Download as pdf or txt
Download as pdf or txt
You are on page 1of 32

Major changes in cereal

biopolymers during ready-to-eat 6


cereal processing
Victor T. Huang 1, a , Alicia A. Perdon 2, a
1
General Mills, Golden Valley, MN, United States; 2Kellogg Company, Battle Creek,
Michigan, United States

Introduction
Whole cereal kernels, cut grains, or even milled-grain flours can be transformed into a
relatively uniform structural matrix of a breakfast cereal, often through the effects of
temperature, moisture, shear, and time treatments. Fig. 6.1 illustrates that processing
can induce either mostly positive or sometimes negative changes to a food matrix at
the molecular, microstructural, and macrostructural levels. Understanding the funda-
mental changes that take place during processing at the molecular and microstructural
levels should enable product developers to design novel breakfast cereals of desired
functional and textural characteristics, and to optimize existing products and processes.
Specifically, the “food polymer science” approach will be used here for describing
those major changes. In the unit-operation chapters, dynamic mapping on a state
diagram will be used to depict the changes in a product matrix during specific cereal
processing unit operations.

Moisture management and food polymer science


In the early 1980s, Drs. Louise Slade and Harry Levine introduced the “food polymer
science” concept, whereby a food system is viewed as a polymer matrix (Fig. 6.2)
(Levine and Slade, 1986; Slade and Levine, 1987a,b,c). As a solid, a polymer matrix
can be either crystalline, where the molecules are packed together in an organized crys-
tal lattice, or amorphous, where the molecules are randomly arranged. An amorphous
matrix can be either glassy or rubbery, depending on temperature and composition,
including moisture content and the amounts and molecular weights of other plasti-
cizers in addition to water.

a
Retired.

Breakfast Cereals and How They Are Made. https://doi.org/10.1016/B978-0-12-812043-9.00006-0


Copyright © 2020 Cereals & Grains Association. Published by Elsevier Inc. All rights reserved.
110
The food structure hierarchy
nm mm mm

Domain Nano Macro


Molecular

Levels Ingredient
Atomic System Finished food

Breakfast Cereals and How They Are Made


Water, cellulose, polymers Water, sugar, grains, Formula, Appearance,
malt extract process, equipment taste, texture

Tools Sensory and physical analyses


Chemistry and “nano” science Polymer and colloidal science

Applied and practical research

Basic and fundamental research


Figure 6.1 The food structure hierarchy with multiscale model.
Major changes in cereal biopolymers during ready-to-eat cereal processing 111

Water

Polymer matrix Sugar, salt, humectant

Figure 6.2 Anatomy of a food system.

Moisture content and water activity definitions


Understanding water and its management is very important, because it affects food
processability, sensory quality, and storage stability. Moisture content is defined as
the amount of water in a food product and can be expressed on a wet or dry basis
(Fig. 6.3).
Typically, moisture content is determined by dehydrating a product for a period of
time at a temperature above the boiling point of water. However, drying at excessively
high temperature may result in product degradation, especially for sugar-containing
cereals. In such a case, drying at a lower temperature under vacuum is recommended.
In the 1950s, the concept of water activity, aw, was proposed for determining food
stability (Scott, 1957, 1958; Salwin, 1959; Labuza et al., 1970; Labuza, 1976, 1980).
aw is defined as the ratio of the partial vapor pressure of water in a food to the vapor
pressure of pure water at the same temperature and pressure (Fig. 6.4). Thus, by
definition, its values range from 0 to 1. The above definition of water activity assumes
that a food is at equilibrium and in equilibrium with the surrounding atmosphere.

Heat w/ or w/o
vacuum

Initial sample Dried sample “Free” and “bound” water

“Bound” “Free”
water water
Matrix

Mass of free and bound water


Moisture content, % as is or wet basis = 100 ×
Mass of initial sample

Mass of free and bound water


Moisture content, % dry basis = 100 ×
Mass of dried sample

Figure 6.3 Moisture content measurement and calculations.


112
Humidity and
temperature
probe

Sample
Sample cup
holder
Initial condition at T At equilibrium at T
pure water- 20 molecules, liquid pure water- 10 molecules, liquid
Shafiur Rahman, 1995
surrounding air- 10 molecules, gas

Water activity, aw, is the ratio of the vapor pressure exerted


psample
aw = by the “free water” in an equilibrated sample, p sample, to the
p0 vapor pressure exerted by pure liquid water in that sample,
p0, measured at constant temperature and pressure

Breakfast Cereals and How They Are Made


“Free”
p0 water vapor p water vapor at equilibrium “Bound” water
water Matrix

Water Sample

p0 pure liquid water p water in food

p0 p sample Equlibrium relative humidity


aw = =1 aw = ∼
=
p0 p0 100

Figure 6.4 Illustration of the definition and measurement of water activity.


Major changes in cereal biopolymers during ready-to-eat cereal processing 113

Food polymer science


In the food polymer science approach, understanding the concepts of “water
dynamics“ and “glass dynamics“ in a food system is very important (Slade and Levine,
1987a, 1991b). Polymers can have different degrees of crystallinity. In a food, the
matrix can be partially crystalline and partially amorphous. Their state transitions
are important during processing. The glassy solid to rubbery liquid transformation
of an amorphous material is called the glass transition. The temperature range over
which this occurs is called the glass transition temperature, or Tg. Water is the most
effective plasticizer and can depress the glass transition temperature of an amorphous
food material (Levine and Slade, 1988b). Water-compatible, low molecular weight
solutes such as sugars are also good plasticizers. Fig. 6.5 is a generic “state diagram”
showing that as the moisture content of a food polymer increases, its Tg decreases.
Changes in material properties accompany the glass transition (Fig. 6.6) and form
the basis for measuring Tg (White and Cakebread, 1966). Typical thermal analysis
techniques are differential scanning calorimetry (DSC), dynamic mechanical thermal
analysis, and thermomechanical analysis. DSC, which measures changes in specific
heat, is the most common method. The glass transition occurs over a range of temper-
atures, from onset, midpoint, to completion. Most studies use the midpoint to represent
T g.
Five different regions of viscoelasticity (Fig. 6.7) can be identified to describe the
rheological properties of any food polymer, which depend on its temperature and
moisture content: glassy solid, leathery, rubbery plateau, elastic or rubbery flow,
and viscous liquid flow regions (Levine and Slade, 1988b, 1989; Sperling, 2005).
These changes occur for different ingredient constituents during cereal processing.

100
Water dynamics Unstable
reactive
Mobility ,Time
Temperature, °C

25

0
Glass
–25 transition
Tg
Glass dynamics
Stable
unreactive Mobility ,Time
–100
0 25 50 75 100
Moisture content, %
Figure 6.5 Generic “state diagram” illustrating the “food polymer science” approach to
moisture management. At temperatures below glass transition, molecular mobility decreases
and reactions occur very slow (longer time). At temperatures above glass transition, molecular
mobility increases and reactions occur very fast (shorter time).
114 Breakfast Cereals and How They Are Made

Tg

Temperature < Tg Temperature > Tg

Glassy Rubbery
High modulus Low modulus
Low diffusivity High diffusivity
Low expansion coefficient High expansion coefficient
Low heat capacity High heat capacity

Temperature
Figure 6.6 Changes in material properties accompanying glass transition. A material is glassy,
has high modulus, low expansion coefficient, and low heat capacity at temperatures below its Tg,
with low molecular diffusivity and mobility. It becomes rubbery, has low modulus, high
expansion coefficient, and high heat capacity as it goes through glass transition.
log Modulus

Tg Tm
Temperature
Figure 6.7 Five regions of viscoelasticity relative to Tg and Tm

Transitions of major constituents


Cereal grains are the major ingredients of breakfast cereals. Starch is the major compo-
nent of cereal grains (% dry weight basis), followed by fiber and protein, as shown in
Table 6.1. Cereal grains contain various cell wall polysaccharides, such as arabinox-
ylans (AX) (0.2%e10.8%), b-glucans (0%e7%), cellulose (0.3%e3.9%), gluco-
mannans, xyloglucans, pectic polysaccharides, callose, fructans, and arabinogalactan
proteins (Collins et al., 2010; Choct, 1997). Cellulose, being insoluble and inert,
is not affected by temperature, moisture content, and shear during breakfast cereal
Major changes in cereal biopolymers during ready-to-eat cereal processing 115

Table 6.1 Components of cereal grains (%db) (Lasztity, 1995).


Cereal grain Starch Protein Fiber Fat Ash

Wheat 71.9 12.2 1.9 1.9 1.7


Oats 55.5 11.3 10.9 5.8 3.2
Corn 70.0 9.1 2.3 4.6 1.3
Rice 75.8 8.1 0.5 1.2 1.4
Barley 73.5 10.9 4.3 2.3 2.4
Rye 71.9 11.6 1.9 1.7 2.0
Sorghum 65.0 11.0 4.9 3.5 2.6
Buckwheat 64.0 11.0 11.0 2.4 1.8
Millet 58.9 10.3 8.7 4.5 4.7

processing. Thus, only the two most abundant polysaccharides, AX and b-glucans,
will be discussed here. Changes in these components are the most important transfor-
mations during cereal processing.

Starch: structure and transformation


Starch structure
Starch is the major polymeric component of cereal grains and exists in the form of
starch granules. Starch morphology (granule size and shape), composition, and prop-
erties vary among the cereal grains, as shown in Table 6.2.
Starch is a partially crystalline, partially amorphous glucose polymer (Slade and
Levine,1989; Biliaderis, 1991). It is composed of two fractions, amylose and amylo-
pectin. Amylose is the linear fraction, with a-1,4 linkages and an average molecular
weight of <0.5  106 Da. It gives a blue color with iodine, with a maximum wave-
length at 640 nm. Amylopectin is the branched polymeric glucose fraction, with linear
a-1,4 and branched a-1,6 linkages and an average molecular weight of
50e500  106 Da. It gives a reddish brown color with iodine, with a maximum
wavelength at 550 nm.
The amylose and amylopectin fractions are systematically arranged in the starch
granules: the rigid crystalline phase, is mostly composed of the amylopectin branches,
the rigid amorphous phase, which is the region adjacent to the crystalline phase, and
the bulk amorphous phase, which is mostly composed of amylose (Fig. 6.8).
116
Table 6.2 Starch morphology and composition of different grains.

High
amylose
Dent corn corn Wheat Oat Rice Potato Tapioca

Polarized light
microscopy

Scanning
electron
microscopy

Granule shape Polygonal, Polygonal, Round, Polyhedral, irregular Polygonal, spherical Oval, Oval,
round round, lenticular compound compound spherical truncated,

Breakfast Cereals and How They Are Made


irregular granules granules “kettle-
drum”
Granule 5e30 5e30 1e45 7e10 1e3 5e100 4e35
diameter, pm
Amylose 25 55e70 25 17e28 0e30 20 17
content, %
Gelatinization 62e80 140e160 58e64 55e65 68e78 58e65 52e65
temperature,
C
Major changes in cereal biopolymers during ready-to-eat cereal processing
Amylopectin

Amylose

Figure 6.8 Morphological and microstructural levels within rice starch.

117
118 Breakfast Cereals and How They Are Made

Starch glass transition


As a polymer, the properties of starch are affected by temperature and the presence of
plasticizers. Water is the most effective plasticizer in such a system, and starch can un-
dergo a glass transition when water is added and/or the material is heated (Slade and
Levine, 1987c,1989). Only the amorphous phase of the starch is directly affected at
this stage. Fig. 6.9 shows the plasticizing effect of water in decreasing the Tg of rice
starch (Perdon et al., 2000).

Starch swelling
Under excess water conditions and at around 104 F/40 C, the amorphous regions of
raw starch granules absorb water. As a result, the granules swell and increase slightly
in size (Tester and Morrison, 1990). However, this is still, at this stage, a reversible
phenomenon.

Starch annealing
Annealing describes the process of perfecting crystalline structures, when starch in
excess water is subjected to a temperature above its glass transition temperature but
below its gelatinization temperature (Slade and Levine, 1989; Tester and Debon,
2000). Annealing is reported to enhance granule stability, crystalline perfection, starch
chain interactions, and increase the gelatinization temperature, but decrease the gela-
tinization temperature range, granule swelling, and amylose leaching, which can affect
the subsequent gelatinization and retrogradation processes.

Starch gelatinization
On further heating above the glass transition temperature, crystalline starch can melt
when water is limiting, or can gelatinize when water is in excess (Slade and Levine,

100

80

60
Temperature, °C

40

20

0 Glass transition, Tg

–20

–40
0 10 20 30 40 50 60 70 80
Moisture content, % db
Figure 6.9 Plasticizing effect of water on the glass transition, Tg, of rice starch (Perdon, 1999).
Major changes in cereal biopolymers during ready-to-eat cereal processing 119

1989). Starch gelatinization has been described as the water-mediated collapse (disrup-
tion) of molecular order within the starch granule, manifested by irreversible changes
in properties. It involves the loss of birefringence and crystallinity, hydration and
swelling of the granules, rupture of the granules, amylose dissolution and diffusion,
and an increase in viscosity/clarity (Atwell et al., 1988). These changes occur over
different temperature ranges and can be monitored (Table 6.3).
Penetration of water results in increased randomness within the starch granule struc-
ture and a decreased number and size of the crystalline regions. Water is first absorbed
into the bulk amorphous regions of the starch granule, which leads to swelling. Water
is then absorbed by and plasticizes the rigid amorphous regions of the amylopectin.
Heat causes the crystalline regions to melt and transform into an amorphous liquid
state (Fig. 6.10). During gelatinization, the intermolecular bonds between starch
molecules break down. Physical stresses caused by the swelling phenomenon eventu-
ally interrupt the structured organization within the granule and allow the leaching of
amylose molecules into the surrounding water. Table 6.3 shows the physical changes
that can occur during gelatinization, and the methods used to monitor those changes.
The gelatinization temperature of starch depends on starch plant origin, the amount
of water present, pH, and type and concentration of salts in the system. Under limited
moisture conditions, usually less than 30%, the crystalline regions of starch can melt
(Zeleznak and Hoseney, 1987). As the moisture content increases, plasticization by
water leads to the melting temperature decreasing. This plasticization by water also
lowers the Tg of pregelatinized starch, compared with the corresponding Tg of raw
starch at similar moisture contents (Fig. 6.11) (Zeleznak and Hoseney, 1987; Seow
and Thevamalar, 1993).

Table 6.3 Monitoring morphological and functionality changes during wheat starch
gelatinization in excess water.

Wheat starch gelatinization 50 60 70 80 90 100∞C Method of analysis

Loss of birefringence Polarized light microscopy

Loss of melting endotherm Differential scanning calorimetry

Loss of X-ray diffraction X-ray diffraction

Solubilization Solvent retention capacity

Swelling power Solvent retention capacity

Light microscopy, scanning electron


Morphological changes microscopy

Paste consistency changes Viscoamylography (RVA, brabender)

50 60 70 80 90 100∞C
120 Breakfast Cereals and How They Are Made

Crystalline region

Heat,
water

Figure 6.10 Disordering of amylopectin crystalline regions as the result of heat treatment in the
presence of excess water.

160

140

120
Temperature (°C)

100

80
Raw starch
60

40
Pregelatinized starch
20

0
5.0 10.0 15.0 20.0 25.0 30.0
Moisture (%)

Figure 6.11 Comparison of raw and pregelatinized wheat starch Tgs at different moisture
contents (Zeleznak and Hoseney, 1987).

During cereal processing, the amount of water present, as well as the source of en-
ergy used to gelatinize the starch, affect starch granule integrity (Fig. 6.12). Cooking in
limited water, without shear, maintains granule shape, whereas processing with shear
destroys the starch granule structure. These changes will have an effect on finished
product quality, especially texture and bowl life.
A state diagram is a plot of the glass transition and melting temperatures (Tg and
Tm, respectively) of a material at corresponding moisture contents (Fig. 6.13) (Levine
and Slade, 1986, 1988a, 1990, 1992; Roos and Karel, 1991; Slade and Levine, 1989,
1991a, 1995a). Such diagrams are used for dynamic mapping to predict material prop-
erty changes under different moisture content and temperature conditions experienced
during processing. Later in this chapter, an example of dynamic mapping during
different cereals unit operations will be shown.
Major changes in cereal biopolymers during ready-to-eat cereal processing 121

Limited water, no shear

Limited water, with shear Excess water, with shear


Figure 6.12 Effects of process conditions on starch granule integrity.

250
Starch melt, Tm
200
Temperature, °C

150

100
Glass transition, Tg

50

–50
0 10 20 30 40 50 60 70 80
Moisture content, % db

Figure 6.13 The state diagram of rice starch (Perdon, 1999).

Starch retrogradation
Starch retrogradation occurs when molecules comprising gelatinized starch begin to
(re)associate into an ordered structure (Biliaderis, 1992; Slade and Levine, 1987c)
as shown in Fig. 6.14. In this initial phase, two or more starch chains may form a sim-
ple juncture point, which then may develop into more extensively ordered regions. Ul-
timately, under favorable conditions, a crystalline order can appear.
122 Breakfast Cereals and How They Are Made

Temperature
moisture content
time

Crystalline Amorphous
melting transition at Tm glass transition at Tg
Figure 6.14 Retrogradation of gelatinized starch.

The most preferred way to analyze starch retrogradation is by using DSC (Slade and
Levine, 1987c; Biliaderis and Juliano, 1993; Perdon et al., 1999). A retrograded
amylopectin starch exhibits a melting endotherm peak at temperatures ranging from
about 113 F/45 C to 140 F/60 C. In contrast, a retrograded amylose starch (type-3
resistant starch (RS)) exhibits a melting endotherm peak at temperatures ranging
from about 302 F/150 C to 320 F/160 C. Another technique that can be used for
analyzing starch retrogradation is X-ray diffraction, whereby retrograded starch
exhibits a V-type structural pattern.
Both amylose and amylopectin can retrograde. However, the types of crystalline
structures formed are different (Fig. 6.15). In limited water, amylose can form an
A-type helix with 9 molecules of water trapped inside, whereas in excess water,
amylose and amylopectin can each form a B-type helix with 36 molecules of water
trapped inside (Biliaderis and Zawistowski, 1990; Wu and Sarko, 1978a,b).
The mechanism of starch retrogradation involves two steps (Fig. 6.16) (Slade and
Levine, 1987c). The first step is nucleation (seeding), with the nucleation rate
increasing with decreasing temperature down to Tg. The second step of retrogradation
is propagation (growth), with the rate of propagation increasing with increasing
temperature up to near Tm (w140 F/60 C).
The effects of temperature and moisture content on starch retrogradation in excess
water are illustrated in Fig. 6.17 (Perdon et al., 1999). Maximized retrogradation, as
manifested by a higher melting enthalpy for retrograded starch, was achieved by hold-
ing the system at a lower temperature, 4 C, for optimum nucleation, followed by hold-
ing at a higher temperature, 42 C, for optimum propagation (Slade and Levine,
1987c). At moisture contents higher than 60%, there was almost no retrogradation
observed.
Major changes in cereal biopolymers during ready-to-eat cereal processing 123

B type A type

Figure 6.15 Molecular arrangements of A- and B-type helical unit cells of crystalline amylose
(Wu and Sarko, 1978a,b).

Starch retrogradation occurs between Tg and Tm

Overall crystallization

Propagation
Nucleation

Rate

Tg Tm

Figure 6.16 Kinetics of starch retrogradationdnucleation rate faster near Tg while propagation
rate faster near Tm.

Dextrinization
Starch, when treated only with dry heat or roasting, can undergo a molecular degrada-
tion process called dextrinization. This transformation can also take place in the pres-
ence of an acidic or alkaline catalyst. The moisture level at which starch can dextrinize
is typically around 5%e7%. This process not only reduces the molecular size of the
starch but it can also alter starch’s molecular arrangement. Dextrinized starch is soluble
in cold water and exhibits reduced viscosity. It also manifests a reduced tendency for
gel formation and an increased reducing power. These properties can be used as a basis
for analyzing starch dextrinization.
Starch fragmentation can also occur during an extrusion cooking process, hence
decreasing the Tg of the extrudate and affecting the physical properties of the finished
124 Breakfast Cereals and How They Are Made

4.500
4.146
4 h at 4 °C, 4 h at 42 °C

3.690 4 h at 4 °C, 24 h at 42 °C
28 h at 4 °C
3.500
28 h at 42 °C
Enthalpy (H), J/g

2.500
1.935 1.941

1.500 1.331

1.010

0.654

0.500
0.000 0.000 0.000 0.000 0.000

–0.500
50 60 70
Moisture content, %

Figure 6.17 Effects of storage moisture content and temperature on the degree of starch
retrogradation, as measured by differential scanning calorimetry (Perdon et al., 1999).

breakfast cereal. Typically, the textural crispness is reduced, as is also the degree of
puffing (Kaletunc and Breslauer, 1993). Fig. 6.18 shows that the Tg curves for starch
hydrolysis products (SHPs) fall below the Tg curve for pregelatinized starch shown in
Fig. 6.11 and that Tg decreases with decreasing molecular weight (increasing dextrose
equivalent, DE) and size of the SHP, as first reported by Levine and Slade (1986).

200

150
Temperature (°C)

100
DE5
50

DE25
0

DE36
–50

–100
0.0 5.0 10.0 15.0 20.0 25.0 30.0
Moisture (%)
Figure 6.18 Tg curves for corn starch hydrolysis products of increasing dextrose equivalent
(DE) values.
Major changes in cereal biopolymers during ready-to-eat cereal processing 125

Cell wall polysaccharides: structure and transformation


Cereal grains contain various nonstarch polysaccharides, such as AX (0.2%e10.8%),
b-glucans (0%e7%), and cellulose (0.3%e3.9%) (Choct, 1997). Because cellulose is
insoluble and inert, no changes happen during breakfast cereal processing. Thus, only
AX and b-glucans will be discussed here.

Arabinoxylans
Among whole grain cereals, rye and wheat contain higher levels of AX at
7.1%e12.2% and 4%e9%, respectively, then followed by barley (3.4%e8%), oats
(2.2%e4.1%), and rice (2.6%), with sorghum having the lowest levels at <2%
(Collins et al., 2010; Fincher and Stone, 1986, 2004; Biliaderis and Izydorczyk, 2006).

Chemical properties
Molecular structure
Typically, cereal AX is a b-(1,4)-linked-D-xylose-based polymer substituted with
a-(1,2) and/or a-(1,3)-linked L-arabinose side chain. The potential number of L-arab-
inose substitution can vary from 0, 1, to 2 per xylose residue. Each hydroxyl group in
the main polymer chain is in equatorial configuration, which enhances the extensive
interpolymer chain hydrogen bonding during gelation.
The degree of substitution, as defined by the molar ratio of arabinose to xylose,
determines the solubility and aggregation properties of AX. The higher the degree
of substitution, the more soluble the AX will be. Therefore, AX which has been treated
with alpha-L-arabinofuranosidase to remove the arabinose side chain aggregates
readily into insoluble complexes (Andrewarths et al., 1979). The arabinose substitu-
tion pattern along the main chain affects the reactivity of AX. In many cereal grains,
the AX substitution patterns are random. The regions with relatively few arabinosyl
substituents are more susceptible to xylanases hydrolysis. In other regions, mono-
and disubstituted xylose residues are often clustered together and separated by one
to two unsubstituted xylose residue (Gruppen et al., 1993; Izydorczyk and Biliaderis,
1994). Some cereal AX also can have hydroxycinnamic acid substituents including
sinapinic acid, ferulic acid, and p-coumaric acid attached to the O-5 of arabinosyl
substituents (Smith and Harley, 1983). The feruloyl moiety is susceptible to oxidative
gelation through cross-linking with neighboring AX chains via peroxidase or
hydrogen peroxideecatalyzed free-radical mechanism (Stone and Morell, 2007).

Molecular weight
The molecular weight of AX depends on the cereal species, the cell wall type, the pro-
cedures for extraction, isolation, and purification, and the method for molecular weight
characterization. Fincher and Stone (2004) reported that the average molecular weight
126 Breakfast Cereals and How They Are Made

ranges from 65 to 5000 KDa with Mw/Mn ratios of 1.3e8.5 for alkali-soluble wheat
AX and rye AX, respectively.

Physical properties
Solubility
Typically, cereal AX is classified into water-soluble or water-extractable arabinoxy-
lans (WE-AX) and water-insoluble or water-unextractable arabinoxylans (WU-AX).
The WE-AX components are usually extracted with water at <40 C, whereas alkaline
solution is needed to extract WU-AX. The solubility of AX is determined by Mw, the
degree and patterns of arabinose substitution, and the various forms of covalent
bonding such as di-feruloyl bridges. WE-AX is only a small proportion of the total
AX for both wheat and rye. Wheat has 0.3%e0.9% WE-AX and 4%e9% total AX,
whereas rye has 0.6%e3.0% WE-AX and 7%e12% of total AX (Stone and Morell,
2007; Bengtsson and Aman, 1990).

Glass transitions
The state diagrams of wheat AX-H2O binary mixtures have been determined by Fessas
and Schiraldi (2001) using two preparations with arabinose/xylose ratios of 0.66 and
0.51, Mn of 56 and 410 KDa, Mw of 235 and 650 KDa, and dispersity values of 4.2
and 1.6, respectively. The Tg0 value, the glass transition temperature of the maximally
frozen concentrate, for high Mw AX was higher (17 C) than for low Mw AX
(35 C), while Cg0 , the solid content of the maximally frozen concentrate, was found
close to 75% w/w for both of them.

Changes due to processing


Extrusion cooking does not change much on the total amount of AX, while part of the
WU-AX is being solubilized, because the processing shear induces physical unentan-
glement and releases some from the entrapped cell wall matrix. The amount of
WE-AX increases by 2%, 14%, and 15% for barley, wheat, and rye, respectively
(Comino et al., 2016). The solubilized fraction of AX had similar chemical structures
to the starting flour, suggesting that increased solubilization was not from any specific
chemical fractions.

b-glucans
A common source of fiber in breakfast cereal is b-glucan. (1-3,1-4)-b-glucan is a major
component of cell wall material in cereal grains. Depending on the growing condition
and variety, the kernels of barley, oats, rye, and wheat has 3%e16%, 3%e7%,
Major changes in cereal biopolymers during ready-to-eat cereal processing 127

1%e2%, and 0.5%e1%, respectively. (Knuckles et al., 1992; Welch and Lloyd, 1989;
Wood et al., 1991a; Havrlentova and Kraic, 2006; Rudi et al., 2006; Lazardous et al.,
2007). Extrusion processing can increase the soluble b-glucan content at the expense
of the insoluble b-glucan, resulting in increased viscosity of barley-containing extru-
dates (Heryford, 1987; Singkhornart et al., 2013; Honců et al., 2016). In one study,
extrusion cooking increased the total dietary fiber and soluble dietary fiber contents
of barley flours, whereas the insoluble dietary fiber was decreased for some barley va-
rieties but increased for others (Yeung, 2001).
b-glucan in cereals typically exists in the subaleurone and endosperm cell walls
(Cui and Wang, 2009). In oat and barley, (1-3,1-4)-b-D-glucan is evenly distributed
throughout the starchy endosperm, whereas in wheat it concentrates more in the
subaleurone layer (Collins et al., 2010). The overall (1-3,1-4)-b-D-glucan content in
rye is much lower than in oats and barley, and it is evenly distributed throughout
the rye kernel (Autio, 2006).

Chemical properties
Molecular structure
Cereal b-glucan is a mix linked (1-3,1-4)-b-d-glucan composed of two major building
blocks: a trisaccharide (DP3) and a tetrasaccharide (DP4) unit. The combination of the
two units is over 90%, with the remaining being longer blocks of 5e20 (1/4)-linked
glucosyl residues (Woodward et al., 1983, 1988; Vårum and Smidsrød, 1988; Wood
et al, 1991b, 1994; Izydorczyk et al., 1998a,b; Papageorgiou et al., 2005). The molar
ratio of DP3/DP4 in these (1-3,1-4)-b-D-glucans are 1.8e2.4 for oats, 2.8e3.3 for
barley, and 3.0e4.5 for wheat (Cui and Wood, 2000; Lazaridou et al., 2004). Oat al-
eurone b-glucan has less of the DP4 units than endospermic b-glucan (Wood et al.,
1994). The ratio of (1-4)- to (1-3)-b-linkages for cereal grains ranges from 2.2 to
2.6 (Fincher and Stone, 1986, 2004).

Molecular weight
A very wide range of weight average molecular weight (Mw) has been reported for
cereal (1-3,1-4)-b-D-glucan at 21e3100 KDa (Fincher and Stone, 2004; Lazaridou
and Biliaderis, 2007; Andersson and B€orjesdotter, 2011). The Mw depends on variety,
the growing environment, the history of the source material, the isolation protocol used
(such as the method of extraction and purification, procedures for minimizing aggre-
gation and depolymerization), and the analytical method used for Mw determination
(Lazaridou and Biliaderis, 2007; Collins et al., 2010; Andersson and B€orjesdotter,
2011). When cereal (1-3,1-4)-b-D-glucan samples prepared in the same manner and
determined by the same methods, the Mw of oat and barley (1-3,1-4)-b-D-glucan is
in the range of 800e2000 KDa, while the Mw of wheat (1-3,1-4)-b-D-glucan is lower
in the range of 350e800 KDa (Rimsten et al., 2003; Li et al., 2006a,b; Lazaridou and
Biliaderis, 2007).
128 Breakfast Cereals and How They Are Made

Physical properties
The distribution of cellulosic oligomers, the linkage pattern, and Mw, along with
temperature and the hydrated concentration, determines the physical properties of b-glu-
cans, such as solubility and rheological property (Lazaridou and Biliaderis, 2007).

Solubility
The solubility of cereal (1-3,1-4)-b-D-glucans depends on the overall molecular
conformation, which in turn is determined by Mw and molecular structure. The irreg-
ularly spaced kinks due to (1-3)-b-linkages along the long molecular chain make
b-glucan much more soluble than cellulosic oligomers (Staudte et al., 1983).
The DP3/DP4 ratio also affects solubility, the (1-3,1-4)-b-D-glucans with ratios
close to 1 are more soluble than those with either very high or very low DP3/DP4
ratios. This is because as the proportion of either the DP3 or DP4 increases, the
polysaccharide would become more regular in shape and this would allow increased
level of molecular alignment, resulting in lower solubility (Lazairdou et al., 2004;
Fincher, 2009). Thus, a significant proportion of (1-3,1-4)-b-D-glucan from barley
(with DP3/DP4 ratio of 2.8e3.3) and oat grain (with the ratio of 1.8e2.4) is extractable
by water at 40 C, and the soluble fraction increases to 50%e70% when extracted at
65 C (Fincher, 1975; Fleming and Kawakami, 1977). On the other hand, the wheat
(1-3,1-4)-b-D-glucan with DP3/DP4 ratio of 3e4.5 is not as extractable even at 65 C
(Beresford and Stone, 1983; Li et al., 2006a).

Rheological properties
b-glucan molecules in aqueous solution behave as fluctuating random coils. The rheo-
logical property of the solution is a function of Mw and concentration. The solution of
b-glucan is Newtonian at low concentration (<0.2%), regardless of its Mw. Viscosity
is an exponential function of either concentration or molecular weight (Ren et al.,
2003). At higher concentration (>0.2%), the solution of low Mw b-glucan tends to
form soft gel, whereas the solution of high Mw b-glucan is pseudoplastic. The pseu-
doplasticity increases with Mw and concentration (Autio et al., 1987; Doublier and
Wood, 1995).
The gelling properties of low Mw cereal b-glucans are determined by DP3/DP4 ratio
and Mw. b-glucans with higher DP3/DP4 ratio, or with lower yet still above minimal
gelation Mw, gel faster and form stronger gels, because they favor the intermolecular
association of b-glucan chains (Li et al., 2006a; Vaikousi and Biliaderisa, 2004).

Changes during processing


The processing conditions, such as temperature, moisture, and shear, can significantly
affect the solubility and Mw of cereal b-glucans.

Solubility
Comino et al. (2016) reported that extrusion cooking does not change the total amount
of b-glucan, while part of the insoluble b-glucan being solubilized, because the
Major changes in cereal biopolymers during ready-to-eat cereal processing 129

processing shear induces physical unentanglement and releases some b-glucan from
the entrapped cell wall matrix. The amount of soluble b-glucan increases by 8%,
14%, and 22% for wheat, rye, and barley, respectively (Comino et al., 2016). The sol-
ubilized fraction has the same DP3/DP4 ratio as the original b-glucan, which means
solubilization is not from any specific fractions nor due to any chemical reaction
(Johansson et al., 2007; Robertson et al., 1997; Comino et al., 2016).

Molecular weight
The Mw of b-glucan in typical extruded flake remains high without much degradation
(Aman et al., 2004). However, under severe extrusion conditions, b-glucan in oats can
be depolymerized with Mw decreasing from 1930 KDa to 251 KDa, the cell wall
integrity is destroyed, and b-glucan is dispersed throughout the cereal. The extruded
cereals with lower Mw glucans are softer in texture, and lower in density
(Tosh et al., 2010).

Protein
Some heating processes, such as extrusion cooking, toasting, and puffing, may cause
protein denaturation, aggregation, and, in some cases, degradation. Denaturation
followed by aggregation leads to reduced water solubility (Li and Lee, 1996). Some
protein loss can be due to degradation and the Maillard reaction between protein
and reducing sugars, which decreases the nutritional value of the protein, due to the
loss of essential amino acids such as lysine. For example, extrusion-cooked rice at
15% moisture, 150 C, and 45 bars has 11% decreased lysine content (IRRI, 1984).
The degree of lysine loss would depend on the raw material type, its composition,
moisture content, and process conditions (Bjoerck et al., 1983; Ilo and Berghofer,
2003). The Maillard reaction is also important for color and flavor development.
Hydroxy-methyl furfural is one of the major intermediates of the Maillard reaction
(Garcia-Villanova and Guerra-Hernande, 1993), while formic acid and acetic acid
are among the Maillard degradation products (Rufian-Henares, 2006).
In conventionally processed corn flakes, the protein bodies were found to be flat-
tened and partially fused together, and the a-zeins were to some degree released,
whereas in extruded flakes, the protein bodies were completely disrupted and the
a-zeins were dispersed (Fig. 6.19). The degree of transformation increased as the
applied specific mechanical energy increased (Batterman-Azcona and Hamaker,
1998; Batterman-Azcona et al., 1999).

Changes during cereals processing


Management of moisture content and processing temperature serves as typical levers
during breakfast cereals production (Fig. 6.20) and changes the material behavior of
the grains. The state diagram is very useful in understanding material behavior during
130
Raw corn flour

Protein
body

Raw corn flour


Plasticorder PL2000, single screw

Boiled corn flour Standard corn Extruded corn

Breakfast Cereals and How They Are Made


flake flake

0.5 mm
35 kJ/kg SME 106 kJ/kg SME 165 kJ/kg SME 387 kJ/kg SME

0.5 mm

Figure 6.19 Changes in corn protein during cereals processing (Batterman-Azcona and Hamaker, 1998).
Major changes in cereal biopolymers during ready-to-eat cereal processing 131

Cook
Shred
Dry and Dry and
temper temper
Toast

Extrude
Mill

H 2O
H2O H2 O
H2 O H2 O
Temperature
Temperature Temperature Temperature Temperature

Figure 6.20 Managing moisture content and temperature in a typical breakfast cereals process.

processing (Caldwell et al., 2000). In the region below the glass transition curve, amor-
phous or partially crystalline rice starch would be a glassy solid and stable (Fig. 6.21).
In the region between the glass transition and crystalline melting curves, the material
would be rubbery and not stable. And, in the region above the melting curve, the ma-
terial would exhibit liquid flow behavior and would be very unstable.
A comparison of dynamic mapping between typical boiling rice for table use and
puff processing for cereals is shown in Fig. 6.22. The following sections discuss

Rice state diagram


250

200
Materila in this res area behaves as s
liquid and flows. Reactions are very,
very fast (arrhenian). Material is very
unstable
150
Temperature, °C

100
Melt, T
Material in this area behaves as a rubbery
solid and diffusion is fast. Reactions are
fast (WLF) and the material is not be stable
50

0 Material in this blue area behaves as a


glassy solid and diffusion is limited.
Reactions are very slow (arrhenian) and the
material is stable. Glass transition, T

–50
0 10 20 30 40 50 60 70
Moisture content, %

Figure 6.21 The effect of temperature and moisture content on the material properties of the
partially crystalline, partially amorphous rice starch (Perdon, 1999).
132 Breakfast Cereals and How They Are Made

Rice state diagram


250

200

Puffed rice
150 Cooked rice
Temperature, °C

Boiled rice
Pu
Cooling

ffin
g

100
Dr ying
Starch melt,T

Cooking

Cooking
50 Dried rice

Adding flavor, water Adding flavor, excess water


Raw rice
0 Cooled puffed
rice

Glass transition,T
–50
0 10 20 30 40 50 60 70
Moisture content, %

Figure 6.22 Dynamic process mapping of rice puffing compared with preparing boiled rice.

how process parameters used at each unit operation can be optimized to accomplish
critical grain material property changes.

Material handling
The stability and handling of a raw cereal ingredient are dependent on its temperature
and moisture content. If the surface of a material is above its Tg, due to either moisture
condensation or warm temperature, agglomeration, or clumping can be an issue during
storage or conveying. When the localized surface is above its critical moisture content
or its temperature is above its Tg, stickiness can occur. The “stickiness zone” is typi-
cally just above the endpoint of the glass transition region, i.e., around 10 to 30 C
above Tg (Levine and Slade, 1986).

Cooking
The primary purpose of cooking is to gelatinize or melt the partially crystalline starch
of a cereal grain into a continuous rubbery-state matrix for downstream processing,
such as flaking or shredding into a desired shape and size. In excess water, typically
above about 27%e35%, starch can gelatinize at temperatures below 100 C
(Slade and Levine,1989). At moisture contents less than about 27%e35%, starch
can melt at temperatures above 100 C. In breakfast cereal processing, cooking is
conducted under limited moisture content conditions, to facilitate subsequent drying
and, in some cases, to maintain the integrity of the cereal kernel. Cooking is sometimes
done under steam pressure, to melt the starch in the granules.
Flavor and color precursors, created by Maillard reactions, are also produced during
cooking. These aspects are important to finished product quality.
Major changes in cereal biopolymers during ready-to-eat cereal processing 133

Drying
Cooked grains need to be dried, to reduce stickiness before further processing.
Efficient drying conditions need to be above the glass transition temperature, where
moisture diffusion is faster, but below the “sticky point” temperature (around 10 to
30 C above Tg). However, drying at such high temperatures may result in a very steep
moisture content gradient, whereby the outer surface of a cooked grain has a lower
moisture content and thus a higher Tg, compared with its center (Fig. 6.23). Such
differences in material properties can result in case hardening, which could slow
down the drying process and cause premature puffing and/or stress cracking.

Tempering
After drying, an in-process sample can be held in suitable storage bins or containers
and allowed to “temper” at a certain temperature. Tempering allows for the equilibra-
tion of the temperature and/or the moisture content of the grains, to facilitate further
processing. This tempering enables both the surface and the center of the dried grain
to have comparable moisture content and thereby exist within the same region of the
state diagram, thus producing a more uniform matrix for further processing, such as
enhanced shreddability properties during shredded wheat production, or leading to
more uniform puffing.
During tempering, gelatinized starch can begin to retrograde, leading to an increase
in crystallinity. As a result, this retrograded starch would become less susceptible to
hydrolysis by amylolytic enzymes. This represents the basis of RS3 formation.

120
Tg line

Center in the rubbery state above Tg


100 Fast diffusion
Low modulus
High expansion coefficient
Temperature, °C

75

Surface in the glassy state below Tg


Slow diffusion
High modulus
50 Low expansion coefficient

25

0
5 10 15 20 25 30

Moisture content, % wb

Figure 6.23 Differences in the material properties across a nonuniform sample during drying.
134 Breakfast Cereals and How They Are Made

Milling and forming


Milling and forming are conducted at moisture contents and temperatures above the
glass transition of a material, for optimum processability. In contrast, below the glass
transition, excessive shear might be required, which could cause starch and/or protein
degradation. Such degradation would typically lead to the formation of lower molec-
ular weight compounds that could affect further processing and finished food quality.

Puffing and toasting


Puffing and toasting are done at high temperatures, for the purposes of expansion,
color, and flavor development, and drying to approximately 2%e4% moisture content.
During puffing, a food matrix would undergo volume changes, either by shrinkage due
to moisture loss or by expansion due to gas generation or pore formation. A foam
structure would be created by the rapid release of superheated steam and the resulting
expansion of the rubbery matrix. This expanded foam would be stabilized, as its mois-
ture content and temperature decrease below its Tg. Thus, puffing is a function of a
product’s moisture content, swelling pressure, and rheological properties that are
affected by its degree of starch gelatinization and degradation. According to the glass
transition theory of “food polymer science,” there would be negligible collapse in a
material, if it is processed below its Tg. But the larger the difference between a higher
processing temperature and a lower Tg, the greater would be the extent of collapse
(Levine and Slade, 1986). For an extrusion-expanded product, Fan et al. (1996)
showed that shrinkage or collapse stopped, when the processing temperature was
below the product’s Tg.
The optimal moisture content for puffing is typically around 10%e13%. Added
sugars in a matrix or excessive starch degradation during processing would have a
negative effect on the degree of puffing. The generation of lower molecular weight
compounds would lower the Tg of a pellet to be puffed, resulting in a reduction in
expansion or an increase in shrinkage during puffing.
Toasting is often carried out after puffing. The toasting step develops browning
flavors and color from the Maillard reaction precursors generated during cooking.

Conclusion
It takes an entire series of unit operations to transform dense, partially crystalline,
starch-containing cereal grains into ready-to-eat breakfast cereal pieces of a desired
density, geometry, color, flavor, texture, and nutritional profile, when consumed in a
bowl of milk. The advancement of the “food polymer science” concept since the early
1980s (Slade and Levine, 1991a) has served as a basis for understanding, in helping to
explain the major physical changes in a structural matrix during each unit operation,
and in helping to design new products with enhanced texture, stability, and processing
Major changes in cereal biopolymers during ready-to-eat cereal processing 135

efficiency. Most current processing techniques require a great deal of mechanical and
thermal energy input, along with the initial addition and subsequent removal of water.
In view of current water and energy sustainability issues, this “food polymer science”
approach can also offer abundant opportunities for future processing innovations.

References
Aman, P., Rimsten, L., Anderrson, R., 2004. Molecular weight distribution of b-glucan in
oat-based foods. Cereal Chem. 81, 356e360.
Andersson, A.A.M., B€orjesdotter, D., 2011. Effects of environment and variety on content and
molecular weight of b-glucan in oats. J. Cereal Sci. 54 (1), 122e128.
Andrewarths, K.A., Phillips, D.R., Stone, B.A., 1979. Solution properties of wheat flour
arabinoxylans and enzymatically modified arabinoxylans. Carbohydr. Res. 77, 191.
Atwell, W.A., Hood, L.F., Lineback, D.R., Varriano-Marston, E., Zobel, H.F., 1988. The
terminology and methodology associated with basic starch phenomena. Cereal Foods
World 33, 310e311, 306, 308.
Autio, K., 2006. Functional aspects of cereal cell-wall polysaccharides. In: Eliasson, A. (Ed.),
Carbohydrates in Food, second ed. CRC Food science and technology, New York.
Autio, K., Myllym€aki, O., M€alkki, Y., 1987. Flow properties of solutions of oat b-glucans.
J. Food Sci. 52, 1354e1366.
Batterman-Azcona, S.J., Hamaker, B.R., 1998. Changes occurring in protein body structure and
a-zein during cornflake processing. Cereal Chem. 75 (2), 217e221.
Batterman-Azcona, S.J., Lawton, J.W., Hamaker, B.R., 1999. Microstructural changes in zein
proteins during extrusion. Scanning 21 (3), 212e216.
Bengtsson, S., Aman, P., 1990. Isolation and chemical characterization of water-soluble ara-
binoxylans in rye grain. Carbohydr. Polym 12, 267e277.
Beresford, G., Stone, B.A., 1983. (1-3),(1-4),-b-D-glucan content of Triticum grains. J. Cereal
Sci. 1, 111e114.
Biliaderis, C.G., 1991. Non-equilibrium phase transitions of aqueous starch systems. In:
Levine, H., Slade, L. (Eds.), Water Relationships in Foods. Plenum Press, New York,
pp. pp251e274.
Biliaderis, C.G., 1992. Structures and phase transitions of starch in food systems. Food Technol.
46 (6), 98e109.
Biliaderis, C.G., Izydorczyk, M.S., 2006. Arabinoxylans: technologically and nutritionally
functional plant polysaccharides. In: Biliaderis, C.G., Izydorczyk, M.S. (Eds.), Functional
Food Carbohydrates. CRC Press, Boca Patton, FL, p. p251.
Biliaderis, C.G., Juliano, B.O., 1993. Thermal and mechanical properties of concentrated rice
starch gels of varying composition. Food Technol. 48 (3), 243e250.
Biliaderis, C.G., Zawistowski, J., 1990. Viscoelastic behavior of ageing starch gels: effects of
concentration, temperature and starch hydrolysates on network properties. Cereal Chem.
67, 240e246.
Bjoerck, I., Noguchi, A., Asp, N.G., Cheftel, J.C., 1983. Protein nutritional value of a biscuit
processed by extrusion cooking: effects on available lysine. J. Agric. Food Chem. 31 (3),
488e492.
136 Breakfast Cereals and How They Are Made

Caldwell, E.F., Fast, R.B., Ievolella, J., Lauhoff, C., Levine, H., Miller, R.C., Slade, L.,
Straham, B.S., Whalen, P.J., 2000. Unit operations and equipment I. Blending and cooking.
In: Fast, R., Caldwell, F. (Eds.), Breakfast Cereals and How They Are Made, second ed.
AACC, St. Paul, pp. p55e131.
Choct, M., June 1997. Feed non-starch polysaccharides: chemical structures and nutritional
significance. Feed Milling Int. (13e26).
Collins, H.M., Burton, R.A., Topping, D.L., Liao, M., Geoffrey, A.B., Fincher, B., 2010.
Variability in fine structures of non-cellulosic cell wall polysaccharides from cereal grains.
Cereal Chem. 87 (4), 272e282.
Comino, P., Collins, H., Lahnstein, J., Gidley, M.J., 2016. Effects of diverse food processing
conditions on the structure and solubility of wheat, barley and rye endosperm dietary fiber.
J. Food Eng. 169, 228e237.
Cui, S.W., Wang, Q., 2009. Cell wall polysaccharides in cereals: chemical structures and
functional properties. Struc. Chem. 20, 291e297.
Cui, S.W., Wood, P.J., 2000. Relationship between structures, molecular weights and rheo-
logical properties of cereal b-glucans. In: Kizhnari, K. (Ed.), Hydrocolloids (1) Physical
Chemistry and Industrial Applications of Gels, Polysaccharides, and Proteins. Elsevier,
Amsterdam, pp. 159e168.
Doublier, J.L., Wood, P.J., 1995. Rheological properties of aqueous solutions of (1–> 3)
(1–> 4)-bD-glucan from oats (Avena sativa L.). Cereal Chemistry 72, 335e340.
Fan, J., Mitchell, J.R., Blanshard, J.M.V., 1996. The effect of sugars on the extrusion of maize
grits: 1. The role of the glass transition in determining product density and shape. Int. J.
Food Sci. Technol. 31 (1), 55e65.
Fessas, D., Schiraldi, A., 2001. State diagrams of arabinoxylan-water binaries. Thermochim.
Acta 370 (1e2), 83e89.
Fincher, G.B., 1975. Morphology and chemical composition of barley endosperm cell walls.
J. Inst. Brew. 81 (2).
Fincher, G.B., 2009. Revolutionary times in our understanding of cell wall biosynthesis and
remodeling in the grasses. Plant Physiol. 149, 27e37.
Fincher, G.B., Stone, B.A., 1986. Cell walls and their components in cereal grain technology. In:
Pomeranz, Y. (Ed.), Advances in Cereal Science and Technology. AACC International, St
Paul, MN, pp. 207e295.
Fincher, G.B., Stone, B.A., 2004. Chemistry of nonstarch polysaccharides. In: Wigley, C.,
Corke, H., Walker, C.E. (Eds.), Encyclopedia of Grain Science. Elsevier, Oxford,
pp. 206e223.
Fleming, M., Kawakami, K., 1977. Studies of the fine structure of b-D-glucan of barley
extracted at different temperatures. Carbohydr. Res. 57, 15e23.
Garcia-Villanova, B., Guerra-Hernande, E., 1993. Liquid chromatography for the determination
of 5-(hydroxymethyl)-2-furaldehyde in breakfast cereals. J. Agric. Food Chem. 41 (8),
1254e1255.
Gruppen, H., Kormelink, F.J.M., Voragen, A.G.J., 1993. Water unextractable cell wall material
from wheat flour. 3. A structural model for arabinoxylans. J. Cereal Sci. 18, 111e128.
Havrlentova, M., Kraic, J., 2006. Content of b-D-glucan in cereal grains. J. Food. Res. Nutr. 45,
97e103.
Heryford, A.G., 1987. The Influence of Extrusion Processing on the Nutritional Value of Barley
for Weanling Pigs and Broiler Chickens. Montana State University, Bozeman. MS thesis.
Honců, I., Slukova, M., Vaculova, K., Sedlackova, I., Wiege, B., Fehling, E., 2016. The effect of
extrusion on the content and properties of dietary fiber components in various barley
cultivars. J. Cereal Sci. 68 (3), 132e139.
Major changes in cereal biopolymers during ready-to-eat cereal processing 137

Ilo, S., Berghofer, E., 2003. Kinetics of lysine and other amino acids loss during extrusion
cooking of maize grits. J. Food Sci. 68 (2), 496e502.
International Rice Research Institute, 1984. Annual Report for 1983. IRRI, Los Banos,
pp. p62e66.
Izydorczyk, M.S., Biliaderis, C.G., 1994. Studies on the structure of wheat endosperm arabi-
noxylans. Cabohydr. Polym 24, 61e71.
Izydorczyk, M.S., Biliaderis, C.G., Macri, L.J., MacGregor, A.W., 1998a. Fractionation of oat
(1-3), (1-4)-b-D-glucans and characterization of the fractions. J. Cereal Sci. 27, 321e325.
Izydorczyk, M.S., Macri, L.J., MacGregor, A.W., 1998b. Structure and physicochemical
properties of barley non-starch polysaccharidesdII. Alkali extractable b-glucans and
arabinoxylans. Carbohydr. Polym. 35, 259e269.
Johansson, L., Tuomainen, P., Anttila, H., Rita, H., Virkki, L., 2007. Effect of processing on the
extractability of oat b-glucan. Food Chem. 105 (4), 1439e1445.
Kaletunc, G., Breslauer, K.J., 1993. Glass transitions of extrudates: relationship with processing-
induced fragmentation and end-product attributes. Cereal Chem. 70 (5), 548e552.
Knuckles, B.E., Chiu, M.M., Betschart, A.A., 1992. b-glucan-enriched fractions from
laboratory-scale dry milling and sieving of barley and oats. Cereal Chem. 69 (2),
198e202.
Labuza, T.P., Tannenbaum, S.R., Karel, M., 1970. Water content and stability of low moisture
and intermediate moisture foods. Food Technol. 24, 543.
Labuza, T.P., 1976. Drying food: technology improves the sun. Food Technol. 30 (6), 37.
Labuza, T.P., 1980. The effect of water activity on reaction kinetics of food deterioration. Food
Technol. 34 (4), 36.
Lazaridou, A., Biliaderis, C.G., 2007. Molecular aspects of cereal b-glucan functionality:
physical properties, technological applications and physiological effects. J. Cereal Sci. 46
(2), 101e118.
Lazaridou, A., Biliaderis, C.G., Micha-Screttas, M., Steele, B.R., 2004. A comparative study on
structureefunction relations of mixed-linkage (1/3), (1/4) linear b-d-glucans. Food
Hydrocoll. 18 (5), 837e855.
Lasztity, R., 1995. The Chemistry of Cereal Proteins. CRC Press, New York.
Levine, H., Slade, L., 1986. A polymer physico-chemical approach to the study of commercial
starch hydrolysis products (SHPs). Carbohydr. Polym. 6 (3), 213e244.
Levine, H., Slade, L., 1988a. Collapse phenomena e a unifying concept for interpreting the
behavior of low-moisture foods. In: Blanshard, J.M.V., Mitchell, J.R. (Eds.), Food
Structure e Its Creation and Evaluation. Butterworths, London, pp. p149e180.
Levine, H., Slade, L., 1988b. Water as a plasticizer: physico-chemical aspects of low-moisture
polymeric systems. In: Franks, F. (Ed.), Water Science Reviews, vol. 3. Cambridge Uni-
versity Press, Cambridge, pp. p79e185.
Levine, H., Slade, L., 1989. Influences of the glassy and rubbery states on the thermal, me-
chanical, and structural properties of doughs and baked products. In: Faridi, H.,
Faubion, J.M. (Eds.), Dough Rheology and Baked Product Texture: Theory and Practice.
Van Nostrand Reinhold/AVI, New York, pp. p157e330.
Levine, H., Slade, L., 1990. Cryostabilization technology: thermoanalytical evaluation of food
ingredients and systems. In: Ma, C.Y., Harwalkar, V.R. (Eds.), Thermal Analysis of Foods.
Elsevier Applied Sci, London, pp. p221e305.
Levine, H., Slade, L., 1992. Glass transitions in foods. In: Schwartzberg, H.G., Hartel, R.W.
(Eds.), Physical Chemistry of Foods. Marcel Dekker, New York, pp. p83e221.
Li, M., Lee, T.C., 1996. Effect of extrusion temperature on solubility and molecular weight
distribution of wheat flour proteins. J. Agric. Food Chem. 44 (3), 763e768.
138 Breakfast Cereals and How They Are Made

Li, W., Wang, Q., Cui, S.W., Huang, X., Kauda, Y., 2006a. Elimination of aggregates of (1/ 3)
(1/ 4)-b-D-glucan in dilute solutions for light scattering and size exclusion chromatog-
raphy study. Food Hydrocolloids 20, 361e368.
Li, W., Cui, S.W., Kakuda, Y., 2006b. Extraction, fractionation, structural and physical char-
acterization of wheat b-D-glucans. Carbohydr. Polym. 63 (3), 408e416.
Papageorgiou, M., Lakhdara, N., Lazaridou, A., 2005. Water extractable (1/ 3, 1/ 4)-b-D-
glucans from barley and oats: an intervarietal study on their structural features and rheo-
logical behavior. J. Cereal Sci. 42, 213e224.
Perdon, A.A., May 1999. Amorphous State Transition in Rice during the Drying Process. 1999.
Ph.D. Dissertation. University of Arkansas.
Perdon, A.A., Siebenmorgen, T.J., Buescher, R.W., Gbur, E.S., 1999. Starch retrogradation and
texture of cooked milled rice during storage. J. Food Sci. 64 (5), 828e832.
Perdon, A.A., Siebenmorgen, T.J., Mauromoustakos, A., 2000. Glassy state transition and rice
drying: development of a brown rice state diagram. Cereal Chem. 71 (6), 708e713.
Ren, Y., Ellis, P.R., Ross-Murphy, S.B., Wang, Q., Wood, P.J., 2003. Dilute and semi-dilute
solution properties of (1-3),(1-4)-b-D-glucan, the endosperm cell wall polysaccharide of
oats (Avena sativa L.). Carbohydr. Polym. 53, 401e408.
Rimsten, L., Stenberg, T., Andersson, R., Andersson, A., Aman, P., 2003. Determination of
b-glucan molecular weight using SEC with calcofluor detection in cereal extracts. Cereal
Chem. 80 (4), 485e490.
Robertson, J.A., Majsak-Newman, G., Ring, S.G., Selvendran, R.R., 1997. Solubilization of
mixed linkage (1/3), (1/4)b-D-glucans from barley: effects of cooking and digestion.
J. Cereal Sci. 25 (3), 275e283.
Roos, Y., Karel, M., 1991; Dec. Applying state diagrams to food processing and development.
Food Technol. 66e71, 107.
Rudi, H., Uhlen, A.K., Harstad, O.M., Munck, L., 2006. Genetic variability in cereal carbo-
hydrate compositions and potentials for improving nutritional value. Anim. Feed Sci.
Technol. 130 (1e2), 55e65.
Rufian-Henares, J.A., 2006. Occurrence of acetic acid and formic acid in breakfast cereals.
J. Sci. Food Agric. 86 (9), 1321e1327.
Salwin, H., 1959. Defining minimum moisture contents for dehydrated foods. Food Technol.
13, 594.
Scott, W.J., 1957. Water relations of food spoilage microorganisms. In: Advances in Food
Research, Vol VII. Academic Press, New York, pp. p83e127.
Scott, W.J., 1958. The effect of residual water on the survival of dried bacteria during storage.
J. Gen. Microbiol. 19, 624e633.
Seow, C.C., Thevamalar, K., 1993. Internal plasticization of granular rice starch by hydrox-
ypropylation: effect on phase transitions associated with gelatinization. St€arke 45 (3),
85e88.
Singkhornart, S., Gu, B.J., Ryu, G.H., 2013. Physicochemical properties of extruded germinated
wheat and barley as modified by CO2 injection and different extrusion conditions. Int. J.
Food Sci. Technol. 48 (2), 290e299.
Slade, L., Levine, H., 1987a. Structural stability of intermediate moisture foods e a new un-
derstanding? In: Blanshard, J.M.V., Mitchell, J.R. (Eds.), Food Structure e Its Creation and
Evaluation. Butterworths, London, pp. p115e147.
Slade, L., Levine, H., 1987b. Polymer-chemical properties of gelatin in foods. In: Pearson, A.M.,
Dutson, T.R., Bailey, A. (Eds.), Advances in Meat Research, Collagen as a Food, vol. 4.
AVI Publ, West Port, pp. p251e266.
Major changes in cereal biopolymers during ready-to-eat cereal processing 139

Slade, L., Levine, H., 1987c. Recent advances in starch retrogradation. In: Stivala, S.S.,
Crescenzi, V., Dea, I.C.M. (Eds.), Industrial Polysaccharides e the Impact of Biotech-
nology and Advanced Methodologies. Gordon and Breach Sci Publ, New York,
pp. p387e430.
Slade, L., Levine, H., 1989. A food polymer science approach to selected aspects of starch
gelatinization and retrogradation. In: Millane, R.P., BeMiller, J.N., Chandrasekaran, R.
(Eds.), Frontiers in Carbohydrate Research-1: Food Applications. Elsevier Applied Sci,
London, pp. p215e270.
Slade, L., Levine, H., 1991a. Beyond water activity: recent advances based on an alternative
approach to the assessment of food quality and safety. Crit. Rev. Food Sci. Nutr. 30 (2e3),
115e360.
Slade, L., Levine, H., 1991b. A polymer science approach to structure-property relationships in
aqueous food systems: non-equilibrium behavior of carbohydrate-water systems. In:
Levine, H., Slade, L. (Eds.), Water Relationships in Foods. Plenum Press, New York,
pp. p29e101.
Slade, L., Levine, H., 1995. Polymer science approach to water relationships in foods. In:
Barbosa-Canovas, G.V., Welti-Chanes, J. (Eds.), Food Preservation by Moisture Control.
Technomic, Lancaster, pp. p33e132.
Smith, M.M., Harley, R.D., 1983. Occurrenace and nature of ferulic acid sunstitution of cell wall
polysaccharides in graminaceous plants. Carbohydr. Res. 118, 65e80.
Sperling, L.H., 2005. Introduction to Physical Polymer Science, fourth ed. John Wiley & Sons,
Inc, Hoboken.
Staudte, R.G., Woodward, J.R., Fincher, G.B., Stone, B.A., 1983. Water soluble (1-3),(1-4)-
b-D-glucan from barley endosperm. III. Distribution of cellotriosyl and cellotetraosyl
residues. Carbohydr. Polym. 3, 299e312.
Stone, B.A., Morell, M., 2007. Carbohydrates. In: Khan, K., Shewry, P. (Eds.), Wheat Chem-
istry and Technology. AACC International, St Paul, MN, pp. P299eP362.
Tester, R.F., Morrison, W.R., 1990. Swelling and gelatinization of cereal starches I. Effects of
amylopectin, amylose and lipids. Cereal Chem. 67 (6), 551e559.
Tester, R.F., Debon, S.J.J., 2000. Annealing of starch e a review. Int. J. Biol. Macromol. 27 (1),
1e12.
Tosh, S.M., Brummer, Y., Miller, S.S., Regand, A., 2010. Processing affects the physico-
chemical properties of b-glucan in oat bran cereal. J. Agric. Food Chem. 58, 7723e7730.
Vaikousi, H., Biliaderisa, C.G., 2004. M.S.Izydorczyk MS. Solution flow behavior and gelling
properties of water-soluble barley (1/3,1/4)-b-glucans varying in molecular size.
J. Cereal Sci. 39 (1), 119e137.
Vårum, K.M., Smidsrød, O., 1988. Partial chemical and physical characterization of
(1 / 3),(1 / 4)-b-d-glucans from oat (Avena sativa L.) aleurone. Carbohydr. Polym. 9
(2), 103e117.
Welch, R.W., Lloyd, J.D., 1989. Kernel (1-3,1-4)-b-D-glucan content of oat genotypes. J. Cereal
Sci. 9, 35e40.
White, G.W., Cakebread, S.H., 1966. The glassy state in certain sugar-containing food products.
Int. J. Food Sci. Technol. 1 (1), 73e82.
Wood, P.J., Weisz, J., Blackwell, B.A., 1991a. Molecular characterization of cereal b -D-
Glucans. Structural analysis of oat b -D-Glucan and rapid structural evaluation of b -D-
Glucans from different sources by high-performance liquid chromatography of oligosac-
charides released by lichenase. Cereal Chem. 68 (1), 31e38.
140 Breakfast Cereals and How They Are Made

Wood, P.J., Weisz, J., Fedec, P., 1991b. Potential for b-glucan enrichment in brans derived from
oat (Avena sativa L.) cultivars of different (1-3,1-4)-b-D-glucan concentrations. Cereal
Chem. 68, 48e51.
Wood, P.J., Weisz, J., Blackwell, B.A., 1994. Structural studies of (1->3), (1->4)-b-D-glucans
by C13-nuclear magnetic resonance spectroscopy and by rapid analysis of cellulose-like
regions using high-performance anion-exchange chromatography of oligosaccharides
released by lichenase. Cereal Chem. 71 (3), 301e307.
Woodward, J.R., Fincher, G.B., Stone, B.A., 1983. Water-soluble (1/3), (1/4)-b-D-glucans
from barley (Hordeum vulgare) endosperm. II. Fine structure. Carbohydr. Polym. 3 (3),
207e225.
Woodward, J.R., Phillips, D.R., Fincher, G.B., 1988. Water-soluble (1 / 3,1 / 4)-b-d-glucans
from barley (Hordeum vulgare) endosperm. IV. Comparison of 40 C and 65 C soluble
fractions. Carbohydr. Polym. 8 (2), 85e97.
Wu, H.C.H., Sarko, A., 1978a. The double helical molecular structure of crystalline A-amylose.
Carbohydr. Res. 61 (1), 27e40.
Wu, H.C.H., Sarko, A., 1978b. The double helical molecular structure of crystalline B-amylose.
Carbohydr. Res. 61 (1), 7e25.
Yeung, J.S., 2001. Extrusion Cooking of Barley Flour with and without Thermostable
A-Amylase. University of Alberta, Edmonton. MS thesis.
Zeleznak, K.J., Hoseney, R.C., 1987. The glass transition in starch. Cereal Chem. 64 (2),
121e124.

Further reading
Shafiur Rahman, M., 1995. Food Properties Handbook. CRC Press, Boca Raton, Florida, USA.
Slade, L., Levine, H., 1995. Glass transitions and water-food structure interactions. In:
Taylor, S.L., Kinsella, J.E. (Eds.), Advances in Food and Nutrition Research, vol. 38.
Academic Press, San Diego, pp. p103e269.

You might also like