
HPV
CPV
PV
HPV
CPV
b
Potato
Native
5751.0
1561.6 2281.4 4881.2 1490.8 2180.6
HP
6780.7
1630.6 2120.6 5070.9 1291.7 1621.4
CL
5031.2
5031.2 7360.9 4670.7 4562.6 7561.2
HP-CL
7321.3
6681.7 7261.7 6621.3 5021.2 7471.2
True yam
Native
1591.9
1120.9 2721.1 1531.7 1060.7 2552.2
HP
2900.7
1280.6 1821.8 2192.7 1131.1 1540.9
CL
1731.4
1481.1 2940.7 1631.4 1420.8 2520.7
HP-CL
2801.0
2021.2 4041.8 2741.0 1901.2 3741.8
Gourd yam Native
2640.9
1770.8 3121.5 2480.9 1470.8 2411.3
HP
1981.9
922.2
1651.2 1761.9 672.2
1151.2
CL
720.7
720.7
1201.0 1400.7 1410.7 2081.0
HP-CL
1831.5
1631.1 2901.4 2321.5 2121.1 3520.4
Taro
Native
1651.2
880.7
1471.9 1931.2 821.2
1381.9
HP
1320.8
631.3
1210.5 1320.8 542.3
880.5
CL
1441.4
1190.9 2100.4 2031.4 1771.9 2851.4
HP-CL
2181.1
1471.8 2520.9 2400.8 1751.8 2910.9
Lotus
Native
6360.9
1201.4 2221.7 6571.6 881.4
1491.7
HP
7201.1
690.8
2530.8 4201.1 500.8
1200.8
CL
941.6
1450.4 1871.6 1871.6 3340.4
941.6
HP-CL
1350.9 2891.2 2500.9 2500.9 5631.2
1360.7
Sweet
Native
1591.7 2381.1 3221.6 1390.7 2121.2
3350.6
potato
HP
3601.5
1460.5 2150.7 3451.5 1360.9 2040.9
781.1
1271.3 1470.9 1472.1 2331.3
CL
781.1
HP-CL
2531.9 4951.6 3460.6 3091.9 5631.6
2840.8
a
PV=peak viscosity, HPV=hot paste viscosity, CPV=cold paste viscosity.
b
Values are means of triplicate determinations standard deviation.

293

Table 4. Gel hardness of native, hydroxypropylated (HP), cross-linked (CL),


and hydroxypropylated cross-linked (HP-CL) tuber and root starches
Starch
Treatment
Gel hardness (g)
Potato
Native
270.2a
HP
CL
440.3
HP-CL
130.2
True yam

Native
HP
CL
HP-CL

1400.2
120.5
710.6
660.2

Gourd yam

Native
HP
CL
HP-CL

660.1
160.3
310.2
90.4

Taro

Native
HP
CL
HP-CL

120.2
160.3
80.4

Lotus

Native
HP
CL
HP-CL

120.5
100.6
70.4
70.4

Sweet potato

Native
280.5
HP
130.5
CL
90.6
HP-CL
70.4
a
Values are means of triplicate determinations standard deviation.

294

Table 5. Percent decrease of peak viscosity of native, hydroxypropylated (HP),


cross-linked (CL), and hydroxypropylated cross-linked (HP-CL) tuber and root
starches after -amylase treatment
Starch
Treatment
% decrease of PV
Potato
Native
870.8a
HP
890.3
CL
400.4
HP-CL
540.9
True yam

Native
HP
CL
HP-CL

100.7
680.2
70.6
420.3

Gourd yam

Native
HP
CL
HP-CL

241.1
950.7
110.8
830.4

Taro

Native
HP
CL
HP-CL

531.3
890.6
140.5
510.4

Lotus

Native
HP
CL
HP-CL

910.8
931.2
130.7
160.4

Sweet potato

Native
900.5
HP
910.3
CL
191.2
HP-CL
650.5
a
Values are means of duplicate determinations standard deviation.

295

Table 6. Retrogradation enthalpy (HR) of native, hydroxypropylated (HP),


cross-linked (CL), and hydroxypropylated cross-linked (HP-CL) tuber and root
starches
Starch
Treatment
HR (J/g)
Potato
Native
3.70.3 a
HP
0.0
CL
3.60.2
HP-CL
0.0
True Yam

Native
HP
CL
HP-CL

9.20.1
2.40.4
8.90.2
2.50.2

Gourd Yam

Native
HP
CL
HP-CL

6.20.1
0.0
5.90.1
0.0

Taro

Native
HP
CL
HP-CL

0.40.1
0.0
0.30.1
0.0

Lotus

Native
HP
CL
HP-CL

1.00.1
0.0
0.90.1
0.0

Native
2.80.2
HP
0.0
CL
2.90.1
HP-CL
0.0
a
Values are means of triplicate determinations standard deviation.
Sweet potato

296

-1.8

-1.9

-2.0

-2.1

-2.2

-2.3

hp

hp

-2.5

-2.4

cl

-2.7

cl

-2.6

-2.9

hp-cl

-2.8

hp-cl
-3.1

-3.0

-3.2

-3.3

40

50

60

70

80

90

100

-1.8

-3.5

40
-1.8

50

60

70

80

90

-2.0

100

110

-2.0

-2.2

hp

Heat flow (W/g)

-2.2

-2.4

hp
-2.6

cl

-2.4

cl

-2.8

-2.6

-3.0

hp-cl
hp-cl

-3.2

-2.8

-3.4

-3.0

50

60

70

80

90

100

110

-3.6

50

60

70

80

90

100

-1.8

-2.1

-2.0

-2.3

hp

-2.2

hp
-2.5

-2.4

cl

cl
-2.7

-2.6

hp-cl
hp-cl
-2.9

-2.8

-3.0

40

50

60

70

80

90

-3.1

40

50

60

70

80

Temperature (C)
Fig. 1. DSC curves of native, hydroxypropylated, cross-linked,
hydroxypropylated cross-linked potato (A), true yam (B), gourd yam (C),
taro (D), lotus (E), and sweet potato (F). n=native, hp=hydroxypropylated,
cl=cross-linked, hp-cl= hydroxypropylated cross-linked.

297

90

100

hp-cl

600

100 500

100

80 400

80

hp

hp-cl
60 300

60

40 200

40

cl

400

200

hp

20 100

A
0
5

10

15

0
0

20

cl

20

25

320

100

0
0

10

15

20

25

100

320

80

240

80
240

hp-cl

hp-cl

cl

160

60

160

hp

40

80

40

80

cl

20

hp

0
0

10

15

20

20

25

0
0

10

15

20

25

100

800

hp

600

80

100

450

80

hp
hp-cl
n

60

60

300
400

40

hp-cl

40

150

200
20

20

cl

cl

0
0

10

15

20

25

0
0

10

15

20

Time ( Mins)

Fig. 2. RVA curves of native, hydroxypropylated, cross-linked, and


hydroxypropylated cross-linked potato (A), true yam (B), gourd yam (C), taro
(D), lotus (E), and sweet potato (F) starches. n=native, hp=hydroxypropylated,
hp-cl=hydroxypropylatedcross-linked.

298

25

Temperature (C)

Viscosity (RVU)

60

100

100

320

B
cl

hp-cl

600

80
240

E+n

n
60
400

Temperature (C)

Viscosity (RVU)

80

60
160

hp

hp-cl

40

200

80

E+cl

E+hp

hp

E+n

E+hp-cl

0
5

10

E+cl

20

E+hp-cl

40

15

20

25

cl
20

E+hp

0
0

10

15

20

25

Time (Mins)
Fig. 3. RVA curves of native, hydroxypropylated, cross-linked, and hydroxypropylated cross-linked potato (A), gourd yam (B) before
and after enzyme treatment. n=native, hp=hydroxypropylated, cl=cross-linked, hp-cl=hydroxypropylated cross-linked. E=-amylase
enzyme.

299

-1.6

-1.8

-1.8

Heat flow (W/g)

-2.0

hp

hp
-2.0

cl

cl

-2.2
-2.2

hp-cl

hp-cl

-2.4

-2.4

50

60

70

80

90

100

40

50

60

70

80

90

100

110

Temperature (C)
Fig. 4. DSC curves of retrograded native, hydroxypropylated, cross-linked and hydroxypropylated cross-linked potato (A), true yam (B)
starches. n=native, hp=hydroxypropylated, cl=cross-linked, hp-cl=hydroxypropylated cross-linked.

300

Summary and Direction for Future Studies


This study showed that -cyclodextrin can disrupt the amylose-lipid complex as well
as inhibit its formation within starch granules, affecting a range of functional
properties. The amylose-lipid complex has shown resistance to acid hydrolysis and
heat-moisture treatment but was susceptible to hydrolysis at higher alkalinity.
Understanding of supermolecular structure and transition of amylose-lipid complex is
technologically important as it affects many functional characteristics of starch and
quality of starch- based food products. -cyclodextrin greatly affects the behavior of
amylose-lipid complex by complexing with starch lipids. So it is worth testing the
effect of -cyclodextrin on dough for bread, noodle and steamed bread making
because interaction of -cyclodextrin with emulsifiers used in the dough could
influence the textural properties of both dough and final products. Furthermore, the
applicability of -cyclodextrin in the processing of starch hydrolysates should be
investigated since it has a great potential to improve the effectiveness of the
hydrolysis, color, aroma and filtration via the disruption of amylose-lipid complex. In
addition, studies should focus on the use of -cyclodextrin as a multi-functional
processing agent for the above food systems. For example flavor retention and
modification of the flavor delivery to the system can be achieved while improving the
effectiveness of processing steps. Studies should further focus on the interaction of
modified -cyclodextrin with different degree of modification (in this study we used
only MS=0.8 hydroxypropyl -cyclodextrin). Isolation and characterization of
complexes formed by -cyclodextrin starch-lipid interaction may reveal more
information on -cyclodextrin-starch interaction. It is of interest to study the multiple
interaction of -cyclodextrin-lipids (different types)-starches (based on granular size)
and its influence on retrogradation and functional properties of starch. Further, the
changes in X-ray diffraction pattern of the V-polymorph (amylose-lipid complex) of

301

cereal starches in the presence of cyclodextrin could be used to examine the influence
of cyclodextrin on amylose-lipid complex.

Prior heat-moisture treatment before hydroxypropylation causes more derivatization


in the crystalline region and results in very high final paste viscosity with greater
pasting stability. This indicates that prior heat-moisture treatment could be applied to
alter the reaction site of hydroxypropylation achieving some novel starch properties.
In this study only one combination of heat and moisture was examined. Thus it is
worthwhile testing many combinations of heat and moisture in heat-moisture
treatment followed by hydroxypropylation with different degree of modification.
Microscopic

examination

of

gels

resulting

from

heat-moisture

treated

hydroxypropylated starch would provide more information on why it resulted in


higher final viscosity with greater pasting stability.

Cleavage of macromolecules in acid treatment increases the proportion of the short


chain fraction which permits realignments and self-association of the molecules to
form double helical-like structures affecting retrogradation, gelatinization and
gelation properties of starch and it could also reduce the number of reaction sites for
acetylation. It can be expected that the sequence of modification could affect the
structural and functional properties of starches. For example acid modification before
and after acetylation could result in different outcomes. Therefore it is worth studying
the effect of sequence of modification and its influence on starch structure and
functional properties. More studies should focus on the change of the reaction site of
hydroxypropylation and acetylation since it could result in novel starch properties
while reducing the amount of reaction reagents needed to achieve the desirable
chemical bonding for the target properties.

302

Mixing of starches showed some promising results towards their use as an alternative
to chemical modification. This study showed the possibility of formulating starch
blends from unmodified and physically modified (heat-moisture treated) starches
resulting in some useful functional characteristics such as higher pasting stability and
soft starch gel. Only a few blends were tested. There is a great deal of work remaining
to be done in this interesting area. Mixing of starches from diverse botanical origin in
different combinations may provide some useful products that can replace those
derived from chemical modification. Some microscopic studies should be carried out
to examine any interactions at granular or molecular levels between individual
components in the mixture at gelatinization and pasting. Understanding of such
interactions may advance the knowledge of the fundamental causes that influence
unique properties exhibited in some starch mixtures such as non-additive pasting
characteristics and formation of very soft gel.

Some excellent properties such as higher gel hardness and resistance to shearthinning exhibited in some under-exploited tuber and root starches are promising for
some food applications. However, many tuber and root starches are not widely used
in food applications due to their poor functional properties. Presently chemical
modification is widely used to tailor the properties of potato and cassava starches.
This study showed that mixing of starches may be an alternative to chemical
modification for altering gelatinizing, pasting and gelling properties. Thus it is worth
testing properties of mixtures of native and physically modified (heat-moisture treated)
under-exploited tuber and root starches to determine their potential uses in food
applications.

303

You might also like