SlideShare a Scribd company logo
1 of 16
Download to read offline
Starch/Stärke 2014, 66, 1–16 DOI 10.1002/star.201300238 1 
REVIEW 
Production, structure, physicochemical and functional properties 
of maize, cassava, wheat, potato and rice starches 
Jasmien Waterschoot, Sara V. Gomand, Ellen Fierens and Jan A. Delcour 
Laboratory of Food Chemistry and Biochemistry, Leuven Food Science and Nutrition Research Centre (LFoRCe), KU Leuven, 
Leuven, Belgium 
In 2012, the world production of starch was 75 million tons. Maize, cassava, wheat and potato are 
the main botanical origins for starch production with only minor quantities of rice and other 
starches being produced. These starches are either used by industry as such or following some 
conversion. When selecting and developing starches for specific purposes, it is important to 
consider the differences between starches of varying botanical origin. Here, an overview is given 
of the production, structure, composition, morphology, swelling, gelatinisation, pasting and 
retrogradation, paste firmness and clarity and freeze–thaw stability of maize, cassava, wheat, 
potato and rice starches. Differences in properties are largely defined by differences in amylose 
and amylopectin structures and contents, granular organisation, presence of lipids, proteins and 
minerals and starch granule size. 
Received: October 3, 2013 
Revised: January 21, 2014 
Accepted: January 23, 2014 
Keywords: 
Gelatinisation / Production / Retrogradation / Starch / Structure 
1 Introduction 
Starch is an important source of carbohydrates in the human 
diet. In addition, it is a versatile and widely used additive in 
the food, paper, chemical and pharmaceutical industries. 
Worldwide, 75 million tons of starch were produced in 2012 
(http://www.zuckerforschung.at/) and marketed as native, 
physically or chemically modified starch but also as liquid 
and solid sweeteners. This paper gives an overview of the 
production, chemical composition, structure and functional 
properties of maize, cassava, wheat, potato and rice starches. 
2 Starch production and uses 
Of the above mentioned world starch production more than 
half was produced in the United States. In the European 
Union, 10 million tons were produced (http://www.zuck-erforschung. 
at/).World production is estimated to increase to 
about 85 million tons by 2015. The most important botanical 
origins for producing starches are maize, cassava, wheat and 
potato, respectively. Table 1 shows the estimated 2015 
production for each starch. Almost 80% of the starch 
production is from maize. In the USA, mainly maize starch is 
produced, although (very) small amounts of wheat, potato and 
rice starches are also manufactured [1]. In Europe, in addition 
to maize (47%) and wheat starch (39%), also potato starch 
(14%) and a very small amount of rice starch (<0.5%) are 
produced [2] (http://www.aaf-eu.org/). Cassava starch is 
mainly produced in Southeast Asia and Brazil [3]. Only a 
small fraction (7% for maize, 4% for cassava, 0.9% for wheat 
and potato and 0.007% for rice) of the raw material crops 
are used for starch production. 
In applications, starch is mainly used as starch derived 
sweeteners and as native and modified starches. In 2011, 
in the European Union, 57% of the produced starch was 
converted to sweeteners, 23% was used as native starch 
and 20% was modified (http://www.aaf-eu.org/). Important 
starch derived sweeteners are glucose (syrups), (high) 
fructose (syrups), and the polyols mannitol, sorbitol and 
maltitol. Maltodextrins and oligosaccharide syrups are also 
produced [1, 4]. Native starch is used because of its thickening 
and gelling capacities. However, for a number of applications, 
properties of native starches fail to meet process or product 
requirements. This is why starches are also chemically or 
Correspondence: Jasmien Waterschoot, Laboratory of Food 
Chemistry and Biochemistry, Leuven Food Science and Nutrition 
Research Centre (LFoRCe), KU Leuven, Kasteelpark Arenberg 20, 
B-3001 Leuven, Belgium 
E-mail: jasmien.waterschoot@biw.kuleuven.be 
Fax: þ32-16-32-19-97 
 2014 WILEY-VCH Verlag GmbH  Co. KGaA, Weinheim www.starch-journal.com
2 J. Waterschoot et al. Starch/Stärke 2014, 66, 1–16 
physically modified. Cross-linking and substitution are 
common modifications for starches used in food production. 
Cross-linking of starch improves its acid, heat and shear 
stability, while the introduction of bulky substituents on the 
starch chains reduces retrogradation [5]. The EU exports 
approximately 10% of its produced starch and starch 
derivatives. Of the remaining 90%, 62% is used by the food 
industry, 1%is used for feed and 37% is used by the non-food 
industry (http://www.aaf-eu.org/). Major non-food starch 
applications are in the paper and board, pharmaceutical (e.g. 
tablet formulations, encapsulating agents), cosmetics, chem-ical 
(e.g. adhesives, starch-based plastics) and textile indus-tries. 
Modifications to meet requirements for non-food 
applications include oxidation, cationisation, copolymerisa-tion, 
hydrolysis and substitution [6]. 
3 Starch production processes 
Starch production processes and their associated costs 
depend on the botanical origin of the starch. Isolation of 
starch from cassava and potato tubers is relatively simple 
due to their tissue structure and their relatively low protein 
and fat contents [7, 8]. Isolation of cereal starches is more 
difficult as higher levels of these components need to be 
removed [2, 9, 10]. 
3.1 Maize starch 
Maize starch is commercially isolated by a wet milling 
procedure. First, contaminating material is removed from the 
bulk mass, after which the maize is steeped in water 
containing a low level of sulphur dioxide for typically 24–40 h 
at 48–52°C to soften the kernels and to obtain optimal milling 
and separation of the maize components during the wet 
milling phase. During steeping, the kernels absorb water and 
sulphur dioxide [10]. The latter induces protein swelling and 
dispersion because it cleaves inter- and intramolecular 
disulphide bonds and thus reduces the average MW and 
increases the solubility of the proteins [11]. During steeping, 
lactic acid bacteria develop and produce lactic acid from the 
available sugars. This causes a drop in pH to 4–5 which is 
optimal for separation of the maize protein from starch. In 
addition, lactic acid bacteria hydrolyse some highMWsoluble 
protein [10]. An important side effect of the steeping process 
is annealing of the starch granules, which changes their 
structural properties and increases the gelatinisation temper-ature 
[12]. After steeping, and wet milling, the germ is 
removed and starch and protein are separated from non-starch 
polysaccharides by sieving. After dietary fibre removal, 
starch is separated from protein by centrifugation or 
sedimentation and flash dried [10]. 
3.2 Cassava starch 
Cassava roots are cleaned, peeled and pulverised into a pulpy 
slurry. Starch is then isolated at ambient temperature. 
The roots contain very small levels of protein (1%) and 
impurities, which can all be removed by decantation. Non-starch 
polysaccharides are removed by passing the slurry 
through extractors with coarse and fine screens to remove 
both large and smaller molecules. The slurry is then 
dewatered by centrifugation and flash dried. A small level 
of sulphur dioxide can be added to the process water to 
control bacterial growth and facilitate the process [13]. 
3.3 Wheat starch 
Wheat is dry milled to separate bran and germ from the 
endosperm which is recovered as flour. Different processes 
are used to separate starch and gluten proteins from wheat 
flour, i.e. dough-ball, batter, dough-batter and high pressure 
disintegration processes [9]. Van Der Borght et al. [14] 
extensively reviewed the main processes. In these processes, 
flour is mixed with different amounts of water to induce 
gluten agglomeration or even gluten network formation. The 
dough-ball and dough-batter process are carried out at 
ambient temperature, while for the other processes warm 
water (30–50°C) is usually used [9, 14, 15]. After formation 
of batter or dough, starch and gluten can be separated based 
on their difference in density (by centrifugation, in hydro-cyclones) 
or particle size (by sieving) [9, 14, 15]. The starch 
is then further purified with the use of hydrocyclones or 
separators and decanters and dried [9]. 
Table 1. Production of starches of different botanical origins 
Maize starch Cassava starch Wheat starch Potato starch Rice starch 
Estimated world starch production 2015 
(million tons/year) 
64.6 10.2 6.0 3.4 0.05 
World production raw material 2011 
(million tons/year) (www.faostat.fao.org) 
880 250 704 374 723 
Main production countries USA, Japan, China, 
South Korea [3] 
Thailand, Indonesia, 
Brazil, China [3] 
France, Germany, 
USA, China [3] 
Netherlands, Germany, 
France, China [2] 
Belgium [2], 
Thailand, Italy 
 2014 WILEY-VCH Verlag GmbH  Co. KGaA, Weinheim www.starch-journal.com
Starch/Stärke 2014, 66, 1–16 3 
3.4 Potato starch 
Potatoes are ground to obtain a mixture of starch granules, 
broken cell walls and the ‘potato juice’, a solution containing 
proteins, amino acids, sugars and salts. Starch granules and 
non-starch polysaccharides are separated from the juice by 
centrifugation. Starch and large non-starch polysaccharides 
are then separated by sieving. However, some smaller non-starch 
polysaccharides and some proteins remain present in 
the starch fraction. The remaining non-starch polysacchar-ides 
can be removed by centrifugation based on the density 
difference between these polysaccharides and starch. The 
soluble protein is removed in a multi-stage, countercurrent 
flow system. After this, water is removed with rotating 
vacuum drum filters and flash drying [3]. 
3.5 Rice starch 
Rice starch is traditionally isolated by an alkaline procedure. 
Broken rice, a by-product of the conversion of brown rice into 
white rice in a process referred to as milling, is steeped in a 
0.3–0.5% sodium hydroxide solution for 12–24 h at 20–50°C. 
As mentioned above, annealing of the starch granules may 
occur at the higher process temperatures. In rice, protein and 
starch are strongly associated. Sodium hydroxide solubilises 
the rice protein and facilitates the isolation of starch during 
subsequent wet milling of the kernels. After wet milling, 
starch is kept in suspension to allow further solubilisation of 
proteins. Non-starch polysaccharides are removed by filtra-tion 
and the slurry is washed to remove the proteins, 
neutralised and dried [2, 16]. 
4 Chemical composition of starch granules 
Starch mainly consists of two polymers of a-D-glucose units 
linked by a-1,4 and a-1,6 bonds. These are the nearly linear 
amylose (AM) and highly branched amylopectin (AP). In 
addition, starch contains minor constituents (lipids, proteins 
and minerals) of which the levels vary with the botanical 
origin (Table 2). Tuber and root starches [e.g. potato (0.1%) 
and cassava starch (0.2%)] usually contain less lipid than 
cereal starches (0.6–1.4%) [17, 18]. Lipid content is positively 
correlated with AM content, i.e. most AM free starches 
Figure 1. Structure of glucose-6-phosphate in an a-(1,4) bound 
glucose chain. 
contain negligible lipid levels [18]. Wheat starch contains a 
high level of LPLs and some glycolipids, whereas lipids of 
maize and rice starches consist of FFA and LPLs [18]. The 
endogenous lipids reduce swelling and leaching of carbo-hydrates 
during heating of starch in excess water by the 
formation of AM lipid complexes [19–21]. In general, protein 
(0.1–0.5%) and ash (0.1–0.3%) contents of starch are very 
low [22–26]. Potato starch contains a relatively high level of 
phosphorus (0.09%) in the form of phosphate monoesters 
that are primarily covalently bound to AP [27]. The phosphate 
is mainly ester linked at the C-6 (61%) (see Fig. 1) and C-3 
(38%) positions, with only 1% linked at the C-2 position [28]. 
The presence of phosphate monoesters in potato starch has 
large consequences for its swelling behaviour. Negatively 
charged phosphate groups cause repulsion between adjacent 
AP chains and allow rapid hydration and large swelling of the 
granule [27]. Phosphorus in cereal starches (0.01–0.07%) is 
mainly present in the form of phospholipids [26, 29]. 
5 Amylose and amylopectin 
The AM content of normal starches varies between 14 and 
29% [30–36]. Table 3 shows AM contents of potato, cassava, 
wheat, maize and rice starches. The AM content of rice 
starches varies from 0 to 40% [30, 37]. Variations in AM 
content can be produced through cross-breeding, mutagene-sis 
or transgenic breeding [38]. AM free starches are called 
‘waxy’ and exist for maize, cassava, wheat, potato and rice 
starches [30, 31, 39–45]. Starches with a high AM content 
(30%) are also available. For maize, starches with an 
AM content ranging from 50 to 90% are commercially 
Table 2. Composition of starches of different botanical origins 
Maize starch Cassava starch Wheat starch Potato starch Rice starch 
Lipids (%) 0.6–0.8 [18, 26] 0.2 [17] 0.8–1.2 [18] 0.1 [17, 26] 0.6–1.4 [18] 
Proteins (%) 0.4 [22, 26] 0.3 [23] 0.2–0.3 [22, 24] 0.1 [22, 26] 0.1–0.5 [25] 
Ash (%) 0.1 [22, 26] 0.3 [23] 0.2 [22] 0.3 [22], 0.2 [26] 0.1 [25] 
Phosphorus (%) 0.02 [29], 0.01 [26] 0.01 [29] 0.05 [29] 0.09 [29], 0.06 [26] 0.07 [29] 
Lipid, protein, ash and phosphorus contents are shown as % of total dry weight. 
 2014 WILEY-VCH Verlag GmbH  Co. KGaA, Weinheim www.starch-journal.com
4 J. Waterschoot et al. Starch/Stärke 2014, 66, 1–16 
Table 3. Contents and structural properties of amylose (AM) and amylopectin (AP) and degree of crystallinity of starches of different 
available [44, 46]. Also for potato starch an AM content from 
56 to 92% was obtained [31, 47–49], while the highest AM 
content obtained so far for wheat (74%) [50] and rice 
(56%) [51] is lower. Mutants with only slightly higher AM 
contents than those of their regular counterparts have been 
developed for wheat (31–56%) [41, 47, 52–54], cassava (28– 
36%) [39, 55] and potato (27–37%) starches [31]. 
In essence, AM is a linear polymer. It contains hardly any 
a-1,6 branch points (1%) [56]. AP is highly branched, with 
5–6% a-1,6 linkages [18]. A starch characteristic is its DP, i.e. 
the number of glucose units in the polysaccharide. Average 
degrees of polymerisation can be calculated based on the 
number or on the weight of the molecules. The number 
average DP (DPn) and the weight average DP (DPw) are given 
by formulas (1) and (2): 
DPn ¼ 
P1 
1 DPiðmi=MiÞ 
P1 
1 
ðmi=MiÞ 
ð1Þ 
DPw ¼ 
P1 
P1 DPimi 1 
1 mi 
ð2Þ 
withmi the mass concentration andMi theMWof chains with 
a DP¼i [57]. DPw is always higher than DPn, except for a 
monodisperse polymer, in which case DPn equals DPw [58]. 
Table 3 lists DPn of AM and AP of the different starches. DPn 
ofAMvaries from 570 to 8025 [59–65], while DPn of AP varies 
from 4700 to 18 000 [62, 63, 66–68]. DPn values of potato and 
botanical origins 
Maize starch Cassava starch Wheat starch Potato starch Rice starch 
AM content (%) 23 [34], 
28 [32] 
18–24 [33], 
18 [31, 34] 
25 [35], 
26 [32, 34] 
19–22 [31], 
17 [34], 
23 [32], 
25–27 [35] 
17–21 [32], 
21 [34], 
14–29 [30], 
16–19 [36], 
4–16 [37] 
DPn 
a) of AM 960 [61], 
830 [63] 
2660 [59], 
3642 [60] 
570 [59] 
3827 [60], 
1290 [61], 
830–1570 [62], 
1200–1500 [67] 
4920 [59], 
8025 [60] 
920–1110 [64], 
847–1118 [65] 
DPn of AP 5100 [63], 
15 900 [66] 
– 5000–9400 [62], 
13 000–18 000 [67] 
11 200 [66] 4700–12 800 [68], 
8200–10 900 [66] 
Number of AM 
molecules per g 
starch 1017 
9–10 2–3 3–11 1–2 6–9 
Number of AP 
molecules per g 
starch 1017 
2–6 – 2–6 3 2–4 
CLb) of AM 335 [61], 
340 [63] 
340 [59] 250–320 [67], 
135–255 [62], 
270 [61], 
300 [59] 
670 [59] 230–370 [64] 
CL of AP 28 [70], 
24 [34], 
20–21 [63], 
20 [72, 75] 
26 [70], 
28 [34], 
18–19 [73], 
19 [72] 
25 [70], 
23 [34], 
19–20 [67], 
19–21 [62], 
19 [72], 
23 [69] 
34 [70], 
29 [34], 
31 [71], 
23 [72] 
25–28 [70], 
23 [34], 
17–18 [73], 
19–22 [68], 
18–19 [72] 
Average number of 
chains per molecule 
of AM 
2.9 [61], 
2.4 [63] 
7.8 [59] 4.4–5.2 [67], 
5.5–6.5 [62], 
4.8 [61], 
1.9 [59] 
7.3 [59] 2.5–4.3 [64] 
Average number of chains 
per molecule of AP 
240 [63] – 660–920 
(based on [67]) 
500 (based on 
[72] and [66]) 
220–700 [74] 
Degree of crystallinity 
as determined with 
X-ray diffraction (%) 
27 [86], 
40 [85] 
24 [86], 
38 [85] 
20 [86], 
36 [85] 
24 [86], 
28 [85] 
38 [85] 
a) DPn, number degree of polymerisation 
b) CL, average chain length 
 2014 WILEY-VCH Verlag GmbH  Co. KGaA, Weinheim www.starch-journal.com
Starch/Stärke 2014, 66, 1–16 5 
cassava AM are higher than those of maize, wheat and rice 
AM. Based on DPn and the contents of AM and AP, the 
number of AM and AP molecules per mass unit of starch can 
be calculated. It can be estimated from data in Table 3 that a 
given mass of regular starch contains more AM molecules 
than AP molecules (except for potato starch), which is due to 
their lower DPn. 
The average chain length (CL) of AM varies from 135 to 
670 glucose units per chain. The CL of potato AM is 670 
glucose units while that of the other starches ranges from 135 
to 370 glucose units [59, 61–64, 67]. Because of the highly 
branched nature of AP, its CL (17–34 glucose units) is much 
smaller than that of AM. Potato and cassava AP have a slightly 
higher CL than AP of the cereal starches [34, 62, 63, 67–75]. 
AP of high AM starches has a similar or higher CL than AP of 
the regular counterparts. In addition, these starches contain 
some material with DPs and branching patterns intermediate 
between those of AM and AP [46, 47, 76–79]. The average 
number of chains per molecule can be calculated from DPn 
and CL and is much smaller for AM (1.9–7.8) than for AP 
(220–920) [59, 61–64, 66, 67, 72, 74]. 
With regard to determination of starch molecular size, it is 
important to note that the analytical results depend on the 
procedure used. Molecular size determinations require 
complete dissolution of the starch polymers, with removal 
of non-starch components that interfere with the analysis and 
without starch loss or degradation during the procedure. 
These requirements are challenging, so when interpreting 
results, caution should be taken when comparing results 
obtained by different procedures [80, 81]. 
6 Different levels of starch granule 
organisation 
Plants synthesise starch in a granular form. Starting from the 
hilum, starch is deposited in alternating amorphous and semi-crystalline 
concentric growth rings [16]. The semi-crystalline 
growth rings consist of amorphous and crystalline lamellae. 
AP is largely responsible for the crystalline character of starch. 
Side chains of AP form double helices which are ordered in 
clusters. The crystalline lamellae contain the double helices, 
while the amorphous lamellae contain the AP branch points, 
which connect the double helices [82]. The amorphous growth 
rings consist of AMand less ordered AP [16]. AM and AP are 
not present in separate regions, but highly intermingled in 
the granule [83, 84]. The degree of crystallinity is usually 
determined with X-ray diffraction [18]. It varies from 20 to 
40% depending on the botanical origin (Table 3) [85, 86]. 
The packing of AP double helices can give rise to different 
crystal structures or polymorphic forms. Cereal starch 
crystals are generally packed according to the A-type packing. 
Such packing is more dense than the B-type packing of e.g. 
potato starch and high AM maize starches. Cassava starch 
contains A-type or C-type crystals (a mixture of A- and B-type 
crystals) [85, 87, 88]. In A-type starches, crystals are packed in 
a monoclinic unit cell (a¼2.124 nm, b¼1.172 nm, c¼1.069 
nm and g¼123.5°) with eight water molecules, whereas in 
B-type starches, crystals are packed in a hexagonal unit cell 
(a¼b¼1.85nm and c¼1.04 nm) with 36 water molecules 
(see Fig. 2). Besides A- and B-type crystals, a third polymorph 
exists, i.e. V-type crystals. In this polymorph, AM single 
helices form inclusion complexes with e.g. iodine, alcohols 
or fatty acids [18]. 
7 Morphology of the starch granule 
The size and shape of starch granules (Table 4) depend on 
the botanical origin and vary widely. Figure 3 shows the 
granular morphology of potato, cassava, wheat, maize and 
rice starches. Potato starch has very large, round or oval 
granules (10–100mm), while rice starch has very small, 
polygonal granules (3–8mm) [32, 89, 90]. Cassava starch has 
round or truncated granules while maize starch granules are 
polygonal. Both starches have granules with somewhat 
similar dimensions (5–20mm for maize starch and 3–32 mm 
for cassava starch) [31–33, 89]. While the shape and size of 
waxy maize starch granules resemble those of regular maize 
starch granules, high AM maize starches contain, in addition 
to the normal polygonal granules, a number of filamentous 
elongated granules [89, 91]. Wheat starch has a bimodal size 
distribution, with small, round B granules (2–10mm) and 
large, lenticular (20–32mm) A granules [32, 89, 92, 93]. 
Figure 2. Monoclinic unit cell of A-type crystals and hexagonal unit 
cell of B-type crystals. Projection of the structure is in the ab plane. 
Reprinted from International Journal of Biological Macromolecules, 
23, Buléon, A., Colonna, P., Planchot, V. and Ball, S., Starch 
granules: structure and biosynthesis, 85–112, Copyright (1998), 
with permission from Elsevier. 
 2014 WILEY-VCH Verlag GmbH  Co. KGaA, Weinheim www.starch-journal.com
6 J. Waterschoot et al. Starch/Stärke 2014, 66, 1–16 
Table 4. Morphological properties of granules of starches of different botanical origins 
Maize starch Cassava starch Wheat starch Potato starch Rice starch 
Shape Round, polygonal [32, 89] Round, truncated [31] Round, lenticular [32, 89] Round, oval [89] Polygonal [32, 89] 
Diameter, range (mm) 5–20 [89] 3–32 [33] A: 20–35 B: 2–10 [32] 10–110 [31] 3–8 [32, 89] 
Volume mean diameter (mm) 15 [32] 17–18 [31] A: 21–23 B: 6–7 [92, 93] 48–60 [31] 6 [90] 
Number mean diameter (mm) (own data) 13 11 A: 17 B: 3 29 2 
Number of granules per g starch 108 6 10 A: 2 B: 80 0.5 1600 
Specific surface area (m2/kg) 270 220 A: 180 B: 670 75 670 
Average granule diameters can be calculated based on the 
number of granules or on their volume (or weight). Number 
mean diameter (D [1,0]) and volume mean diameter (D [4,3]) 
can be calculated with formulas (3) and (4), 
D½1; 0 ¼ 
P 
n1 
di 
n 
ð3Þ 
D½4; 3 ¼ 
P 
n1 
d4i 
P 
n1 
d3i 
ð4Þ 
where di is the diameter of particle i and n is the total number 
of particles. Based on the number mean diameter, the 
number of granules per gram starch can be determined. This 
allows estimating that a given weight of rice starch contains 
about 3000 times more granules than a same given weight of 
potato starch. The specific surface area, i.e. the total surface 
area per unit of volume, of rice starch is about ten times that 
of potato starch. The values for specific surface area in Table 4 
are probably an underestimation, as the presence of channels 
and pores on the granule surface is not included in the 
calculation. However, to the best of our knowledge, no 
information is available on this. Especially cereal starches 
contain pores [94, 95], while for potato and cassava starches 
also depressions and protrusions on the granule surface 
have been observed [96]. 
8 Gelatinisation properties 
Heating starch in excess water (1:2 starch:water) above a 
certain temperature, the ‘gelatinisation temperature’, dis-rupts 
the molecular order of the granules and melts the 
crystallites [97]. When relatively less water (1:2 starch:water) 
is available, gelatinisation is partly postponed to higher 
temperatures [16]. Table 5 lists gelatinisation onset (To), peak 
(Tp) and conclusion (Tc) temperatures and melting enthalpies 
(DH) of the different starches in excess water. DH represents 
the amount of energy needed to melt all the crystals [16]. 
Wheat starch has the lowest gelatinisation temperature, 
followed by potato, cassava and maize starches [22, 25, 26, 30, 
31, 34, 71, 98–106]. Rice starches show a high variation in 
gelatinisation temperature, which is at least partly due to the 
high variation of AM content in regular rice starches [25, 30, 
Figure 3. SEM pictures of maize (a), cassava (b), wheat (c), potato (d) and rice (e) starches (own data). Bars represent 10mm. 
 2014 WILEY-VCH Verlag GmbH  Co. KGaA, Weinheim www.starch-journal.com
Starch/Stärke 2014, 66, 1–16 7 
Table 5. Gelatinisation properties of starches of different botanical origins in excess water measured with DSC 
Reference Starch-to-water ratio To (°C)a) TP (°C)b) Tc (°C)c) TcTo (°C) DH (J/g)d) 
Maize starch [22, 26, 34] 1:3 64–67 68–71 72–75 8–11 11–12 
[98] 3:7 63 67 72 9 9 
[103] 1:4 67 71 – – 12 
[104] 1:9 66 71 – – 12 
Cassava starch [31, 34] 1:3 55–64 61–68 71–74 10–16 15–19 
[98] 3:7 60 65 75 15 11 
[103] 1:4 64–66 68–70 – – 11–14 
[104] 1:9 65 71 – – 13 
Wheat starch [22, 34, 100] 1:3 53–62 61–65 64–69 7–12 9–12 
[98, 105, 108] 3:7 48–57 54–62 58–68 9–11 7–15 
[103] 1:4 60 65 – – 10 
Potato starch [22, 26, 31, 34, 71] 1:3 58–63 61–68 68–73 8–11 15–24 
[98, 99, 101, 102] 3:7 57–66 61–70 67–75 7–12 12–18 
[103] 1:4 62 65 – – 17 
[104] 1:9 59 64 – – 17 
Rice starch [34] 1:3 70 76 80 10 13 
[37] 3:7 61–76 67–79 72–85 8–13 8–14 
[25] 3:11 51–70 58–74 64–79 9–13 8–12 
[104] 1:9 58 65 – – 12 
[30] 1:2 57–76 63–79 71–83 8–19 17–20 
34, 37, 104]. For a single starch granule, loss of molecular 
order typically takes place over a very small temperature range 
(1°C) [107], while gelatinisation occurs over a wider interval 
for a population of granules. The gelatinisation temperature 
range of starches mostly varies from 8 to 12°C [22, 25, 26, 30, 
31, 34, 37, 71, 98–105, 108], although some exceptions with a 
larger temperature range have also been described [30, 31, 
98]. DH of potato starch is slightly higher than that of the 
other starches. Wheat and maize starches have a low DH [22, 
25, 26, 30, 31, 34, 37, 71, 98–105, 108]. 
The gelatinisation temperature is a measure of crystal 
quality, while DH is a measure of both crystal quality and 
quantity [109]. The presence of a relatively high amount 
of short AP chains (DP14) reduces the gelatinisation 
temperature, while a relatively high amount of longer chains 
leads to an increased gelatinisation temperature [30, 31, 105]. 
According to Gidley and Bulpin [110], a chain DP of at least 10 
is needed to form double helices. Consequently, starches with 
a relatively high level of short AP chains have lower crystalline 
order, which leads to a lower gelatinisation temperature. 
Relatively high amounts of longer chains may be responsible 
for both better stabilisation of the crystal structure over a 
longer distance as well as for a higher gelatinisation 
temperature [31]. Although CL of potato AP (23–34) is 
higher than CL of the other starches (17–28) (Table 3), its 
gelatinisation temperature is relatively low (To¼57–66°C) 
(Table 5), probably because of the presence of phosphate 
monoesters and the more open crystal structure of B-type 
than of A-type starches [26, 34, 88]. The importance of the AP 
chain length for gelatinisation temperature is expressed by 
the Gibbs–Thomson Eq. (5) for lamellar crystallites 
  
m 1  2g 
Tm ¼ T0 
DHlc 
ð5Þ 
This equation relates the melting temperature (Tm) to the 
average crystalline layer thickness (lc), the melting tempera-ture 
of an ideal crystal with an infinite crystal size (T0 
m), the 
lamellar surface free energy (g) and DH. Longer AP chains 
lead to increased lc and thus also Tm. A high Tm is obtained 
when branch chain lengths are relatively long (high lc) and 
when crystal quality is high (low g and high DH) [111]. 
Branch chain length can also influence DH. Starches with 
higher CL (e.g. potato starch) appear to have a higher DH[34]. 
DH is also correlated with crystallinity, i.e. a higher degree of 
crystallinity (e.g. waxy starches) leads to a higher DH [26, 31, 
34, 105]. Furthermore, the presence of lipids in cereal 
starches might also explain the difference in DH between 
potato and cereal starches. The exothermic formation of AM 
lipid complexes can occur simultaneously with gelatinisation, 
thereby lowering the measured DH [22]. 
9 Swelling power and solubility 
At roomtemperature, starch granules can absorb up to 30% of 
their weight in excess water without swelling noticeably [112]. 
a) To, gelatinisation onset temperature; 
b) Tp, gelatinisation peak temperature, 
c) Tc, gelatinisation conclusion temperature; 
d) DH, gelatinisation enthalpy. 
 2014 WILEY-VCH Verlag GmbH  Co. KGaA, Weinheim www.starch-journal.com
8 J. Waterschoot et al. Starch/Stärke 2014, 66, 1–16 
Table 6. Swelling power and solubility of starches of different botanical origins 
Reference Starch concentration (%) Temperature (°C) Swelling power (g/g) Solubility (%) 
Maize starch [103] 0.5 84 16 – 
[115] 1.0 95 23 17 
[122] 2.0 95 11 18 
Cassava starch [103] 0.5 84 40–60 – 
[114] 0.3 90 50 13 
Wheat starch [103] 0.5 84 40 – 
[105] 2.0 90 13–25 – 
[122] 2.0 95 9 20 
Potato starch [103] 0.5 84 168 – 
[116] – 90 26–49 3–37 
[114] 0.3 90 60–130 12–17 
[115] 1.0 95 91 12 
Rice starch [117] 1.0 95 25–45 13–32 
[37] 2.0 90 17–39 – 
However, during heating, starch granules absorb much 
more water and swell. At higher temperatures, part of 
the polysaccharides go into solution and leach out of the 
granules [113]. Table 6 lists the swelling powers and 
solubilities of the different starches. The former are the 
amounts of water a starch can absorb per gram starch at a 
certain temperature and at a certain starch concentration, 
while the solubilities represent the percentages of leached AM 
and AP at this temperature. 
Potato starch has a much higher swelling power than 
other starches [103, 114–116]. As mentioned above, this is 
largely due to its negatively charged phosphate monoesters. 
The swelling power of cassava starch is also higher than that 
of the cereal starches [37, 103, 105, 114, 115, 117]. Granule 
swelling is mainly attributed to AP and is inhibited by 
AM [118]. As a result, waxy starches have a higher swelling 
power than their AM containing counterparts [34, 114, 117, 
118]. Removal of endogenous starch lipids reduces formation 
of AM lipid complexes and increases the swelling power of 
wheat and maize starches but not up to the level of swelling 
that characterises tuber starches [21, 118]. The impact of 
exogenous lipids on swelling of starch granules depends on 
the type of lipid added and the temperature [19, 119–121]. The 
absence of AM lipid complexes in potato and cassava starches 
also contributes to their extensive swelling power [34, 114]. 
Starch solubility, determined at temperatures from 84 to 
95°C, varies from 3 to 37%, with values for most starches 
between 10 and 20% (Table 6) [103, 105, 114–117, 122]. Both 
AM and AP leach out of granules of regular starches, while 
evidently only AP leaches from waxy starches. Usually, in 
excess water, AM leaching starts at relatively low temper-atures 
(70°C), while AP only leaches out at higher 
temperatures (90°C). This has been observed for potato 
and cassava [114], rice [117], maize [123] and wheat [21, 119, 
123] starches. 
10 Pasting properties 
Pasting can be described as a term encompassing the events 
that occur after gelatinisation in a starch suspension, i.e. 
further swelling of the granules, leaching of polysaccharides, 
increase in viscosity and formation of an AM gel network [97]. 
Changes in viscosity depend on the concentration of the 
starch suspension and can be measured with a Rapid Visco 
Analyser or a Viscoamylograph. Figure 4 shows typical 
pasting profiles for maize, cassava, wheat, potato and rice 
starches (own data). The pasting temperature is that at which 
an onset in viscosity rise can be observed. Table 7 lists pasting 
temperatures, and peak and cold paste viscosities of the 
different starches. While the term peak viscosity speaks for 
itself, cold paste viscosity is that measured at the end of the 
Figure 4. Pasting profiles of 8.0% suspensions of potato, cassava, 
maize, rice and wheat starches in water measured with a rapid visco 
analyser. The following temperature-time profile was used at a 
stirring speed of 160 rpm: 1 min at 50°C, heating from 50 to 95°C 
in 9 min, 10 min at 95°C, cooling from 95 to 50°C in 15 min, 
10 min at 50°C (own data). 
 2014 WILEY-VCH Verlag GmbH  Co. KGaA, Weinheim www.starch-journal.com
Starch/Stärke 2014, 66, 1–16 9 
heating and cooling cycle. Potato and cassava starches have 
lower pasting temperatures than cereal starches. Due to the 
high level of negatively charged phosphate groups in potato 
starch, viscosity development starts at lower temperatures. 
Potato starch reaches a very high peak viscosity. The peak 
viscosities of cereal starches are rather low because of their 
lower swelling power [34, 37, 105, 116, 124, 125]. Waxy 
starches reach higher peak viscosities than their AM 
containing counterparts due to their higher swelling 
power [117]. It is of note that starches with a high swelling 
power are more sensitive to granule breakdown at high 
temperatures than those with a lower swelling power. Pasting 
profiles of potato starch show a large breakdown during the 
holding phase at 95°C [114, 124].Waxy starches also have low 
thermal stability and quickly disintegrate during pasting at 
high temperatures [22]. 
During cooling of a starch suspension, viscosity 
increases as the leached AM forms a gel network [126]. 
Cold paste viscosities of regular wheat, maize and rice 
starches are at least as high as their peak viscosities, while 
cold paste viscosities of potato and cassava starches 
aremuch lower than their peak viscosities due to significant 
granule breakdown [34, 105, 116, 124, 125]. After some 
hours, AM forms stable crystalline structures, which are 
acid resistant and have a melting temperature of about 
150 °C [16, 126]. AMcrystallisation is largely responsible for 
Table 7. Pasting properties of starches of different botanical origins 
Reference Starch concentration (%) Pasting temperature (°C) Peak viscosity (mPa s) Cold paste viscosity (mPa s) 
Maize starch [124] 8 82 2100 2000 
[34] 8 82 1800 2000 
Cassava starch [124] 8 67 2300 1400 
[34] 8 68 2100 1300 
Wheat starch [34] 8 89 1250 1850 
[125] 9 – 2100 2950 
[105] 11 82–90 2250–3400 2600–4000 
Potato starch [124] 8 67 9500 3400 
[34] 8 64 8400 2750 
[116] 11 65–70 4100–7200 2300–3400 
Rice starch [124] 8 71 2500 2050 
[34] 8 80 1350 1900 
[37] 6 72–80 1000–2400 1500–3500 
Table 8. Retrogradation properties of starches of different botanical origins measured with DSC 
Reference Starch-to-water ratio Storage conditions [time (days)/temperature (°C)] DH (J/g)a) 
Maize starch [34] 1:3 7/4 5.8 
[98] 3:7 7/22 3.0 
Cassava starch [114] 1:3 28/20 No retrogradation detected 
[34] 1:3 7/4 3.7 
[98] 3:7 7/22 No retrogradation detected 
Wheat starch [34] 1:3 7/4 3.6 
[98] 3:7 7/22 2.0 
[105] 3:7 7/4 0.7–3.0 
[108] 3:7 28/5 10.1–10.6 
Potato starch [114] 1:3 7/20 2.8–7.0 
[114] 1:3 28/20 6.3–9.9 
[34] 1:3 7/4 7.5 
[98] 3:7 7/22 4.2 
[99] 3:7 14/4 6.4–8.6 
Rice starch [34] 1:3 7/4 5.3 
[140] 1:2 28/RTb) 1.4–3.1, 7.8–11.6 
[140] 1:2 28/6 5.7–11.1 
a) DH, melting enthalpy of retrograded amylopectin; 
b) RT, room temperature. 
 2014 WILEY-VCH Verlag GmbH  Co. KGaA, Weinheim www.starch-journal.com
10 J. Waterschoot et al. Starch/Stärke 2014, 66, 1–16 
the initial firmness of a starch gel, while AP retrogradation 
(cf. infra) is responsible for the long-term changes in gel 
firmness [127]. 
11 Amylopectin retrogradation 
Retrogradation is the process of crystallisation of AP 
molecules in a starch paste [16]. The term retrogradation, 
which means ‘to go back’, is here only used for AP and not 
for AM, as strictly speaking only AP molecules can go back 
to a crystalline entity [16]. AP crystals are acid labile and 
melt at 50–60°C [87]. Table 8 shows the melting enthalpies 
of retrograded amylopectin. The extent of retrogradation 
highly depends on the storage time and temperature, starch 
concentration and starch structural properties [45, 128]. 
It consists of three steps: nucleation, propagation and 
maturation. Nucleation occurs faster at a relatively low 
temperature, close to the Tg of the starch, while propagation 
and maturation occur faster at a temperature close to the 
melting temperature [129]. Besides storage temperature, 
also the starch-to-water ratio has an important effect on 
retrogradation. Water content should neither be too high 
(80%) nor too low (30%) to allow retrogradation [130]. 
Evidently, AP structure also impacts retrogradation. Short 
AP chains (DP 6–9) are less susceptible to retrogradation 
than chains with DP 14–24 [114, 131]. Asmentioned earlier, 
at least a DP of 10 is needed for forming double 
helices [110]. This could explain the higher extent of 
retrogradation of potato starch than of cereal starches under 
the same conditions (Table 8) [35, 132]. During storage of 
starch gels, interactions or even co-crystallisation with AM 
can occur, especially when the AM content is relatively 
high [133–136]. 
12 Starch functionality 
12.1 Paste firmness 
As mentioned above, due to AM crystallisation (short-term) 
and AP retrogradation (long-term), firmness of starch gels 
increases with time [115, 127]. This is an important aspect of 
starch functionality in food products. Gel firmness can be 
measured by compressing a gel by a certain percentage. It 
depends on starch concentration, storage time and tempera-ture 
[137]. Table 9 lists the firmness of gels of maize, cassava, 
wheat, potato and rice starches. As different starch concen-trations 
and storage conditions are used, it is difficult to 
compare results. When comparing firmness values obtained 
under the same conditions, maize, wheat and potato starches 
show similar firmness values which are higher than that of 
a cassava starch gel [138, 139]. This could be due to the 
relatively low AM content of cassava starch and the large loss 
of granular integrity in the gel [139]. Rice starch gels have a 
broad range of firmness values (0.09–7N) [140]. AM content 
is positively correlated with gel firmness. Waxy starch gels 
have only a low firmness as a result of their poor gel network 
formation [140]. In contrast, lipid content is negatively 
correlated with gel firmness. The formation of AM lipid 
complexes reduces the amount of AM available for network 
formation [122, 141, 142]. 
12.2 Paste clarity 
Another important characteristic in many starch applications 
in food systems is paste clarity. The presence of relatively 
short chains of AM or AP adds to opacity in food products. 
While for a range of products including sauces, dressings and 
puddings this is not a problem, products such as fruit fillings 
Table 9. Firmness of gels of starches of different botanical origins 
Reference Starch content (%) Storage conditions [time (days)/temperature (°C)] Compressed proportion of gel (%) Gel firmness (N) 
Maize starch [138] 8 1/RTa) 40 0.64 
[138] 8 7/4 40 1.03 
[139] 6 1/4 33 0.93 
[141] 6 0.2/25 11 0.04 
[141] 6 1/4 11 0.05 
Cassava starch [138] 8 1/RTa) 40 Not measurable 
[138] 8 7/4 40 0.35 
[139] 6 1/4 33 0.19 
Wheat starch [138] 8 1/RTa) 40 0.69 
[138] 8 7/4 40 0.70 
[125] 9 1/4 Not reported 0.50 
Potato starch [138] 8 1/RTa) 40 0.56 
[138] 8 7/4 40 0.71 
Rice starch [140] 8 2/6 10 0.09–4.17 
[140] 8 14/6 10 0.12–7.05 
a) RT, room temperature. 
 2014 WILEY-VCH Verlag GmbH  Co. KGaA, Weinheim www.starch-journal.com
Starch/Stärke 2014, 66, 1–16 11 
and jellies require the starch pastes to be of high clarity [143]. 
Paste clarity can be determined by measuring the light 
transmittance (at 650 nm) of a 1.0% starch paste. Potato 
starch (42–96% light transmittance) has the highest 
paste clarity, followed by cassava starch (51–81% light 
transmittance) and the cereal starches (13–62% light 
transmittance) [144–148]. The clarity of waxy cereal starch 
gels is better than that of their AM containing counter-parts 
[144–146]. The very high paste clarity of potato starch 
gels is caused by the absence of granule remnants in the 
gel (cf. high swelling power and high granular break-down) 
[144]. Pastes of cassava and waxy starches also 
show few granule remnants. However, more interactions 
between leached material lead to a higher opacity of these 
gels. The low clarity of pastes of regular cereal starches 
is caused by swollen granule remnants [144] and by the 
presence of AM lipid complexes [149]. During storage 
of starch gels, their paste clarity decreases as more and 
more association of AM and/or AP molecules takes 
place [146, 150]. 
12.3 Freeze–thaw stability 
Freeze–thaw stability is an important quality aspect of starch 
gels. When a starch gel undergoes repeated freezing and 
thawing cycles, it releases water. The gel is then said to 
undergo syneresis. The extent of syneresis is a measure for its 
freeze–thaw stability [128, 151]. During freezing, phase 
separation takes place as ice crystals are formed. As a result, 
starch is concentrated in the unfrozen matrix. During 
thawing of the gel, association of AM and AP takes place in 
the more concentrated zones. As a result, increasingly 
insoluble aggregates are formed. Ice crystals melt and the 
water is not reabsorbed by the starch gel. A sponge like 
structure develops and the melted water readily separates 
from the gel [152]. 
Different factors have an influence on the freeze–thaw 
stability. These include the freezing rate, the botanical origin 
of the starch and its AM content, the starch concentration, the 
number of freeze–thaw cycles and the sample prepara-tion 
[128, 153]. The freezing rate is an important factor 
affecting syneresis. Slow freezing favours formation of large 
ice crystals which disrupt the gel structure to a larger extent 
than small ice crystals [128, 152, 154]. Slow freezing also leads 
to maximal ice formation and AM self-association which 
result in significant structure loss and syneresis. The latter is 
positively correlated with AM content [153]. Native, AM 
containing starches have very low freeze–thaw stability. 
Syneresis readings after several freeze–thaw cycles for maize 
starch (47–79%) [153, 155, 156], cassava starch (50–67%) [153, 
155], wheat starch (44–68%) [153, 155, 156] and potato starch 
(60–76%) [153, 155, 157] are very high. Rice starch shows a 
broad range of syneresis values (7–75%) [153, 155, 158, 159]. 
Waxy starches are intrinsically more stable than AM 
containing starches, although they can also undergo strong 
textural changes during freeze–thaw cycles [151, 155]. Waxy 
rice starch has a very good freeze–thaw stability. A relatively 
high proportion of AP side chains with DP 6–12, which is the 
case for waxy rice starch, is believed to be at the basis of a 
reduced extent of syneresis [153, 157]. In contrast, syneresis 
of waxy maize starch gels is comparable to that of gels from 
regular maize starch [153, 155]. For different waxy maize 
starches, no correlation was observed between syneresis and 
retrogradation enthalpy as measured with DSC. Syneresis 
values were already maximal, when little, if any, double helical 
order was present. This indicates that the early stages of 
AM and/or AP association (before the formation of real 
crystal structures) already cause syneresis due to network 
formation [160]. 
The presence of lipids in cereal starches may also 
contribute to a relatively high freeze–thaw stability. Granular 
swelling and leaching of AM are reduced in the presence of 
lipids. As a result, starch molecules remain close to each 
other in the granules. This facilitates their reassociation and 
in this way it may contribute to a low freeze–thaw 
stability [153]. 
In conclusion, there are significant differences in 
structural, physicochemical and functional properties of 
starches of different botanical origins. Starch is widely used 
by the food and non-food industry in a broad range of 
products. The overview of starch properties provided in this 
review can be of assistance when developing starches for 
specific purposes. The botanical origin of the used starch has 
a great impact on final product properties. Within a botanical 
origin, in some instances the starch AM content is a major 
determinant of its properties and starch should therefore be 
selected with great care. 
The authors gratefully acknowledge Flanders’ FOOD (Brus-sels, 
Belgium) and the Methusalem programme ‘Food for the 
future’ (KU Leuven) for financial support. J.A.D. is W.K. Kellogg 
Chair in Cereal Science and Nutrition at the KU Leuven. 
The authors have declared no conflict of interest. 
13 References 
[1] Schwartz, D., Whistler, R., in: BeMiller, J., Whistler, R. 
(Eds.), Starch: Chemistry and Technology,Academic Press, 
New York 2009, pp. 1–10. 
[2] Mitchell, C. R., in: BeMiller, J., Whistler, R. (Eds.), Starch: 
Chemistry and Technology, Academic Press, New York 
2009, pp. 569–578. 
[3] Grommers, H. E., van der Krogt, D. A., in: BeMiller, J., 
Whistler, R. (Eds.), Starch: Chemistry and Technology, 
Academic Press, New York 2009, pp. 511–539. 
[4] Hobbs, L., in: BeMiller, J., Whistler, R. (Eds.), Starch: 
Chemistry and Technology, Academic Press, New York 
2009, pp. 797–832. 
 2014 WILEY-VCH Verlag GmbH  Co. KGaA, Weinheim www.starch-journal.com
12 J. Waterschoot et al. Starch/Stärke 2014, 66, 1–16 
[5] Mason, W. R., in: BeMiller, J., Whistler, R. (Eds.), Starch: 
Chemistry and Technology, Academic Press, New York 
2009, pp. 745–795. 
[6] Chiu, C., Solarek, D., in: BeMiller, J., Whistler, R. (Eds.), 
Starch: Chemistry and Technology, Academic Press, New 
York 2009, pp. 629–655. 
[7] Moorthy, S. N., in: Eliasson, A. C. (Ed.), Starch in food: 
Structure, Function and Applications, Woodhead Publish-ing, 
Cambridge 2004, pp. 321–359. 
[8] Bergthaller, W., in: Eliasson, A. C. (Ed.), Starch in food: 
Structure, Function and Applications, Woodhead Publish-ing, 
Cambridge 2004, pp. 241–257. 
[9] Maningat, C. C., Seib, P. A., Bassi, S. D., Woo, K. S., 
Lasater, G. D., in: BeMiller, J., Whistler, R. (Eds.), Starch: 
Chemistry and Technology, Academic Press, New York 
2009, pp. 441–510. 
[10] Eckhoff, S. R., Watson, S. A., in: BeMiller, J., Whistler, R. 
(Eds.), Starch: Chemistry and Technology,Academic Press, 
New York 2009, pp. 373–439. 
[11] Boundy, J. A., Turner, J. E., Wall, J. S., Dimler, R. J., 
Influence of commercial processing on composition and 
properties of corn zein. Cereal Chem. 1967, 44, 281–287. 
[12] Jacobs, H., Delcour, J. A., Hydrothermal modifications of 
granular starch, with retention of the granular structure: A 
review. J. Agric. Food Chem. 1998, 46, 2895–2905. 
[13] Breuninger, W. F., Piyachomkwan, K., Sriroth, K., in: 
BeMiller, J., Whistler, R. (Eds.), Starch: Chemistry and 
Technology, Academic Press, New York 2009, pp. 541– 
568. 
[14] Van Der Borght, A., Goesaert, H., Veraverbeke, W. S., 
Delcour, J. A., Fractionation of wheat and wheat flour into 
starch and gluten: Overview of the main processes and the 
factors involved. J. Cereal Sci. 2005, 41, 221–237. 
[15] Cornell, H., in: Eliasson, A. C. (Ed.), Starch in Food: 
Structure, Function and Applications, Woodhead Publish-ing, 
Cambridge 2004, pp. 211–240. 
[16] Delcour, J. A., Hoseney, R. C., Principles of Cereal Science 
and Technology, AACC International, Inc., St. Paul, MN 
2010. 
[17] Vasanthan, T., Hoover, R., A comparative study of the 
composition of lipids associated with starch granules from 
various botanical sources. Food Chem. 1992, 43, 19–27. 
[18] Buléon, A., Colonna, P., Planchot, V., Ball, S., Starch 
granules: Structure and biosynthesis. Int. J. Biol.Macromol. 
1998, 23, 85–112. 
[19] Raphaelides, S. N., Georgiadis, N., Effect of fatty acids on 
the rheological behaviour of maize starch dispersions during 
heating. Carbohydr. Polym. 2006, 65, 81–92. 
[20] Putseys, J. A., Lamberts, L., Delcour, J. A., Amylose-inclusion 
complexes: Formation, identity and physico-chemical 
properties. J. Cereal Sci. 2010, 51, 238–247. 
[21] Shi, Y.C., Seib, P.A., Lu, S.P.W., in: Levine, H., Slade, L., 
(Eds.), Water relationships in foods: Advances in the 1980s 
and trends for the 1990s, Plenum Press, New York 1991, 
pp. 667–686. 
[22] Schirmer,M., Höchstötter,A., Jekle, M.,Arendt, E., Becker, 
T., Physicochemical and morphological characterization of 
different starches with variable amylose/amylopectin ratio. 
Food Hydrocolloids 2013, 32, 52–63. 
[23] Leelavathi, K., Indrani, D., Sidhu, J. S., Amylograph pasting 
behaviour of cereal and tuber starches. Starch/Stärke 1987, 
39, 378–381. 
[24] Skerritt, J. H., Frend, A. J., Robson, L. G., Greenwell, P., 
Immunological homologies between wheat gluten and 
starch granule proteins. J. Cereal Sci. 1990, 12, 123–136. 
[25] Hoover, R., Sailaja, Y., Sosulski, F. W., Characterization of 
starches fromwild and long grain brown rice. Food Res. Int. 
1996, 29, 99–107. 
[26] Dhital, S., Shrestha, A. K., Hasjim, J., Gidley, M. J., 
Physicochemical and structural properties of maize and 
potato starches as a function of granule size. J. Agric. Food 
Chem. 2011, 59, 10151–10161. 
[27] Singh, N., Singh, J., Kaur, L., Sodhi, N. S., Gill, B. S., 
Morphological, thermal and rheological properties of 
starches from different botanical sources. Food Chem. 
2003, 81, 219–231. 
[28] Tabata, S., Hizukuri, S., Studies on starch phosphate. Part 
2. Isolation of glucose 3-phosphate and maltose phosphate 
by acid hydrolysis of potato starch. Starch/Stärke 1971, 23, 
267–272. 
[29] Lim, S. T., Kasemsuwan, T., Jane, J. L., Characterization of 
phosphorus in starch by 31P-nuclear magnetic resonance 
spectroscopy. Cereal Chem. 1994, 71, 488–493. 
[30] Vandeputte, G. E., Vermeylen, R., Geeroms, J., Delcour, 
J. A., Rice starches, I., Structural aspects provide insight 
into crystallinity characteristics and gelatinisation behaviour 
of granular starch. J. Cereal Sci. 2003, 38, 43–52. 
[31] Gomand, S. V., Lamberts, L., Derde, L. J., Goesaert, H. 
et al., Structural properties and gelatinisation characteristics 
of potato and cassava starches and mutants thereof. Food 
Hydrocolloids 2010, 24, 307–317. 
[32] Lineback, D. R., The starch granule – organization and 
properties. Bakers Dig. 1984, 58, 16–21. 
[33] Defloor, I., Dehing, I., Delcour, J. A., Physico-chemical 
properties of cassava starch. Starch/Stärke 1998, 50, 
58–64. 
[34] Jane, J. L., Chen, Y. Y., Lee, L. F., McPherson, A. E. et al., 
Effects of amylopectin branch chain length and amylose 
content on the gelatinization and pasting properties of 
starch. Cereal Chem. 1999, 76, 629–637. 
[35] Fredriksson, H., Silverio, J., Andersson, R., Eliasson, A. C., 
Aman, P., The influence of amylose and amylopectin 
characteristics on gelatinization and retrogradation proper-ties 
of different starches. Carbohydr. Polym. 1998, 35, 
119–134. 
[36] Juliano, B. O., Villareal, R. M., Perez, C. M., Villareal, 
C. P. et al., Varietal differences in properties among 
high amylose rice starches. Starch/Stärke 1987, 39, 
390–393. 
[37] Singh, N., Kaur, L., Sandhu, K. S., Kaur, J., Nishinari, K., 
Relationships between physicochemical, morphological, 
thermal, rheological properties of rice starches. Food 
Hydrocolloids 2006, 20, 532–542. 
[38] Lafiandra, D., Sestili, F., D’Ovidio, R., Janni, M. et al., 
Approaches formodification of starch composition in durum 
wheat. Cereal Chem. 2010, 87, 28–34. 
[39] Rolland-Sabaté, A., Sánchez, T., Buléon, A., Colonna, P. 
et al., Structural characterization of novel cassava starches 
with low and high-amylose contents in comparison with 
other commercial sources. Food Hydrocolloids 2012, 27, 
161–174. 
[40] Raemakers, K., Schreuder,M., Suurs, L., Furrer-Verhorst,H. 
et al., Improved cassava starch by antisense inhibition of 
granule-bound starch synthase I. Mol. Breed. 2005, 16, 
163–172. 
 2014 WILEY-VCH Verlag GmbH  Co. KGaA, Weinheim www.starch-journal.com
Starch/Stärke 2014, 66, 1–16 13 
[41] Hung, P. V.,Maeda, T.,Morita, N., Study on physicochemi-cal 
characteristics of waxy and high-amylose wheat 
starches in comparison with normal wheat starch. 
Starch/Stärke 2007, 59, 125–131. 
[42] Wang, Y. J., Wang, L. F., Structures of four waxy rice 
starches in relation to thermal, pasting, and textural 
properties. Cereal Chem. 2002, 79, 252–256. 
[43] Sanchez, T., Dufour, D., Moreno, I. X., Ceballos, H., 
Comparison of pasting and gel stabilities of waxy and 
normal starches from potato, maize, and rice with those of a 
novel waxy cassava starch under thermal, chemical, and 
mechanical stress. J. Agric. Food Chem. 2010, 58, 5093– 
5099. 
[44] Matveev, Y. I., van Soest, J. J. G., Nieman, C., 
Wasserman, L. A. et al., The relationship between 
thermodynamic and structural properties of low and high 
amylose maize starches. Carbohydr. Polym. 2001, 44, 
151–160. 
[45] Yuan,R.C.,Thompson,D. B.,Boyer,C.D., Finestructureof 
amylopectin in relation to gelatinization and retrogradation 
behavior of maize starches from three wx-containing 
genotypes in two inbred lines. Cereal Chem. 1993, 70, 
81–89. 
[46] Shi, Y. C., Capitani, T., Trzasko, P., Jeffcoat, R., Molecular 
structure of a low-amylopectin starch and other high-amylose 
maize starches. J. Cereal Sci. 1998, 27, 
289–299. 
[47] Schwall, G. P., Safford, R., Westcott, R. J., Jeffcoat, R. 
et al., Production of very-high-amylose potato starch by 
inhibition of SBE A and B. Nat. Biotechnol. 2000, 18, 551– 
554. 
[48] Karlsson, M. E., Leeman,A.M., Bjorck, I. M. E., Eliasson,A.- 
C., Some physical and nutritional characteristics of 
genetically modified potatoes varying in amylose/amylo-pectin 
ratios. Food Chem. 2007, 100, 136–146. 
[49] Uwer, U., Frohberg, C., Pilling, J., Landschütze, V., US 
Patent 7112718, 2006. 
[50] Regina, A., Bird, A., Topping, D., Bowden, S. et al., High-amylose 
wheat generated by RNA interference improves 
indices of large-bowel health in rats. Proc. Natl. Acad. Sci. 
USA 2006, 103, 3546–3551. 
[51] Zhu, L., Gu, M., Meng, X., Cheung, S. C. K. et al., High-amylose 
rice improves indices of animal health in normal 
and diabetic rats. Plant Biotechnol. J. 2012, 10, 
353–362. 
[52] Yamamori,M., Fujita, S., Hayakawa, K.,Matsuki, J., Yasui, 
T., Genetic elimination of a starch granule protein, SGP-1, of 
wheat generates an altered starch with apparent high 
amylose. Theor. Appl. Genet. 2000, 101, 21–29. 
[53] Hogg, A. C., Gause, K., Hofer, P., Martin, J. M. et al., 
Creation of a high-amylose durum wheat through mutagen-esis 
of starch synthase II (SSIIa). J. Cereal Sci. 2013, 57, 
377–383. 
[54] Konik-Rose, C., Thistleton, J., Chanvrier, H., Tan, I. et al., 
Effects of starch synthase IIa gene dosage on grain, protein 
and starch in endosperm of wheat. Theor. Appl. Genet. 
2007, 115, 1053–1065. 
[55] Ceballos, H., Sanchez, T., Denyer, K., Tofino, A. P. et al., 
Induction and identification of a small-granule, high-amylose 
mutant in cassava (Manihot esculenta Crantz). J. Agric. 
Food Chem. 2008, 56, 7215–7222. 
[56] Ball, S., Guan, H. P., James, M., Myers, A. et al., From 
glycogen to amylopectin: A model for the biogenesis of the 
plant starch granule. Cell 1996, 86, 349–352. 
[57] Gelders, G. G., Vanderstukken, T. C., Goesaert, H., 
Delcour, J. A., Amylose–lipid complexation: A new 
fractionation method. Carbohydr. Polym. 2004, 56, 
447–458. 
[58] Billmeyer, F. W. J., Textbook of Polymer Science, John 
Wiley  Sons, Inc., New York 1984. 
[59] Takeda, Y., Shirasaka, K., Hizukuri, S., Examination of the 
purity and structure of amylose by gel-permeation chroma-tography. 
Carbohydr. Res. 1984, 132, 83–92. 
[60] Ong,M.H.,Jumel, K.,Tokarczuk,P.F., Blanshard, J.M. V., 
Harding, S. E., Simultaneous determinations of the 
molecular weight distributions of amyloses and the fine 
structures of amylopectins of native starches. Carbohydr. 
Res. 1994, 260, 99–117. 
[61] Takeda, Y., Hizukuri, S., Takeda, C., Suzuki, A., Structures 
of branched molecules of amyloses of various origins and 
molar fractions of branched and unbranched molecules. 
Carbohydr. Res. 1987, 165, 139–145. 
[62] Shibanuma, K., Takeda, Y., Hizukuri, S., Shibata, S., 
Molecular structures of some wheat starches. Carbohydr. 
Polym. 1994, 25, 111–116. 
[63] Takeda, Y., Preiss, J., Structures of B90 (sugary) andW64a 
(normal) maize starches. Carbohydr. Res. 1993, 240, 
265–275. 
[64] Takeda, Y., Hizukuri, S., Juliano, B. O., Purification and 
structure of amylose from rice starch. Carbohydr. Res. 
1986, 148, 299–308. 
[65] Chen, M. H., Bergman, C. J., Method for determining the 
amylose content, molecular weights, and weight- and 
molar-based distributions of degree of polymerization of 
amylose and fine-structure of amylopectin. Carbohydr. 
Polym. 2007, 69, 562–578. 
[66] Takeda, Y., Shibahara, S., Hanashiro, I., Examination of the 
structure of amylopectin molecules by fluorescent labeling. 
Carbohydr. Res. 2003, 338, 471–475. 
[67] Shibanuma, Y., Takeda, Y., Hizukuri, S., Molecular and 
pasting properties of some wheat starches. Carbohydr. 
Polym. 1996, 29, 253–261. 
[68] Takeda, Y., Hizukuri, S., Juliano, B. O., Structures of rice 
amylopectins with low and high affinities for iodine. 
Carbohydr. Res. 1987, 168, 79–88. 
[69] Gomand, S. V., Verwimp, T., Goesaert, H., Delcour, J. A., 
Structural and physicochemical characterisation of rye 
starch. Carbohydr. Res. 2011, 346, 2727–2735. 
[70] Hizukuri, S., Relationship between the distribution of 
the chain length of amylopectin and the crystalline 
structure of starch granules. Carbohydr. Res. 1985, 141, 
295–306. 
[71] Yoo, S. H., Perera, C., Shen, J., Ye, L. et al., Molecular 
structure of selected tuber and root starches and effect of 
amylopectin structure on their physical properties. J. Agric. 
Food Chem. 2009, 57, 1556–1564. 
[72] Hizukuri, S., Kaneko, T., Takeda, Y., Measurement of the 
chain length of amylopectin and its relevance to the origin of 
crystalline polymorphism of starch granules. Biochim. 
Biophys. Acta 1983, 760, 188–191. 
[73] Laohaphatanaleart, K., Piyachomkwan, K., Sriroth, K., 
Santisopasri, V., Bertoft, E.,Astudy of the internal structure 
in cassava and rice amylopectin. Starch/Stärke 2009, 61, 
557–569. 
[74] Hizukuri, S., Takeda, Y., Maruta, N., Juliano, B. O., 
Molecular structures of rice starch. Carbohydr. Res. 
1989, 189, 227–235. 
 2014 WILEY-VCH Verlag GmbH  Co. KGaA, Weinheim www.starch-journal.com
14 J. Waterschoot et al. Starch/Stärke 2014, 66, 1–16 
[75] Zhu, F., Bertoft, E., Kallman, A., Myers, A. M., Seethara-man, 
K., Molecular structure of starches from maize 
mutants deficient in starch synthase III. J. Agric. Food 
Chem. 2013, 61, 9899–9907. 
[76] Jobling, S. A., Schwall, G. P.,Westcott, R. J., Sidebottom, 
C. M. et al., A minor form of starch branching enzyme in 
potato (SolanumtuberosumL.) tubers has a major effect on 
starch structure: Cloning and characterisation of multiple 
forms of SBE A. Plant J. 1999, 18, 163–171. 
[77] Baba, T., Arai, Y., Structural features of amylomaize starch. 
3. Structural characterization of amylopectin and intermedi-ate 
material in amylomaize starch granules. Agric. Biol. 
Chem. 1984, 48, 1763–1775. 
[78] Takeda, C., Takeda, Y., Hizukuri, S., Structure of the 
amylopectin fraction of amylomaize. Carbohydr. Res. 1993, 
246, 273–281. 
[79] Song, Y., Jane, J., Characterization of barley starches of 
waxy, normal, and high amylose varieties. Carbohydr. 
Polym. 2000, 41, 365–377. 
[80] Vilaplana, F., Gilbert, R. G., Characterization of branched 
polysaccharides using multiple-detection size separation 
techniques. J. Sep. Sci. 2010, 33, 3537–3554. 
[81] Cave, R. A., Seabrook, S. A., Gidley, M. J., Gilbert, R. G., 
Characterization of starch by size-exclusion chromatogra-phy: 
The limitations imposed by shear scission. Biomacro-molecules 
2009, 10, 2245–2253. 
[82] Thompson, D. B., On the non-random nature of 
amylopectin branching. Carbohydr. Polym. 2000, 43, 
223–239. 
[83] Jane, J. L., Xu, A., Radosavljevic, M., Seib, P. A., Location 
of amylose in normal starch granules. I. Susceptibility of 
amylose and amylopectin to cross-linking reagents. Cereal 
Chem. 1992, 69, 405–409. 
[84] Kasemsuwan, T., Jane, J., Location of amylose in normal 
starch granules. II. Locations of phosphodiester cross-linking 
revealed by Phosphorus-31 nuclear magnetic resonance. 
Cereal Chem. 1994, 71, 282–287. 
[85] Zobel, H. F.,Molecules to granules:Acomprehensive starch 
review. Starch/Stärke 1988, 40, 44–50. 
[86] Cooke, D., Gidley, M. J., Loss of crystalline and 
molecular order during starch gelatinization – Origin of 
the enthalpic transition. Carbohydr. Res. 1992, 227, 
103–112. 
[87] Colonna, P., Buléon, A., 9th International Cereal and Bread 
Congress, ICC, Paris 1992. 
[88] Vermeylen, R., Goderis, B., Reynaers, H., Delcour, J. A., 
Amylopectin molecular structure reflected in macromolecu-lar 
organization of granular starch. Biomacromolecules 
2004, 5, 1775–1786. 
[89] Jane, J. L., Kasemsuwan, T., Leas, S., Zobel, H., Robyt, 
J. F., Anthology of starch granule morphology by scanning 
electron microscopy. Starch/Stärke 1994, 46, 121–129. 
[90] Jacquier, J. C., Kar, A., Lyng, J. G., Morgan, D. J., 
McKenna, B. M., Influence of granule size on the flow 
behaviour of heated rice starch dispersions in excesswater. 
Carbohydr. Polym. 2006, 66, 425–434. 
[91] Jiang, H., Horner, H. T., Pepper, T. M., Blanco, M. et al., 
Formation of elongated starch granules in high-amylose 
maize. Carbohydr. Polym. 2010, 80, 533–538. 
[92] Kim, H. S., Huber, K. C., Physicochemical properties and 
amylopectin fine structures of A- and B-type granules of 
waxy and normal soft wheat starch. J. Cereal Sci. 2010, 
51, 256–264. 
[93] Kim, H. S., Huber, K. C., Channels within soft wheat 
starch A- and B-type granules. J. Cereal Sci. 2008, 48, 
159–172. 
[94] Fannon, J. E.,Hauber, R. J., Bemiller, J. N., Surface pores of 
starch granules. Cereal Chem. 1992, 69, 284–288. 
[95] Juszczak, L., Fortuna, T., Krok, F., Non-contact atomic 
force microscopy of starch granules surface. Part II. 
Selected cereal starches. Starch/Stärke 2003, 55, 8–16. 
[96] Juszczak, L., Fortuna, T., Krok, F., Non-contact atomic 
force microscopy of starch granules surface. Part I. Potato 
and tapioca starches. Starch/Stärke 2003, 55, 1–7. 
[97] Atwell, W. A., Hood, L. F., Lineback, D. R., Varriano- 
Marston, E., Zobel, H. F., The terminology andmethodology 
associated with basic starch phenomena. Cereal Foods 
World 1988, 33, 306–311. 
[98] Teng, L. Y., Chin, N. L., Yusof, Y. A., Rheological and 
textural studies of fresh and freeze-thawed native 
sago starch-sugar gels. II. Comparisons with other starch 
sources and reheating effects. Food Hydrocolloids 2013, 
31, 156–165. 
[99] Singh, J., Singh, N., Studies on the morphological, thermal 
and rheological properties of starch separated from some 
Indian potato cultivars. Food Chem. 2001, 75, 67–77. 
[100] Ao, Z., Jane, J. L., Characterization and modeling of the A-and 
B-granule starches of wheat, triticale, and barley. 
Carbohydr. Polym. 2007, 67, 46–55. 
[101] Kim, Y. S., Wiesenborn, D. P., Orr, P. H., Grant, L. A., 
Screening potato starch for novel properties using differen-tial 
scanning calorimetry. J. Food Sci. 1995, 60, 1060– 
1065. 
[102] Kaur, L., Singh, N., Sodhi, N. S., Some properties of 
potatoes and their starches. II. Morphological, thermal and 
rheological properties of starches. Food Chem. 2002, 79, 
183–192. 
[103] Anggraini, V., Sudarmonowati, E., Hartati, N. S., Suurs, L., 
Visser, R. G. F., Characterization of cassava starch 
attributes of different genotypes. Starch/Stärke 2009, 
61, 472–481. 
[104] Li, J. Y., Yeh, A. I., Relationships between thermal, 
rheological characteristics and swelling power for various 
starches. J. Food Eng. 2001, 50, 141–148. 
[105] Singh, S., Singh, N., Isono, N., Noda, T., Relationship of 
granule size distribution and amylopectin structure with 
pasting, thermal, and retrogradation properties in wheat 
starch. J. Agric. Food Chem. 2010, 58, 1180–1188. 
[106] Sasaki, T., Yasui, T.,Matsuki, J., Effect of amylose content 
on gelatinization, retrogradation, and pasting properties of 
starches from waxy and nonwaxy wheat and their F1 
seeds. Cereal Chem. 2000, 77, 58–63. 
[107] Liu, H., Lelièvre, J., Ayoungchee, W., A study of starch 
gelatinization using differential scanning calorimetry, X-ray 
and birefringence measurements. Carbohydr. Res. 1991, 
210, 79–87. 
[108] Sasaki, T., Yasui, T., Matsuki, J., Satake, T., Rheological 
properties of mixed gels using waxy and non-waxy wheat 
starch. Starch/Stärke 2002, 54, 410–414. 
[109] Tester, R. F., Morrison, W. R., Swelling and gelatinization of 
cereal starches. 2.Waxy rice starches. Cereal Chem. 1990, 
67, 558–563. 
[110] Gidley, M. J., Bulpin, P. V., Crystallisation of malto-oligosaccharides 
as models of the crystalline forms of 
starch:Minimum chain-length requirement for the formation 
of double helices. Carbohydr. Res. 1987, 161, 291–300. 
 2014 WILEY-VCH Verlag GmbH  Co. KGaA, Weinheim www.starch-journal.com
Starch/Stärke 2014, 66, 1–16 15 
[111] Gomand, S. V., Lamberts, L., Gommes, C. J., Visser, R. G. 
F. et al., Molecular and morphological aspects of annealing-induced 
stabilization of starch crystallites. Biomacromole-cules 
2012, 13, 1361–1370. 
[112] Kerr, R. W., Chemistry and Industry of Starch, Academic 
Press, New York 1950. 
[113] Hermansson, A. M., Svegmark, K., Developments in the 
understanding of starch functionality. Trends Food Sci. 
Technol. 1996, 7, 345–353. 
[114] Gomand, S. V., Lamberts, L., Visser, R. G. F.,Delcour, J.A., 
Physicochemical properties of potato and cassava starches 
and their mutants in relation to their structural properties. 
Food Hydrocolloids 2010, 24, 424–433. 
[115] Ring, S. G., Some studies on starch gelation. Starch/Stärke 
1985, 37, 80–83. 
[116] Kaur, A., Singh, N., Ezekiel, R., Guraya, H. S., Physico-chemical, 
thermal and pasting properties of starches 
separated from different potato cultivars grown at different 
locations. Food Chem. 2007, 101, 643–651. 
[117] Vandeputte,G. E., Derycke, V.,Geeroms, J., Delcour, J.A., 
Rice starches. II. Structural aspects provide insight into 
swelling and pasting properties. J. Cereal Sci. 2003, 38, 
53–59. 
[118] Tester, R. F., Morrison, W. R., Swelling and gelatinization of 
cereal starches. 1. Effects of amylopectin, amylose and 
lipids. Cereal Chem. 1990, 67, 551–557. 
[119] Ghiasi, K., Hoseney, R. C., Varrianomarston, E., Gelatiniza-tion 
of wheat starch. 1. Excess-water systems. Cereal 
Chem. 1982, 59, 81–85. 
[120] Van Steertegem, B., Pareyt, B., Brijs, K., Delcour, 
J. A., Combined impact of Bacillus stearothermophilus 
maltogenic alpha-amylase and surfactants on starch pasting 
and gelation properties. Food Chem. 2013, 139, 1113– 
1120. 
[121] Roach, R. R., Hoseney, R. C., Effect of certain surfactants 
on the swelling, solubility and amylograph consistency of 
starch. Cereal Chem. 1995, 72, 571–577. 
[122] Takahashi, S., Seib, P. A., Paste and gel properties of prime 
corn and wheat starches with and without native lipids. 
Cereal Chem. 1988, 65, 474–483. 
[123] Doublier, J. L., Llamas, G., Lemeur, M., A rheological 
investigation of cereal starch pastes and gels – effect 
of pasting procedures. Carbohydr. Polym. 1987, 7, 
251–275. 
[124] Srichuwong, S., Sunarti, T. C., Mishima, T., Isono, N., 
Hisamatsu,M., Starches from different botanical sources II: 
Contribution of starch structure to swelling and pasting 
properties. Carbohydr. Polym. 2005, 62, 25–34. 
[125] Zhu, F., Corke, H., Gelatinization, pasting, and gelling 
properties of sweetpotato and wheat starch blends. Cereal 
Chem. 2011, 88, 302–309. 
[126] Putseys, J. A., Gommes, C. J., Van Puyvelde, P., Delcour, 
J. A., Goderis, B., In situ SAXS under shear unveils the 
gelation of aqueous starch suspensions and the impact of 
added amylose–lipid complexes. Carbohydr. Polym. 2011, 
84, 1141–1150. 
[127] Miles, M. J., Morris, V. J., Orford, P. D., Ring, S. G., The 
roles of amylose and amylopectin in the gelation and 
retrogradation of starch. Carbohydr. Res. 1985, 135, 271– 
281. 
[128] Eliasson, A. C., Gudmundsson, M., in: Eliasson, A. C. (Ed.), 
Carbohydrates in food, Marcel Dekker, New York 1996, 
pp. 431–503. 
[129] Slade, L., Levine, H., in: Stivala, S. S., Crescenzi, V., Dea, 
I. C. M. (Eds.), Industrial polysaccharides, the impact of 
biotechnology and advanced methodologies, Gordon  
Breach, New York 1987, pp. 387–430. 
[130] Zeleznak, K. J., Hoseney, R. C., The role of water in the 
retrogradation ofwheat starch gels and bread crumb. Cereal 
Chem. 1986, 63, 407–411. 
[131] Lin, Y. S., Yeh, A. I., Lii, C. Y., Correlation between starch 
retrogradation and water mobility as determined by 
differential scanning calorimetry (DSC) and nuclear magnet-ic 
resonance (NMR). Cereal Chem. 2001, 78, 647–653. 
[132] Kalichevsky, M. T., Orford, P. D., Ring, S. G., The 
retrogradation and gelation of amylopectins from various 
botanical sources. Carbohydr. Res. 1990, 198, 49–55. 
[133] Gudmundsson, M., Eliasson, A. C., Retrogradation of 
amylopectin and the effects of amylose and added 
surfactants/emulsifiers. Carbohydr. Polym. 1990, 13, 
295–315. 
[134] Klucinec, J. D., Thompson, D. B., Amylose and amylopectin 
interact in retrogradation of dispersed high-amylose 
starches. Cereal Chem. 1999, 76, 282–291. 
[135] Parovuori, P.,Manelius, R., Suortti, T., Bertoft, E., Autio, K., 
Effects of enzymically modified amylopectin on the 
rheological properties of amylose–amylopectin mixed gels. 
Food Hydrocolloids 1997, 11, 471–477. 
[136] Jane, J. L., Chen, J. F., Effect of amylosemolecular size and 
amylopectin branch chain length on paste properties of 
starch. Cereal Chem. 1992, 69, 60–65. 
[137] Biliaderis, C. G., in: BeMiller, J., Whistler, R. (Eds.), Starch: 
Chemistry and Technology, Academic Press, New York 
2009, pp. 293–372. 
[138] Dhillon, S., Seetharaman, K., Rheology and texture of starch 
gels containing iodine. J. Cereal Sci. 2011, 54, 374–379. 
[139] Karam, L. B., Grossmann, M. V. E., Silva, R., Ferrero, C., 
Zaritzky, N. E., Gel textural characteristics of corn, cassava 
and yam starch blends: A mixture surface response 
methodology approach. Starch/Stärke 2005, 57, 62–70. 
[140] Vandeputte, G. E., Vermeylen, R., Geeroms, J., Delcour, 
J. A., Rice starches. III. Structural aspects provide insight in 
amylopectin retrogradation properties and gel texture. J. 
Cereal Sci. 2003, 38, 61–68. 
[141] Wang, L. Z., White, P. J., Functional properties of oat 
starches and relationships among functional and structural 
characteristics. Cereal Chem. 1994, 71, 451–458. 
[142] Hibi, Y., Roles of water-soluble and water-insoluble 
carbohydrates in the gelatinization and retrogradation of 
rice starch. Starch/Stärke 1998, 50, 474–478. 
[143] Luallen, T., in: Eliasson, A. C. (Ed.), Starch in Food: 
Structure, Function and Applications, Woodhead Publish-ing, 
Cambridge 2004, pp. 393–424. 
[144] Craig, S. A. S.,Maningat, C. C., Seib, P. A., Hoseney, R. C., 
Starch paste clarity. Cereal Chem. 1989, 66, 173–182. 
[145] Jane, J. L., Kasemsuwan, T., Chen, J. F., Juliano, B. O., 
Phosphorus in rice and other starches. Cereal Foods World 
1996, 41, 827–832. 
[146] Jacobson,M. R., Obanni,M., BeMiller, J. N., Retrogradation 
of starches from different botanical sources. Cereal Chem. 
1997, 74, 511–518. 
[147] Nuwamanya, E., Baguma, Y., Wembabazi, E., Rubaihayo, 
P., A comparative study of the physicochemical properties 
of starches from root, tuber and cereal crops. Afr. J. 
Biotechnol. 2011, 10, 12018–12030. 
 2014 WILEY-VCH Verlag GmbH  Co. KGaA, Weinheim www.starch-journal.com
16 J. Waterschoot et al. Starch/Stärke 2014, 66, 1–16 
[148] Aviara, N. A., Igbeka, J. C., Nwokocha, L. M., Effect of 
drying temperature on physicochemical properties of 
cassava starch. Int. Agrophys. 2010, 24, 219–225. 
[149] Bello-Perez, L. A., Ortiz-Maldonado, F., Villagomez-Mendez, 
J., Toro-Vazquez, J. F., Effect of fatty acids on clarity of 
starch pastes. Starch/Stärke 1998, 50, 383–386. 
[150] Sodhi, N. S., Singh, N., Morphological, thermal and 
rheological properties of starches separated from rice 
cultivars grown in India. Food Chem. 2003, 80, 99–108. 
[151] Goff, H. D., in: Eliasson, A. C. (Ed.) Starch in Food: 
Structure, Function and Applications, Woodhead Publish-ing, 
Cambridge 2004, pp. 425–440. 
[152] Ferrero, C., Zaritzky, N. E., Effect of freezing rate and frozen 
storage on starch–sucrose-hydrocolloid systems. J. Sci. 
Food Agric. 2000, 80, 2149–2158. 
[153] Srichuwong, S., Isono, N., Jiang, H., Mishima, T., 
Hisamatsu, M., Freeze–thaw stability of starches from 
different botanical sources: Correlation with structural 
features. Carbohydr. Polym. 2012, 87, 1275–1279. 
[154] Navarro, A. S., Martino, M. N., Zaritzky, N. E., Viscoelastic 
properties of frozen starch–triglycerides systems. J. Food 
Eng. 1997, 34, 411–427. 
[155] Zheng, G. H., Sosulski, F. W., Determination of water 
separation from cooked starch and flour pastes after 
refrigeration and freeze–thaw. J. Food Sci. 1998, 63, 
134–139. 
[156] Li, L. L., Kim, Y., Huang, W. N., Jia, C. L., Xu, B. C., Effects 
of ice structuring proteins on freeze–thaw stability of corn 
and wheat starch gels. Cereal Chem. 2010, 87, 497–503. 
[157] Jobling, S. A., Westcott, R. J., Tayal, A., Jeffcoat, R., 
Schwall, G. P., Production of a freeze–thaw-stable potato 
starch by antisense inhibition of three starch synthase 
genes. Nat. Biotechnol. 2002, 20, 295–299. 
[158] Charoenrein, S., Tatirat, O., Rengsutthi, K., Thongngam, 
M., Effect of konjac glucomannan on syneresis, textural 
properties and the microstructure of frozen rice starch gels. 
Carbohydr. Polym. 2011, 83, 291–296. 
[159] Arunyanart, T., Charoenrein, S., Effect of sucrose on the 
freeze–thaw stability of rice starch gels: Correlation with 
microstructure and freezable water. Carbohydr. Polym. 
2008, 74, 514–518. 
[160] Yuan, R. C., Thompson, D. B., Freeze–thaw stability of 
three waxymaize starch pastes measured by centrifugation 
and calorimetry. Cereal Chem. 1998, 75, 571–573. 
 2014 WILEY-VCH Verlag GmbH  Co. KGaA, Weinheim www.starch-journal.com

More Related Content

What's hot (20)

Profile on production of baker’s
Profile on production of baker’s Profile on production of baker’s
Profile on production of baker’s
 
whey utilisation
whey utilisationwhey utilisation
whey utilisation
 
Fermentation Process in Yogurt Industry
Fermentation Process in Yogurt IndustryFermentation Process in Yogurt Industry
Fermentation Process in Yogurt Industry
 
Production of food grade yeasts
Production of food grade yeastsProduction of food grade yeasts
Production of food grade yeasts
 
Fermented milk products
Fermented milk productsFermented milk products
Fermented milk products
 
Dairy industry
Dairy industryDairy industry
Dairy industry
 
Evolution in Extrusion of Aquatic Feeds
Evolution in Extrusion of Aquatic FeedsEvolution in Extrusion of Aquatic Feeds
Evolution in Extrusion of Aquatic Feeds
 
Feed Manufacturing Basics
Feed Manufacturing BasicsFeed Manufacturing Basics
Feed Manufacturing Basics
 
PROTEIN ISOLATES
PROTEIN ISOLATESPROTEIN ISOLATES
PROTEIN ISOLATES
 
Industrial Preparation of Yogurt
Industrial Preparation of YogurtIndustrial Preparation of Yogurt
Industrial Preparation of Yogurt
 
YEAST AND YEAST PRODUCESS
YEAST AND YEAST PRODUCESSYEAST AND YEAST PRODUCESS
YEAST AND YEAST PRODUCESS
 
Fermented dairy foods
Fermented     dairy foodsFermented     dairy foods
Fermented dairy foods
 
Useful microorganisms
Useful microorganismsUseful microorganisms
Useful microorganisms
 
Plant proteins ppt
Plant proteins pptPlant proteins ppt
Plant proteins ppt
 
PRESENTATION FINAL
PRESENTATION FINALPRESENTATION FINAL
PRESENTATION FINAL
 
Traditional fermented milk products of india by Geeta Chauhan
Traditional fermented milk products of  india by Geeta ChauhanTraditional fermented milk products of  india by Geeta Chauhan
Traditional fermented milk products of india by Geeta Chauhan
 
Extruded Snack Food
Extruded Snack FoodExtruded Snack Food
Extruded Snack Food
 
Dairy science
Dairy scienceDairy science
Dairy science
 
Waste utilization
Waste utilizationWaste utilization
Waste utilization
 
Fermented milk products
Fermented milk productsFermented milk products
Fermented milk products
 

Viewers also liked

13 instalaciones 167 176
13 instalaciones 167 17613 instalaciones 167 176
13 instalaciones 167 176Edinson Cordova
 
13 instalaciones 167 176
13 instalaciones 167 17613 instalaciones 167 176
13 instalaciones 167 176Edinson Cordova
 
Instalaciones hidráulicas uptc
Instalaciones hidráulicas uptcInstalaciones hidráulicas uptc
Instalaciones hidráulicas uptcRichars Montañezz
 
Instalaciones hidraulicas pvc, cobre, galvanizado
Instalaciones hidraulicas pvc, cobre, galvanizadoInstalaciones hidraulicas pvc, cobre, galvanizado
Instalaciones hidraulicas pvc, cobre, galvanizadoAndy Camarena
 
Vivienda de interes social materiales de construccion
Vivienda de interes social materiales de construccionVivienda de interes social materiales de construccion
Vivienda de interes social materiales de construccionBUSINESS GROUP CORP
 
Instalaciones sanitarias para edificios
Instalaciones sanitarias para edificiosInstalaciones sanitarias para edificios
Instalaciones sanitarias para edificiosWalter Rodriguez
 
Instalaciones hidraulicas y sanitarias en edificios
Instalaciones hidraulicas y sanitarias en edificiosInstalaciones hidraulicas y sanitarias en edificios
Instalaciones hidraulicas y sanitarias en edificiosJeeveth Jackelinne I IT
 
73516654 manual-de-albanileria-las-instalaciones-sanitarias-de-la-casa
73516654 manual-de-albanileria-las-instalaciones-sanitarias-de-la-casa73516654 manual-de-albanileria-las-instalaciones-sanitarias-de-la-casa
73516654 manual-de-albanileria-las-instalaciones-sanitarias-de-la-casaIng. Alberto
 
instalación de agua en una vivienda (fría, caliente y saneamiento)
instalación de agua en una vivienda (fría, caliente y saneamiento)instalación de agua en una vivienda (fría, caliente y saneamiento)
instalación de agua en una vivienda (fría, caliente y saneamiento)Arturo Iglesias Castro
 

Viewers also liked (9)

13 instalaciones 167 176
13 instalaciones 167 17613 instalaciones 167 176
13 instalaciones 167 176
 
13 instalaciones 167 176
13 instalaciones 167 17613 instalaciones 167 176
13 instalaciones 167 176
 
Instalaciones hidráulicas uptc
Instalaciones hidráulicas uptcInstalaciones hidráulicas uptc
Instalaciones hidráulicas uptc
 
Instalaciones hidraulicas pvc, cobre, galvanizado
Instalaciones hidraulicas pvc, cobre, galvanizadoInstalaciones hidraulicas pvc, cobre, galvanizado
Instalaciones hidraulicas pvc, cobre, galvanizado
 
Vivienda de interes social materiales de construccion
Vivienda de interes social materiales de construccionVivienda de interes social materiales de construccion
Vivienda de interes social materiales de construccion
 
Instalaciones sanitarias para edificios
Instalaciones sanitarias para edificiosInstalaciones sanitarias para edificios
Instalaciones sanitarias para edificios
 
Instalaciones hidraulicas y sanitarias en edificios
Instalaciones hidraulicas y sanitarias en edificiosInstalaciones hidraulicas y sanitarias en edificios
Instalaciones hidraulicas y sanitarias en edificios
 
73516654 manual-de-albanileria-las-instalaciones-sanitarias-de-la-casa
73516654 manual-de-albanileria-las-instalaciones-sanitarias-de-la-casa73516654 manual-de-albanileria-las-instalaciones-sanitarias-de-la-casa
73516654 manual-de-albanileria-las-instalaciones-sanitarias-de-la-casa
 
instalación de agua en una vivienda (fría, caliente y saneamiento)
instalación de agua en una vivienda (fría, caliente y saneamiento)instalación de agua en una vivienda (fría, caliente y saneamiento)
instalación de agua en una vivienda (fría, caliente y saneamiento)
 

Similar to Artigo amido revisão

Alfa Laval solutions for processing starch.pdf
Alfa Laval solutions for processing starch.pdfAlfa Laval solutions for processing starch.pdf
Alfa Laval solutions for processing starch.pdfAndrewOseni1
 
Production and DSP of Bakers Yeast
Production and DSP of Bakers YeastProduction and DSP of Bakers Yeast
Production and DSP of Bakers YeastHazem Hussein
 
PRODUCTION OF BAKER’S YEAST.pptx
PRODUCTION OF BAKER’S YEAST.pptxPRODUCTION OF BAKER’S YEAST.pptx
PRODUCTION OF BAKER’S YEAST.pptxMeeraSingh83
 
Probiotic weaning food by monika keshavrao tambakhe
Probiotic weaning food by monika keshavrao tambakheProbiotic weaning food by monika keshavrao tambakhe
Probiotic weaning food by monika keshavrao tambakheMonika Tambakhe
 
Protein Extraction and Purification of Soybean Flakes and Meals Using a Lime ...
Protein Extraction and Purification of Soybean Flakes and Meals Using a Lime ...Protein Extraction and Purification of Soybean Flakes and Meals Using a Lime ...
Protein Extraction and Purification of Soybean Flakes and Meals Using a Lime ...IJMER
 
1.-Status-availability-and-issues-involved-in-the-utilization-of-dairy-by-pro...
1.-Status-availability-and-issues-involved-in-the-utilization-of-dairy-by-pro...1.-Status-availability-and-issues-involved-in-the-utilization-of-dairy-by-pro...
1.-Status-availability-and-issues-involved-in-the-utilization-of-dairy-by-pro...ManteeKumari
 
Single cell protein.
Single cell protein.Single cell protein.
Single cell protein.FSNutri
 
Production of lactic acid from sweet meat industry waste by lactobacillus del...
Production of lactic acid from sweet meat industry waste by lactobacillus del...Production of lactic acid from sweet meat industry waste by lactobacillus del...
Production of lactic acid from sweet meat industry waste by lactobacillus del...eSAT Publishing House
 
Production of lactic acid from sweet meat industry waste by lactobacillus del...
Production of lactic acid from sweet meat industry waste by lactobacillus del...Production of lactic acid from sweet meat industry waste by lactobacillus del...
Production of lactic acid from sweet meat industry waste by lactobacillus del...eSAT Journals
 
Production of Wheat Starch and Gluten
Production of Wheat Starch and GlutenProduction of Wheat Starch and Gluten
Production of Wheat Starch and GlutenAjjay Kumar Gupta
 
Extraction and Modification of Sorghum Starch and its Characterization
Extraction and Modification of Sorghum Starch and its CharacterizationExtraction and Modification of Sorghum Starch and its Characterization
Extraction and Modification of Sorghum Starch and its Characterizationijtsrd
 
Technology of food product bread
Technology of food product bread Technology of food product bread
Technology of food product bread SujataRao11
 
Grain structure of major cereals, pulses and oilseed
Grain structure of  major cereals, pulses and oilseedGrain structure of  major cereals, pulses and oilseed
Grain structure of major cereals, pulses and oilseedpooja1452
 
Starch manufacturing from Corn and Industrial Application
Starch manufacturing from Corn and Industrial ApplicationStarch manufacturing from Corn and Industrial Application
Starch manufacturing from Corn and Industrial ApplicationAbdul Raouf
 
Grain structure of major cereals, pulses and oilseed
Grain structure of  major cereals, pulses and oilseedGrain structure of  major cereals, pulses and oilseed
Grain structure of major cereals, pulses and oilseedpooja1452
 
PROCESS OF EXTRACTION OF LACTOPEROXIDASE PROTEIN FROM YOGURT WHEY
PROCESS OF EXTRACTION OF LACTOPEROXIDASE PROTEIN FROM YOGURT WHEYPROCESS OF EXTRACTION OF LACTOPEROXIDASE PROTEIN FROM YOGURT WHEY
PROCESS OF EXTRACTION OF LACTOPEROXIDASE PROTEIN FROM YOGURT WHEYAin Nur Syazwani
 

Similar to Artigo amido revisão (20)

Alfa Laval solutions for processing starch.pdf
Alfa Laval solutions for processing starch.pdfAlfa Laval solutions for processing starch.pdf
Alfa Laval solutions for processing starch.pdf
 
Production and DSP of Bakers Yeast
Production and DSP of Bakers YeastProduction and DSP of Bakers Yeast
Production and DSP of Bakers Yeast
 
PRODUCTION OF BAKER’S YEAST.pptx
PRODUCTION OF BAKER’S YEAST.pptxPRODUCTION OF BAKER’S YEAST.pptx
PRODUCTION OF BAKER’S YEAST.pptx
 
Probiotic weaning food by monika keshavrao tambakhe
Probiotic weaning food by monika keshavrao tambakheProbiotic weaning food by monika keshavrao tambakhe
Probiotic weaning food by monika keshavrao tambakhe
 
10.11648.j.ijnfs.20150403.19
10.11648.j.ijnfs.20150403.1910.11648.j.ijnfs.20150403.19
10.11648.j.ijnfs.20150403.19
 
Physico-Chemical, Functional and Sensory Properties of Composite Bread prepar...
Physico-Chemical, Functional and Sensory Properties of Composite Bread prepar...Physico-Chemical, Functional and Sensory Properties of Composite Bread prepar...
Physico-Chemical, Functional and Sensory Properties of Composite Bread prepar...
 
Protein Extraction and Purification of Soybean Flakes and Meals Using a Lime ...
Protein Extraction and Purification of Soybean Flakes and Meals Using a Lime ...Protein Extraction and Purification of Soybean Flakes and Meals Using a Lime ...
Protein Extraction and Purification of Soybean Flakes and Meals Using a Lime ...
 
1.-Status-availability-and-issues-involved-in-the-utilization-of-dairy-by-pro...
1.-Status-availability-and-issues-involved-in-the-utilization-of-dairy-by-pro...1.-Status-availability-and-issues-involved-in-the-utilization-of-dairy-by-pro...
1.-Status-availability-and-issues-involved-in-the-utilization-of-dairy-by-pro...
 
Single cell protein.
Single cell protein.Single cell protein.
Single cell protein.
 
Production of lactic acid from sweet meat industry waste by lactobacillus del...
Production of lactic acid from sweet meat industry waste by lactobacillus del...Production of lactic acid from sweet meat industry waste by lactobacillus del...
Production of lactic acid from sweet meat industry waste by lactobacillus del...
 
Production of lactic acid from sweet meat industry waste by lactobacillus del...
Production of lactic acid from sweet meat industry waste by lactobacillus del...Production of lactic acid from sweet meat industry waste by lactobacillus del...
Production of lactic acid from sweet meat industry waste by lactobacillus del...
 
Production of Wheat Starch and Gluten
Production of Wheat Starch and GlutenProduction of Wheat Starch and Gluten
Production of Wheat Starch and Gluten
 
Extraction and Modification of Sorghum Starch and its Characterization
Extraction and Modification of Sorghum Starch and its CharacterizationExtraction and Modification of Sorghum Starch and its Characterization
Extraction and Modification of Sorghum Starch and its Characterization
 
Technology of food product bread
Technology of food product bread Technology of food product bread
Technology of food product bread
 
Dairy industry
Dairy industryDairy industry
Dairy industry
 
Grain structure of major cereals, pulses and oilseed
Grain structure of  major cereals, pulses and oilseedGrain structure of  major cereals, pulses and oilseed
Grain structure of major cereals, pulses and oilseed
 
Starch manufacturing from Corn and Industrial Application
Starch manufacturing from Corn and Industrial ApplicationStarch manufacturing from Corn and Industrial Application
Starch manufacturing from Corn and Industrial Application
 
Dairy waste management
Dairy waste managementDairy waste management
Dairy waste management
 
Grain structure of major cereals, pulses and oilseed
Grain structure of  major cereals, pulses and oilseedGrain structure of  major cereals, pulses and oilseed
Grain structure of major cereals, pulses and oilseed
 
PROCESS OF EXTRACTION OF LACTOPEROXIDASE PROTEIN FROM YOGURT WHEY
PROCESS OF EXTRACTION OF LACTOPEROXIDASE PROTEIN FROM YOGURT WHEYPROCESS OF EXTRACTION OF LACTOPEROXIDASE PROTEIN FROM YOGURT WHEY
PROCESS OF EXTRACTION OF LACTOPEROXIDASE PROTEIN FROM YOGURT WHEY
 

Artigo amido revisão

  • 1. Starch/Stärke 2014, 66, 1–16 DOI 10.1002/star.201300238 1 REVIEW Production, structure, physicochemical and functional properties of maize, cassava, wheat, potato and rice starches Jasmien Waterschoot, Sara V. Gomand, Ellen Fierens and Jan A. Delcour Laboratory of Food Chemistry and Biochemistry, Leuven Food Science and Nutrition Research Centre (LFoRCe), KU Leuven, Leuven, Belgium In 2012, the world production of starch was 75 million tons. Maize, cassava, wheat and potato are the main botanical origins for starch production with only minor quantities of rice and other starches being produced. These starches are either used by industry as such or following some conversion. When selecting and developing starches for specific purposes, it is important to consider the differences between starches of varying botanical origin. Here, an overview is given of the production, structure, composition, morphology, swelling, gelatinisation, pasting and retrogradation, paste firmness and clarity and freeze–thaw stability of maize, cassava, wheat, potato and rice starches. Differences in properties are largely defined by differences in amylose and amylopectin structures and contents, granular organisation, presence of lipids, proteins and minerals and starch granule size. Received: October 3, 2013 Revised: January 21, 2014 Accepted: January 23, 2014 Keywords: Gelatinisation / Production / Retrogradation / Starch / Structure 1 Introduction Starch is an important source of carbohydrates in the human diet. In addition, it is a versatile and widely used additive in the food, paper, chemical and pharmaceutical industries. Worldwide, 75 million tons of starch were produced in 2012 (http://www.zuckerforschung.at/) and marketed as native, physically or chemically modified starch but also as liquid and solid sweeteners. This paper gives an overview of the production, chemical composition, structure and functional properties of maize, cassava, wheat, potato and rice starches. 2 Starch production and uses Of the above mentioned world starch production more than half was produced in the United States. In the European Union, 10 million tons were produced (http://www.zuck-erforschung. at/).World production is estimated to increase to about 85 million tons by 2015. The most important botanical origins for producing starches are maize, cassava, wheat and potato, respectively. Table 1 shows the estimated 2015 production for each starch. Almost 80% of the starch production is from maize. In the USA, mainly maize starch is produced, although (very) small amounts of wheat, potato and rice starches are also manufactured [1]. In Europe, in addition to maize (47%) and wheat starch (39%), also potato starch (14%) and a very small amount of rice starch (<0.5%) are produced [2] (http://www.aaf-eu.org/). Cassava starch is mainly produced in Southeast Asia and Brazil [3]. Only a small fraction (7% for maize, 4% for cassava, 0.9% for wheat and potato and 0.007% for rice) of the raw material crops are used for starch production. In applications, starch is mainly used as starch derived sweeteners and as native and modified starches. In 2011, in the European Union, 57% of the produced starch was converted to sweeteners, 23% was used as native starch and 20% was modified (http://www.aaf-eu.org/). Important starch derived sweeteners are glucose (syrups), (high) fructose (syrups), and the polyols mannitol, sorbitol and maltitol. Maltodextrins and oligosaccharide syrups are also produced [1, 4]. Native starch is used because of its thickening and gelling capacities. However, for a number of applications, properties of native starches fail to meet process or product requirements. This is why starches are also chemically or Correspondence: Jasmien Waterschoot, Laboratory of Food Chemistry and Biochemistry, Leuven Food Science and Nutrition Research Centre (LFoRCe), KU Leuven, Kasteelpark Arenberg 20, B-3001 Leuven, Belgium E-mail: jasmien.waterschoot@biw.kuleuven.be Fax: þ32-16-32-19-97 2014 WILEY-VCH Verlag GmbH Co. KGaA, Weinheim www.starch-journal.com
  • 2. 2 J. Waterschoot et al. Starch/Stärke 2014, 66, 1–16 physically modified. Cross-linking and substitution are common modifications for starches used in food production. Cross-linking of starch improves its acid, heat and shear stability, while the introduction of bulky substituents on the starch chains reduces retrogradation [5]. The EU exports approximately 10% of its produced starch and starch derivatives. Of the remaining 90%, 62% is used by the food industry, 1%is used for feed and 37% is used by the non-food industry (http://www.aaf-eu.org/). Major non-food starch applications are in the paper and board, pharmaceutical (e.g. tablet formulations, encapsulating agents), cosmetics, chem-ical (e.g. adhesives, starch-based plastics) and textile indus-tries. Modifications to meet requirements for non-food applications include oxidation, cationisation, copolymerisa-tion, hydrolysis and substitution [6]. 3 Starch production processes Starch production processes and their associated costs depend on the botanical origin of the starch. Isolation of starch from cassava and potato tubers is relatively simple due to their tissue structure and their relatively low protein and fat contents [7, 8]. Isolation of cereal starches is more difficult as higher levels of these components need to be removed [2, 9, 10]. 3.1 Maize starch Maize starch is commercially isolated by a wet milling procedure. First, contaminating material is removed from the bulk mass, after which the maize is steeped in water containing a low level of sulphur dioxide for typically 24–40 h at 48–52°C to soften the kernels and to obtain optimal milling and separation of the maize components during the wet milling phase. During steeping, the kernels absorb water and sulphur dioxide [10]. The latter induces protein swelling and dispersion because it cleaves inter- and intramolecular disulphide bonds and thus reduces the average MW and increases the solubility of the proteins [11]. During steeping, lactic acid bacteria develop and produce lactic acid from the available sugars. This causes a drop in pH to 4–5 which is optimal for separation of the maize protein from starch. In addition, lactic acid bacteria hydrolyse some highMWsoluble protein [10]. An important side effect of the steeping process is annealing of the starch granules, which changes their structural properties and increases the gelatinisation temper-ature [12]. After steeping, and wet milling, the germ is removed and starch and protein are separated from non-starch polysaccharides by sieving. After dietary fibre removal, starch is separated from protein by centrifugation or sedimentation and flash dried [10]. 3.2 Cassava starch Cassava roots are cleaned, peeled and pulverised into a pulpy slurry. Starch is then isolated at ambient temperature. The roots contain very small levels of protein (1%) and impurities, which can all be removed by decantation. Non-starch polysaccharides are removed by passing the slurry through extractors with coarse and fine screens to remove both large and smaller molecules. The slurry is then dewatered by centrifugation and flash dried. A small level of sulphur dioxide can be added to the process water to control bacterial growth and facilitate the process [13]. 3.3 Wheat starch Wheat is dry milled to separate bran and germ from the endosperm which is recovered as flour. Different processes are used to separate starch and gluten proteins from wheat flour, i.e. dough-ball, batter, dough-batter and high pressure disintegration processes [9]. Van Der Borght et al. [14] extensively reviewed the main processes. In these processes, flour is mixed with different amounts of water to induce gluten agglomeration or even gluten network formation. The dough-ball and dough-batter process are carried out at ambient temperature, while for the other processes warm water (30–50°C) is usually used [9, 14, 15]. After formation of batter or dough, starch and gluten can be separated based on their difference in density (by centrifugation, in hydro-cyclones) or particle size (by sieving) [9, 14, 15]. The starch is then further purified with the use of hydrocyclones or separators and decanters and dried [9]. Table 1. Production of starches of different botanical origins Maize starch Cassava starch Wheat starch Potato starch Rice starch Estimated world starch production 2015 (million tons/year) 64.6 10.2 6.0 3.4 0.05 World production raw material 2011 (million tons/year) (www.faostat.fao.org) 880 250 704 374 723 Main production countries USA, Japan, China, South Korea [3] Thailand, Indonesia, Brazil, China [3] France, Germany, USA, China [3] Netherlands, Germany, France, China [2] Belgium [2], Thailand, Italy 2014 WILEY-VCH Verlag GmbH Co. KGaA, Weinheim www.starch-journal.com
  • 3. Starch/Stärke 2014, 66, 1–16 3 3.4 Potato starch Potatoes are ground to obtain a mixture of starch granules, broken cell walls and the ‘potato juice’, a solution containing proteins, amino acids, sugars and salts. Starch granules and non-starch polysaccharides are separated from the juice by centrifugation. Starch and large non-starch polysaccharides are then separated by sieving. However, some smaller non-starch polysaccharides and some proteins remain present in the starch fraction. The remaining non-starch polysacchar-ides can be removed by centrifugation based on the density difference between these polysaccharides and starch. The soluble protein is removed in a multi-stage, countercurrent flow system. After this, water is removed with rotating vacuum drum filters and flash drying [3]. 3.5 Rice starch Rice starch is traditionally isolated by an alkaline procedure. Broken rice, a by-product of the conversion of brown rice into white rice in a process referred to as milling, is steeped in a 0.3–0.5% sodium hydroxide solution for 12–24 h at 20–50°C. As mentioned above, annealing of the starch granules may occur at the higher process temperatures. In rice, protein and starch are strongly associated. Sodium hydroxide solubilises the rice protein and facilitates the isolation of starch during subsequent wet milling of the kernels. After wet milling, starch is kept in suspension to allow further solubilisation of proteins. Non-starch polysaccharides are removed by filtra-tion and the slurry is washed to remove the proteins, neutralised and dried [2, 16]. 4 Chemical composition of starch granules Starch mainly consists of two polymers of a-D-glucose units linked by a-1,4 and a-1,6 bonds. These are the nearly linear amylose (AM) and highly branched amylopectin (AP). In addition, starch contains minor constituents (lipids, proteins and minerals) of which the levels vary with the botanical origin (Table 2). Tuber and root starches [e.g. potato (0.1%) and cassava starch (0.2%)] usually contain less lipid than cereal starches (0.6–1.4%) [17, 18]. Lipid content is positively correlated with AM content, i.e. most AM free starches Figure 1. Structure of glucose-6-phosphate in an a-(1,4) bound glucose chain. contain negligible lipid levels [18]. Wheat starch contains a high level of LPLs and some glycolipids, whereas lipids of maize and rice starches consist of FFA and LPLs [18]. The endogenous lipids reduce swelling and leaching of carbo-hydrates during heating of starch in excess water by the formation of AM lipid complexes [19–21]. In general, protein (0.1–0.5%) and ash (0.1–0.3%) contents of starch are very low [22–26]. Potato starch contains a relatively high level of phosphorus (0.09%) in the form of phosphate monoesters that are primarily covalently bound to AP [27]. The phosphate is mainly ester linked at the C-6 (61%) (see Fig. 1) and C-3 (38%) positions, with only 1% linked at the C-2 position [28]. The presence of phosphate monoesters in potato starch has large consequences for its swelling behaviour. Negatively charged phosphate groups cause repulsion between adjacent AP chains and allow rapid hydration and large swelling of the granule [27]. Phosphorus in cereal starches (0.01–0.07%) is mainly present in the form of phospholipids [26, 29]. 5 Amylose and amylopectin The AM content of normal starches varies between 14 and 29% [30–36]. Table 3 shows AM contents of potato, cassava, wheat, maize and rice starches. The AM content of rice starches varies from 0 to 40% [30, 37]. Variations in AM content can be produced through cross-breeding, mutagene-sis or transgenic breeding [38]. AM free starches are called ‘waxy’ and exist for maize, cassava, wheat, potato and rice starches [30, 31, 39–45]. Starches with a high AM content (30%) are also available. For maize, starches with an AM content ranging from 50 to 90% are commercially Table 2. Composition of starches of different botanical origins Maize starch Cassava starch Wheat starch Potato starch Rice starch Lipids (%) 0.6–0.8 [18, 26] 0.2 [17] 0.8–1.2 [18] 0.1 [17, 26] 0.6–1.4 [18] Proteins (%) 0.4 [22, 26] 0.3 [23] 0.2–0.3 [22, 24] 0.1 [22, 26] 0.1–0.5 [25] Ash (%) 0.1 [22, 26] 0.3 [23] 0.2 [22] 0.3 [22], 0.2 [26] 0.1 [25] Phosphorus (%) 0.02 [29], 0.01 [26] 0.01 [29] 0.05 [29] 0.09 [29], 0.06 [26] 0.07 [29] Lipid, protein, ash and phosphorus contents are shown as % of total dry weight. 2014 WILEY-VCH Verlag GmbH Co. KGaA, Weinheim www.starch-journal.com
  • 4. 4 J. Waterschoot et al. Starch/Stärke 2014, 66, 1–16 Table 3. Contents and structural properties of amylose (AM) and amylopectin (AP) and degree of crystallinity of starches of different available [44, 46]. Also for potato starch an AM content from 56 to 92% was obtained [31, 47–49], while the highest AM content obtained so far for wheat (74%) [50] and rice (56%) [51] is lower. Mutants with only slightly higher AM contents than those of their regular counterparts have been developed for wheat (31–56%) [41, 47, 52–54], cassava (28– 36%) [39, 55] and potato (27–37%) starches [31]. In essence, AM is a linear polymer. It contains hardly any a-1,6 branch points (1%) [56]. AP is highly branched, with 5–6% a-1,6 linkages [18]. A starch characteristic is its DP, i.e. the number of glucose units in the polysaccharide. Average degrees of polymerisation can be calculated based on the number or on the weight of the molecules. The number average DP (DPn) and the weight average DP (DPw) are given by formulas (1) and (2): DPn ¼ P1 1 DPiðmi=MiÞ P1 1 ðmi=MiÞ ð1Þ DPw ¼ P1 P1 DPimi 1 1 mi ð2Þ withmi the mass concentration andMi theMWof chains with a DP¼i [57]. DPw is always higher than DPn, except for a monodisperse polymer, in which case DPn equals DPw [58]. Table 3 lists DPn of AM and AP of the different starches. DPn ofAMvaries from 570 to 8025 [59–65], while DPn of AP varies from 4700 to 18 000 [62, 63, 66–68]. DPn values of potato and botanical origins Maize starch Cassava starch Wheat starch Potato starch Rice starch AM content (%) 23 [34], 28 [32] 18–24 [33], 18 [31, 34] 25 [35], 26 [32, 34] 19–22 [31], 17 [34], 23 [32], 25–27 [35] 17–21 [32], 21 [34], 14–29 [30], 16–19 [36], 4–16 [37] DPn a) of AM 960 [61], 830 [63] 2660 [59], 3642 [60] 570 [59] 3827 [60], 1290 [61], 830–1570 [62], 1200–1500 [67] 4920 [59], 8025 [60] 920–1110 [64], 847–1118 [65] DPn of AP 5100 [63], 15 900 [66] – 5000–9400 [62], 13 000–18 000 [67] 11 200 [66] 4700–12 800 [68], 8200–10 900 [66] Number of AM molecules per g starch 1017 9–10 2–3 3–11 1–2 6–9 Number of AP molecules per g starch 1017 2–6 – 2–6 3 2–4 CLb) of AM 335 [61], 340 [63] 340 [59] 250–320 [67], 135–255 [62], 270 [61], 300 [59] 670 [59] 230–370 [64] CL of AP 28 [70], 24 [34], 20–21 [63], 20 [72, 75] 26 [70], 28 [34], 18–19 [73], 19 [72] 25 [70], 23 [34], 19–20 [67], 19–21 [62], 19 [72], 23 [69] 34 [70], 29 [34], 31 [71], 23 [72] 25–28 [70], 23 [34], 17–18 [73], 19–22 [68], 18–19 [72] Average number of chains per molecule of AM 2.9 [61], 2.4 [63] 7.8 [59] 4.4–5.2 [67], 5.5–6.5 [62], 4.8 [61], 1.9 [59] 7.3 [59] 2.5–4.3 [64] Average number of chains per molecule of AP 240 [63] – 660–920 (based on [67]) 500 (based on [72] and [66]) 220–700 [74] Degree of crystallinity as determined with X-ray diffraction (%) 27 [86], 40 [85] 24 [86], 38 [85] 20 [86], 36 [85] 24 [86], 28 [85] 38 [85] a) DPn, number degree of polymerisation b) CL, average chain length 2014 WILEY-VCH Verlag GmbH Co. KGaA, Weinheim www.starch-journal.com
  • 5. Starch/Stärke 2014, 66, 1–16 5 cassava AM are higher than those of maize, wheat and rice AM. Based on DPn and the contents of AM and AP, the number of AM and AP molecules per mass unit of starch can be calculated. It can be estimated from data in Table 3 that a given mass of regular starch contains more AM molecules than AP molecules (except for potato starch), which is due to their lower DPn. The average chain length (CL) of AM varies from 135 to 670 glucose units per chain. The CL of potato AM is 670 glucose units while that of the other starches ranges from 135 to 370 glucose units [59, 61–64, 67]. Because of the highly branched nature of AP, its CL (17–34 glucose units) is much smaller than that of AM. Potato and cassava AP have a slightly higher CL than AP of the cereal starches [34, 62, 63, 67–75]. AP of high AM starches has a similar or higher CL than AP of the regular counterparts. In addition, these starches contain some material with DPs and branching patterns intermediate between those of AM and AP [46, 47, 76–79]. The average number of chains per molecule can be calculated from DPn and CL and is much smaller for AM (1.9–7.8) than for AP (220–920) [59, 61–64, 66, 67, 72, 74]. With regard to determination of starch molecular size, it is important to note that the analytical results depend on the procedure used. Molecular size determinations require complete dissolution of the starch polymers, with removal of non-starch components that interfere with the analysis and without starch loss or degradation during the procedure. These requirements are challenging, so when interpreting results, caution should be taken when comparing results obtained by different procedures [80, 81]. 6 Different levels of starch granule organisation Plants synthesise starch in a granular form. Starting from the hilum, starch is deposited in alternating amorphous and semi-crystalline concentric growth rings [16]. The semi-crystalline growth rings consist of amorphous and crystalline lamellae. AP is largely responsible for the crystalline character of starch. Side chains of AP form double helices which are ordered in clusters. The crystalline lamellae contain the double helices, while the amorphous lamellae contain the AP branch points, which connect the double helices [82]. The amorphous growth rings consist of AMand less ordered AP [16]. AM and AP are not present in separate regions, but highly intermingled in the granule [83, 84]. The degree of crystallinity is usually determined with X-ray diffraction [18]. It varies from 20 to 40% depending on the botanical origin (Table 3) [85, 86]. The packing of AP double helices can give rise to different crystal structures or polymorphic forms. Cereal starch crystals are generally packed according to the A-type packing. Such packing is more dense than the B-type packing of e.g. potato starch and high AM maize starches. Cassava starch contains A-type or C-type crystals (a mixture of A- and B-type crystals) [85, 87, 88]. In A-type starches, crystals are packed in a monoclinic unit cell (a¼2.124 nm, b¼1.172 nm, c¼1.069 nm and g¼123.5°) with eight water molecules, whereas in B-type starches, crystals are packed in a hexagonal unit cell (a¼b¼1.85nm and c¼1.04 nm) with 36 water molecules (see Fig. 2). Besides A- and B-type crystals, a third polymorph exists, i.e. V-type crystals. In this polymorph, AM single helices form inclusion complexes with e.g. iodine, alcohols or fatty acids [18]. 7 Morphology of the starch granule The size and shape of starch granules (Table 4) depend on the botanical origin and vary widely. Figure 3 shows the granular morphology of potato, cassava, wheat, maize and rice starches. Potato starch has very large, round or oval granules (10–100mm), while rice starch has very small, polygonal granules (3–8mm) [32, 89, 90]. Cassava starch has round or truncated granules while maize starch granules are polygonal. Both starches have granules with somewhat similar dimensions (5–20mm for maize starch and 3–32 mm for cassava starch) [31–33, 89]. While the shape and size of waxy maize starch granules resemble those of regular maize starch granules, high AM maize starches contain, in addition to the normal polygonal granules, a number of filamentous elongated granules [89, 91]. Wheat starch has a bimodal size distribution, with small, round B granules (2–10mm) and large, lenticular (20–32mm) A granules [32, 89, 92, 93]. Figure 2. Monoclinic unit cell of A-type crystals and hexagonal unit cell of B-type crystals. Projection of the structure is in the ab plane. Reprinted from International Journal of Biological Macromolecules, 23, Buléon, A., Colonna, P., Planchot, V. and Ball, S., Starch granules: structure and biosynthesis, 85–112, Copyright (1998), with permission from Elsevier. 2014 WILEY-VCH Verlag GmbH Co. KGaA, Weinheim www.starch-journal.com
  • 6. 6 J. Waterschoot et al. Starch/Stärke 2014, 66, 1–16 Table 4. Morphological properties of granules of starches of different botanical origins Maize starch Cassava starch Wheat starch Potato starch Rice starch Shape Round, polygonal [32, 89] Round, truncated [31] Round, lenticular [32, 89] Round, oval [89] Polygonal [32, 89] Diameter, range (mm) 5–20 [89] 3–32 [33] A: 20–35 B: 2–10 [32] 10–110 [31] 3–8 [32, 89] Volume mean diameter (mm) 15 [32] 17–18 [31] A: 21–23 B: 6–7 [92, 93] 48–60 [31] 6 [90] Number mean diameter (mm) (own data) 13 11 A: 17 B: 3 29 2 Number of granules per g starch 108 6 10 A: 2 B: 80 0.5 1600 Specific surface area (m2/kg) 270 220 A: 180 B: 670 75 670 Average granule diameters can be calculated based on the number of granules or on their volume (or weight). Number mean diameter (D [1,0]) and volume mean diameter (D [4,3]) can be calculated with formulas (3) and (4), D½1; 0 ¼ P n1 di n ð3Þ D½4; 3 ¼ P n1 d4i P n1 d3i ð4Þ where di is the diameter of particle i and n is the total number of particles. Based on the number mean diameter, the number of granules per gram starch can be determined. This allows estimating that a given weight of rice starch contains about 3000 times more granules than a same given weight of potato starch. The specific surface area, i.e. the total surface area per unit of volume, of rice starch is about ten times that of potato starch. The values for specific surface area in Table 4 are probably an underestimation, as the presence of channels and pores on the granule surface is not included in the calculation. However, to the best of our knowledge, no information is available on this. Especially cereal starches contain pores [94, 95], while for potato and cassava starches also depressions and protrusions on the granule surface have been observed [96]. 8 Gelatinisation properties Heating starch in excess water (1:2 starch:water) above a certain temperature, the ‘gelatinisation temperature’, dis-rupts the molecular order of the granules and melts the crystallites [97]. When relatively less water (1:2 starch:water) is available, gelatinisation is partly postponed to higher temperatures [16]. Table 5 lists gelatinisation onset (To), peak (Tp) and conclusion (Tc) temperatures and melting enthalpies (DH) of the different starches in excess water. DH represents the amount of energy needed to melt all the crystals [16]. Wheat starch has the lowest gelatinisation temperature, followed by potato, cassava and maize starches [22, 25, 26, 30, 31, 34, 71, 98–106]. Rice starches show a high variation in gelatinisation temperature, which is at least partly due to the high variation of AM content in regular rice starches [25, 30, Figure 3. SEM pictures of maize (a), cassava (b), wheat (c), potato (d) and rice (e) starches (own data). Bars represent 10mm. 2014 WILEY-VCH Verlag GmbH Co. KGaA, Weinheim www.starch-journal.com
  • 7. Starch/Stärke 2014, 66, 1–16 7 Table 5. Gelatinisation properties of starches of different botanical origins in excess water measured with DSC Reference Starch-to-water ratio To (°C)a) TP (°C)b) Tc (°C)c) TcTo (°C) DH (J/g)d) Maize starch [22, 26, 34] 1:3 64–67 68–71 72–75 8–11 11–12 [98] 3:7 63 67 72 9 9 [103] 1:4 67 71 – – 12 [104] 1:9 66 71 – – 12 Cassava starch [31, 34] 1:3 55–64 61–68 71–74 10–16 15–19 [98] 3:7 60 65 75 15 11 [103] 1:4 64–66 68–70 – – 11–14 [104] 1:9 65 71 – – 13 Wheat starch [22, 34, 100] 1:3 53–62 61–65 64–69 7–12 9–12 [98, 105, 108] 3:7 48–57 54–62 58–68 9–11 7–15 [103] 1:4 60 65 – – 10 Potato starch [22, 26, 31, 34, 71] 1:3 58–63 61–68 68–73 8–11 15–24 [98, 99, 101, 102] 3:7 57–66 61–70 67–75 7–12 12–18 [103] 1:4 62 65 – – 17 [104] 1:9 59 64 – – 17 Rice starch [34] 1:3 70 76 80 10 13 [37] 3:7 61–76 67–79 72–85 8–13 8–14 [25] 3:11 51–70 58–74 64–79 9–13 8–12 [104] 1:9 58 65 – – 12 [30] 1:2 57–76 63–79 71–83 8–19 17–20 34, 37, 104]. For a single starch granule, loss of molecular order typically takes place over a very small temperature range (1°C) [107], while gelatinisation occurs over a wider interval for a population of granules. The gelatinisation temperature range of starches mostly varies from 8 to 12°C [22, 25, 26, 30, 31, 34, 37, 71, 98–105, 108], although some exceptions with a larger temperature range have also been described [30, 31, 98]. DH of potato starch is slightly higher than that of the other starches. Wheat and maize starches have a low DH [22, 25, 26, 30, 31, 34, 37, 71, 98–105, 108]. The gelatinisation temperature is a measure of crystal quality, while DH is a measure of both crystal quality and quantity [109]. The presence of a relatively high amount of short AP chains (DP14) reduces the gelatinisation temperature, while a relatively high amount of longer chains leads to an increased gelatinisation temperature [30, 31, 105]. According to Gidley and Bulpin [110], a chain DP of at least 10 is needed to form double helices. Consequently, starches with a relatively high level of short AP chains have lower crystalline order, which leads to a lower gelatinisation temperature. Relatively high amounts of longer chains may be responsible for both better stabilisation of the crystal structure over a longer distance as well as for a higher gelatinisation temperature [31]. Although CL of potato AP (23–34) is higher than CL of the other starches (17–28) (Table 3), its gelatinisation temperature is relatively low (To¼57–66°C) (Table 5), probably because of the presence of phosphate monoesters and the more open crystal structure of B-type than of A-type starches [26, 34, 88]. The importance of the AP chain length for gelatinisation temperature is expressed by the Gibbs–Thomson Eq. (5) for lamellar crystallites m 1 2g Tm ¼ T0 DHlc ð5Þ This equation relates the melting temperature (Tm) to the average crystalline layer thickness (lc), the melting tempera-ture of an ideal crystal with an infinite crystal size (T0 m), the lamellar surface free energy (g) and DH. Longer AP chains lead to increased lc and thus also Tm. A high Tm is obtained when branch chain lengths are relatively long (high lc) and when crystal quality is high (low g and high DH) [111]. Branch chain length can also influence DH. Starches with higher CL (e.g. potato starch) appear to have a higher DH[34]. DH is also correlated with crystallinity, i.e. a higher degree of crystallinity (e.g. waxy starches) leads to a higher DH [26, 31, 34, 105]. Furthermore, the presence of lipids in cereal starches might also explain the difference in DH between potato and cereal starches. The exothermic formation of AM lipid complexes can occur simultaneously with gelatinisation, thereby lowering the measured DH [22]. 9 Swelling power and solubility At roomtemperature, starch granules can absorb up to 30% of their weight in excess water without swelling noticeably [112]. a) To, gelatinisation onset temperature; b) Tp, gelatinisation peak temperature, c) Tc, gelatinisation conclusion temperature; d) DH, gelatinisation enthalpy. 2014 WILEY-VCH Verlag GmbH Co. KGaA, Weinheim www.starch-journal.com
  • 8. 8 J. Waterschoot et al. Starch/Stärke 2014, 66, 1–16 Table 6. Swelling power and solubility of starches of different botanical origins Reference Starch concentration (%) Temperature (°C) Swelling power (g/g) Solubility (%) Maize starch [103] 0.5 84 16 – [115] 1.0 95 23 17 [122] 2.0 95 11 18 Cassava starch [103] 0.5 84 40–60 – [114] 0.3 90 50 13 Wheat starch [103] 0.5 84 40 – [105] 2.0 90 13–25 – [122] 2.0 95 9 20 Potato starch [103] 0.5 84 168 – [116] – 90 26–49 3–37 [114] 0.3 90 60–130 12–17 [115] 1.0 95 91 12 Rice starch [117] 1.0 95 25–45 13–32 [37] 2.0 90 17–39 – However, during heating, starch granules absorb much more water and swell. At higher temperatures, part of the polysaccharides go into solution and leach out of the granules [113]. Table 6 lists the swelling powers and solubilities of the different starches. The former are the amounts of water a starch can absorb per gram starch at a certain temperature and at a certain starch concentration, while the solubilities represent the percentages of leached AM and AP at this temperature. Potato starch has a much higher swelling power than other starches [103, 114–116]. As mentioned above, this is largely due to its negatively charged phosphate monoesters. The swelling power of cassava starch is also higher than that of the cereal starches [37, 103, 105, 114, 115, 117]. Granule swelling is mainly attributed to AP and is inhibited by AM [118]. As a result, waxy starches have a higher swelling power than their AM containing counterparts [34, 114, 117, 118]. Removal of endogenous starch lipids reduces formation of AM lipid complexes and increases the swelling power of wheat and maize starches but not up to the level of swelling that characterises tuber starches [21, 118]. The impact of exogenous lipids on swelling of starch granules depends on the type of lipid added and the temperature [19, 119–121]. The absence of AM lipid complexes in potato and cassava starches also contributes to their extensive swelling power [34, 114]. Starch solubility, determined at temperatures from 84 to 95°C, varies from 3 to 37%, with values for most starches between 10 and 20% (Table 6) [103, 105, 114–117, 122]. Both AM and AP leach out of granules of regular starches, while evidently only AP leaches from waxy starches. Usually, in excess water, AM leaching starts at relatively low temper-atures (70°C), while AP only leaches out at higher temperatures (90°C). This has been observed for potato and cassava [114], rice [117], maize [123] and wheat [21, 119, 123] starches. 10 Pasting properties Pasting can be described as a term encompassing the events that occur after gelatinisation in a starch suspension, i.e. further swelling of the granules, leaching of polysaccharides, increase in viscosity and formation of an AM gel network [97]. Changes in viscosity depend on the concentration of the starch suspension and can be measured with a Rapid Visco Analyser or a Viscoamylograph. Figure 4 shows typical pasting profiles for maize, cassava, wheat, potato and rice starches (own data). The pasting temperature is that at which an onset in viscosity rise can be observed. Table 7 lists pasting temperatures, and peak and cold paste viscosities of the different starches. While the term peak viscosity speaks for itself, cold paste viscosity is that measured at the end of the Figure 4. Pasting profiles of 8.0% suspensions of potato, cassava, maize, rice and wheat starches in water measured with a rapid visco analyser. The following temperature-time profile was used at a stirring speed of 160 rpm: 1 min at 50°C, heating from 50 to 95°C in 9 min, 10 min at 95°C, cooling from 95 to 50°C in 15 min, 10 min at 50°C (own data). 2014 WILEY-VCH Verlag GmbH Co. KGaA, Weinheim www.starch-journal.com
  • 9. Starch/Stärke 2014, 66, 1–16 9 heating and cooling cycle. Potato and cassava starches have lower pasting temperatures than cereal starches. Due to the high level of negatively charged phosphate groups in potato starch, viscosity development starts at lower temperatures. Potato starch reaches a very high peak viscosity. The peak viscosities of cereal starches are rather low because of their lower swelling power [34, 37, 105, 116, 124, 125]. Waxy starches reach higher peak viscosities than their AM containing counterparts due to their higher swelling power [117]. It is of note that starches with a high swelling power are more sensitive to granule breakdown at high temperatures than those with a lower swelling power. Pasting profiles of potato starch show a large breakdown during the holding phase at 95°C [114, 124].Waxy starches also have low thermal stability and quickly disintegrate during pasting at high temperatures [22]. During cooling of a starch suspension, viscosity increases as the leached AM forms a gel network [126]. Cold paste viscosities of regular wheat, maize and rice starches are at least as high as their peak viscosities, while cold paste viscosities of potato and cassava starches aremuch lower than their peak viscosities due to significant granule breakdown [34, 105, 116, 124, 125]. After some hours, AM forms stable crystalline structures, which are acid resistant and have a melting temperature of about 150 °C [16, 126]. AMcrystallisation is largely responsible for Table 7. Pasting properties of starches of different botanical origins Reference Starch concentration (%) Pasting temperature (°C) Peak viscosity (mPa s) Cold paste viscosity (mPa s) Maize starch [124] 8 82 2100 2000 [34] 8 82 1800 2000 Cassava starch [124] 8 67 2300 1400 [34] 8 68 2100 1300 Wheat starch [34] 8 89 1250 1850 [125] 9 – 2100 2950 [105] 11 82–90 2250–3400 2600–4000 Potato starch [124] 8 67 9500 3400 [34] 8 64 8400 2750 [116] 11 65–70 4100–7200 2300–3400 Rice starch [124] 8 71 2500 2050 [34] 8 80 1350 1900 [37] 6 72–80 1000–2400 1500–3500 Table 8. Retrogradation properties of starches of different botanical origins measured with DSC Reference Starch-to-water ratio Storage conditions [time (days)/temperature (°C)] DH (J/g)a) Maize starch [34] 1:3 7/4 5.8 [98] 3:7 7/22 3.0 Cassava starch [114] 1:3 28/20 No retrogradation detected [34] 1:3 7/4 3.7 [98] 3:7 7/22 No retrogradation detected Wheat starch [34] 1:3 7/4 3.6 [98] 3:7 7/22 2.0 [105] 3:7 7/4 0.7–3.0 [108] 3:7 28/5 10.1–10.6 Potato starch [114] 1:3 7/20 2.8–7.0 [114] 1:3 28/20 6.3–9.9 [34] 1:3 7/4 7.5 [98] 3:7 7/22 4.2 [99] 3:7 14/4 6.4–8.6 Rice starch [34] 1:3 7/4 5.3 [140] 1:2 28/RTb) 1.4–3.1, 7.8–11.6 [140] 1:2 28/6 5.7–11.1 a) DH, melting enthalpy of retrograded amylopectin; b) RT, room temperature. 2014 WILEY-VCH Verlag GmbH Co. KGaA, Weinheim www.starch-journal.com
  • 10. 10 J. Waterschoot et al. Starch/Stärke 2014, 66, 1–16 the initial firmness of a starch gel, while AP retrogradation (cf. infra) is responsible for the long-term changes in gel firmness [127]. 11 Amylopectin retrogradation Retrogradation is the process of crystallisation of AP molecules in a starch paste [16]. The term retrogradation, which means ‘to go back’, is here only used for AP and not for AM, as strictly speaking only AP molecules can go back to a crystalline entity [16]. AP crystals are acid labile and melt at 50–60°C [87]. Table 8 shows the melting enthalpies of retrograded amylopectin. The extent of retrogradation highly depends on the storage time and temperature, starch concentration and starch structural properties [45, 128]. It consists of three steps: nucleation, propagation and maturation. Nucleation occurs faster at a relatively low temperature, close to the Tg of the starch, while propagation and maturation occur faster at a temperature close to the melting temperature [129]. Besides storage temperature, also the starch-to-water ratio has an important effect on retrogradation. Water content should neither be too high (80%) nor too low (30%) to allow retrogradation [130]. Evidently, AP structure also impacts retrogradation. Short AP chains (DP 6–9) are less susceptible to retrogradation than chains with DP 14–24 [114, 131]. Asmentioned earlier, at least a DP of 10 is needed for forming double helices [110]. This could explain the higher extent of retrogradation of potato starch than of cereal starches under the same conditions (Table 8) [35, 132]. During storage of starch gels, interactions or even co-crystallisation with AM can occur, especially when the AM content is relatively high [133–136]. 12 Starch functionality 12.1 Paste firmness As mentioned above, due to AM crystallisation (short-term) and AP retrogradation (long-term), firmness of starch gels increases with time [115, 127]. This is an important aspect of starch functionality in food products. Gel firmness can be measured by compressing a gel by a certain percentage. It depends on starch concentration, storage time and tempera-ture [137]. Table 9 lists the firmness of gels of maize, cassava, wheat, potato and rice starches. As different starch concen-trations and storage conditions are used, it is difficult to compare results. When comparing firmness values obtained under the same conditions, maize, wheat and potato starches show similar firmness values which are higher than that of a cassava starch gel [138, 139]. This could be due to the relatively low AM content of cassava starch and the large loss of granular integrity in the gel [139]. Rice starch gels have a broad range of firmness values (0.09–7N) [140]. AM content is positively correlated with gel firmness. Waxy starch gels have only a low firmness as a result of their poor gel network formation [140]. In contrast, lipid content is negatively correlated with gel firmness. The formation of AM lipid complexes reduces the amount of AM available for network formation [122, 141, 142]. 12.2 Paste clarity Another important characteristic in many starch applications in food systems is paste clarity. The presence of relatively short chains of AM or AP adds to opacity in food products. While for a range of products including sauces, dressings and puddings this is not a problem, products such as fruit fillings Table 9. Firmness of gels of starches of different botanical origins Reference Starch content (%) Storage conditions [time (days)/temperature (°C)] Compressed proportion of gel (%) Gel firmness (N) Maize starch [138] 8 1/RTa) 40 0.64 [138] 8 7/4 40 1.03 [139] 6 1/4 33 0.93 [141] 6 0.2/25 11 0.04 [141] 6 1/4 11 0.05 Cassava starch [138] 8 1/RTa) 40 Not measurable [138] 8 7/4 40 0.35 [139] 6 1/4 33 0.19 Wheat starch [138] 8 1/RTa) 40 0.69 [138] 8 7/4 40 0.70 [125] 9 1/4 Not reported 0.50 Potato starch [138] 8 1/RTa) 40 0.56 [138] 8 7/4 40 0.71 Rice starch [140] 8 2/6 10 0.09–4.17 [140] 8 14/6 10 0.12–7.05 a) RT, room temperature. 2014 WILEY-VCH Verlag GmbH Co. KGaA, Weinheim www.starch-journal.com
  • 11. Starch/Stärke 2014, 66, 1–16 11 and jellies require the starch pastes to be of high clarity [143]. Paste clarity can be determined by measuring the light transmittance (at 650 nm) of a 1.0% starch paste. Potato starch (42–96% light transmittance) has the highest paste clarity, followed by cassava starch (51–81% light transmittance) and the cereal starches (13–62% light transmittance) [144–148]. The clarity of waxy cereal starch gels is better than that of their AM containing counter-parts [144–146]. The very high paste clarity of potato starch gels is caused by the absence of granule remnants in the gel (cf. high swelling power and high granular break-down) [144]. Pastes of cassava and waxy starches also show few granule remnants. However, more interactions between leached material lead to a higher opacity of these gels. The low clarity of pastes of regular cereal starches is caused by swollen granule remnants [144] and by the presence of AM lipid complexes [149]. During storage of starch gels, their paste clarity decreases as more and more association of AM and/or AP molecules takes place [146, 150]. 12.3 Freeze–thaw stability Freeze–thaw stability is an important quality aspect of starch gels. When a starch gel undergoes repeated freezing and thawing cycles, it releases water. The gel is then said to undergo syneresis. The extent of syneresis is a measure for its freeze–thaw stability [128, 151]. During freezing, phase separation takes place as ice crystals are formed. As a result, starch is concentrated in the unfrozen matrix. During thawing of the gel, association of AM and AP takes place in the more concentrated zones. As a result, increasingly insoluble aggregates are formed. Ice crystals melt and the water is not reabsorbed by the starch gel. A sponge like structure develops and the melted water readily separates from the gel [152]. Different factors have an influence on the freeze–thaw stability. These include the freezing rate, the botanical origin of the starch and its AM content, the starch concentration, the number of freeze–thaw cycles and the sample prepara-tion [128, 153]. The freezing rate is an important factor affecting syneresis. Slow freezing favours formation of large ice crystals which disrupt the gel structure to a larger extent than small ice crystals [128, 152, 154]. Slow freezing also leads to maximal ice formation and AM self-association which result in significant structure loss and syneresis. The latter is positively correlated with AM content [153]. Native, AM containing starches have very low freeze–thaw stability. Syneresis readings after several freeze–thaw cycles for maize starch (47–79%) [153, 155, 156], cassava starch (50–67%) [153, 155], wheat starch (44–68%) [153, 155, 156] and potato starch (60–76%) [153, 155, 157] are very high. Rice starch shows a broad range of syneresis values (7–75%) [153, 155, 158, 159]. Waxy starches are intrinsically more stable than AM containing starches, although they can also undergo strong textural changes during freeze–thaw cycles [151, 155]. Waxy rice starch has a very good freeze–thaw stability. A relatively high proportion of AP side chains with DP 6–12, which is the case for waxy rice starch, is believed to be at the basis of a reduced extent of syneresis [153, 157]. In contrast, syneresis of waxy maize starch gels is comparable to that of gels from regular maize starch [153, 155]. For different waxy maize starches, no correlation was observed between syneresis and retrogradation enthalpy as measured with DSC. Syneresis values were already maximal, when little, if any, double helical order was present. This indicates that the early stages of AM and/or AP association (before the formation of real crystal structures) already cause syneresis due to network formation [160]. The presence of lipids in cereal starches may also contribute to a relatively high freeze–thaw stability. Granular swelling and leaching of AM are reduced in the presence of lipids. As a result, starch molecules remain close to each other in the granules. This facilitates their reassociation and in this way it may contribute to a low freeze–thaw stability [153]. In conclusion, there are significant differences in structural, physicochemical and functional properties of starches of different botanical origins. Starch is widely used by the food and non-food industry in a broad range of products. The overview of starch properties provided in this review can be of assistance when developing starches for specific purposes. The botanical origin of the used starch has a great impact on final product properties. Within a botanical origin, in some instances the starch AM content is a major determinant of its properties and starch should therefore be selected with great care. The authors gratefully acknowledge Flanders’ FOOD (Brus-sels, Belgium) and the Methusalem programme ‘Food for the future’ (KU Leuven) for financial support. J.A.D. is W.K. Kellogg Chair in Cereal Science and Nutrition at the KU Leuven. The authors have declared no conflict of interest. 13 References [1] Schwartz, D., Whistler, R., in: BeMiller, J., Whistler, R. (Eds.), Starch: Chemistry and Technology,Academic Press, New York 2009, pp. 1–10. [2] Mitchell, C. R., in: BeMiller, J., Whistler, R. (Eds.), Starch: Chemistry and Technology, Academic Press, New York 2009, pp. 569–578. [3] Grommers, H. E., van der Krogt, D. A., in: BeMiller, J., Whistler, R. (Eds.), Starch: Chemistry and Technology, Academic Press, New York 2009, pp. 511–539. [4] Hobbs, L., in: BeMiller, J., Whistler, R. (Eds.), Starch: Chemistry and Technology, Academic Press, New York 2009, pp. 797–832. 2014 WILEY-VCH Verlag GmbH Co. KGaA, Weinheim www.starch-journal.com
  • 12. 12 J. Waterschoot et al. Starch/Stärke 2014, 66, 1–16 [5] Mason, W. R., in: BeMiller, J., Whistler, R. (Eds.), Starch: Chemistry and Technology, Academic Press, New York 2009, pp. 745–795. [6] Chiu, C., Solarek, D., in: BeMiller, J., Whistler, R. (Eds.), Starch: Chemistry and Technology, Academic Press, New York 2009, pp. 629–655. [7] Moorthy, S. N., in: Eliasson, A. C. (Ed.), Starch in food: Structure, Function and Applications, Woodhead Publish-ing, Cambridge 2004, pp. 321–359. [8] Bergthaller, W., in: Eliasson, A. C. (Ed.), Starch in food: Structure, Function and Applications, Woodhead Publish-ing, Cambridge 2004, pp. 241–257. [9] Maningat, C. C., Seib, P. A., Bassi, S. D., Woo, K. S., Lasater, G. D., in: BeMiller, J., Whistler, R. (Eds.), Starch: Chemistry and Technology, Academic Press, New York 2009, pp. 441–510. [10] Eckhoff, S. R., Watson, S. A., in: BeMiller, J., Whistler, R. (Eds.), Starch: Chemistry and Technology,Academic Press, New York 2009, pp. 373–439. [11] Boundy, J. A., Turner, J. E., Wall, J. S., Dimler, R. J., Influence of commercial processing on composition and properties of corn zein. Cereal Chem. 1967, 44, 281–287. [12] Jacobs, H., Delcour, J. A., Hydrothermal modifications of granular starch, with retention of the granular structure: A review. J. Agric. Food Chem. 1998, 46, 2895–2905. [13] Breuninger, W. F., Piyachomkwan, K., Sriroth, K., in: BeMiller, J., Whistler, R. (Eds.), Starch: Chemistry and Technology, Academic Press, New York 2009, pp. 541– 568. [14] Van Der Borght, A., Goesaert, H., Veraverbeke, W. S., Delcour, J. A., Fractionation of wheat and wheat flour into starch and gluten: Overview of the main processes and the factors involved. J. Cereal Sci. 2005, 41, 221–237. [15] Cornell, H., in: Eliasson, A. C. (Ed.), Starch in Food: Structure, Function and Applications, Woodhead Publish-ing, Cambridge 2004, pp. 211–240. [16] Delcour, J. A., Hoseney, R. C., Principles of Cereal Science and Technology, AACC International, Inc., St. Paul, MN 2010. [17] Vasanthan, T., Hoover, R., A comparative study of the composition of lipids associated with starch granules from various botanical sources. Food Chem. 1992, 43, 19–27. [18] Buléon, A., Colonna, P., Planchot, V., Ball, S., Starch granules: Structure and biosynthesis. Int. J. Biol.Macromol. 1998, 23, 85–112. [19] Raphaelides, S. N., Georgiadis, N., Effect of fatty acids on the rheological behaviour of maize starch dispersions during heating. Carbohydr. Polym. 2006, 65, 81–92. [20] Putseys, J. A., Lamberts, L., Delcour, J. A., Amylose-inclusion complexes: Formation, identity and physico-chemical properties. J. Cereal Sci. 2010, 51, 238–247. [21] Shi, Y.C., Seib, P.A., Lu, S.P.W., in: Levine, H., Slade, L., (Eds.), Water relationships in foods: Advances in the 1980s and trends for the 1990s, Plenum Press, New York 1991, pp. 667–686. [22] Schirmer,M., Höchstötter,A., Jekle, M.,Arendt, E., Becker, T., Physicochemical and morphological characterization of different starches with variable amylose/amylopectin ratio. Food Hydrocolloids 2013, 32, 52–63. [23] Leelavathi, K., Indrani, D., Sidhu, J. S., Amylograph pasting behaviour of cereal and tuber starches. Starch/Stärke 1987, 39, 378–381. [24] Skerritt, J. H., Frend, A. J., Robson, L. G., Greenwell, P., Immunological homologies between wheat gluten and starch granule proteins. J. Cereal Sci. 1990, 12, 123–136. [25] Hoover, R., Sailaja, Y., Sosulski, F. W., Characterization of starches fromwild and long grain brown rice. Food Res. Int. 1996, 29, 99–107. [26] Dhital, S., Shrestha, A. K., Hasjim, J., Gidley, M. J., Physicochemical and structural properties of maize and potato starches as a function of granule size. J. Agric. Food Chem. 2011, 59, 10151–10161. [27] Singh, N., Singh, J., Kaur, L., Sodhi, N. S., Gill, B. S., Morphological, thermal and rheological properties of starches from different botanical sources. Food Chem. 2003, 81, 219–231. [28] Tabata, S., Hizukuri, S., Studies on starch phosphate. Part 2. Isolation of glucose 3-phosphate and maltose phosphate by acid hydrolysis of potato starch. Starch/Stärke 1971, 23, 267–272. [29] Lim, S. T., Kasemsuwan, T., Jane, J. L., Characterization of phosphorus in starch by 31P-nuclear magnetic resonance spectroscopy. Cereal Chem. 1994, 71, 488–493. [30] Vandeputte, G. E., Vermeylen, R., Geeroms, J., Delcour, J. A., Rice starches, I., Structural aspects provide insight into crystallinity characteristics and gelatinisation behaviour of granular starch. J. Cereal Sci. 2003, 38, 43–52. [31] Gomand, S. V., Lamberts, L., Derde, L. J., Goesaert, H. et al., Structural properties and gelatinisation characteristics of potato and cassava starches and mutants thereof. Food Hydrocolloids 2010, 24, 307–317. [32] Lineback, D. R., The starch granule – organization and properties. Bakers Dig. 1984, 58, 16–21. [33] Defloor, I., Dehing, I., Delcour, J. A., Physico-chemical properties of cassava starch. Starch/Stärke 1998, 50, 58–64. [34] Jane, J. L., Chen, Y. Y., Lee, L. F., McPherson, A. E. et al., Effects of amylopectin branch chain length and amylose content on the gelatinization and pasting properties of starch. Cereal Chem. 1999, 76, 629–637. [35] Fredriksson, H., Silverio, J., Andersson, R., Eliasson, A. C., Aman, P., The influence of amylose and amylopectin characteristics on gelatinization and retrogradation proper-ties of different starches. Carbohydr. Polym. 1998, 35, 119–134. [36] Juliano, B. O., Villareal, R. M., Perez, C. M., Villareal, C. P. et al., Varietal differences in properties among high amylose rice starches. Starch/Stärke 1987, 39, 390–393. [37] Singh, N., Kaur, L., Sandhu, K. S., Kaur, J., Nishinari, K., Relationships between physicochemical, morphological, thermal, rheological properties of rice starches. Food Hydrocolloids 2006, 20, 532–542. [38] Lafiandra, D., Sestili, F., D’Ovidio, R., Janni, M. et al., Approaches formodification of starch composition in durum wheat. Cereal Chem. 2010, 87, 28–34. [39] Rolland-Sabaté, A., Sánchez, T., Buléon, A., Colonna, P. et al., Structural characterization of novel cassava starches with low and high-amylose contents in comparison with other commercial sources. Food Hydrocolloids 2012, 27, 161–174. [40] Raemakers, K., Schreuder,M., Suurs, L., Furrer-Verhorst,H. et al., Improved cassava starch by antisense inhibition of granule-bound starch synthase I. Mol. Breed. 2005, 16, 163–172. 2014 WILEY-VCH Verlag GmbH Co. KGaA, Weinheim www.starch-journal.com
  • 13. Starch/Stärke 2014, 66, 1–16 13 [41] Hung, P. V.,Maeda, T.,Morita, N., Study on physicochemi-cal characteristics of waxy and high-amylose wheat starches in comparison with normal wheat starch. Starch/Stärke 2007, 59, 125–131. [42] Wang, Y. J., Wang, L. F., Structures of four waxy rice starches in relation to thermal, pasting, and textural properties. Cereal Chem. 2002, 79, 252–256. [43] Sanchez, T., Dufour, D., Moreno, I. X., Ceballos, H., Comparison of pasting and gel stabilities of waxy and normal starches from potato, maize, and rice with those of a novel waxy cassava starch under thermal, chemical, and mechanical stress. J. Agric. Food Chem. 2010, 58, 5093– 5099. [44] Matveev, Y. I., van Soest, J. J. G., Nieman, C., Wasserman, L. A. et al., The relationship between thermodynamic and structural properties of low and high amylose maize starches. Carbohydr. Polym. 2001, 44, 151–160. [45] Yuan,R.C.,Thompson,D. B.,Boyer,C.D., Finestructureof amylopectin in relation to gelatinization and retrogradation behavior of maize starches from three wx-containing genotypes in two inbred lines. Cereal Chem. 1993, 70, 81–89. [46] Shi, Y. C., Capitani, T., Trzasko, P., Jeffcoat, R., Molecular structure of a low-amylopectin starch and other high-amylose maize starches. J. Cereal Sci. 1998, 27, 289–299. [47] Schwall, G. P., Safford, R., Westcott, R. J., Jeffcoat, R. et al., Production of very-high-amylose potato starch by inhibition of SBE A and B. Nat. Biotechnol. 2000, 18, 551– 554. [48] Karlsson, M. E., Leeman,A.M., Bjorck, I. M. E., Eliasson,A.- C., Some physical and nutritional characteristics of genetically modified potatoes varying in amylose/amylo-pectin ratios. Food Chem. 2007, 100, 136–146. [49] Uwer, U., Frohberg, C., Pilling, J., Landschütze, V., US Patent 7112718, 2006. [50] Regina, A., Bird, A., Topping, D., Bowden, S. et al., High-amylose wheat generated by RNA interference improves indices of large-bowel health in rats. Proc. Natl. Acad. Sci. USA 2006, 103, 3546–3551. [51] Zhu, L., Gu, M., Meng, X., Cheung, S. C. K. et al., High-amylose rice improves indices of animal health in normal and diabetic rats. Plant Biotechnol. J. 2012, 10, 353–362. [52] Yamamori,M., Fujita, S., Hayakawa, K.,Matsuki, J., Yasui, T., Genetic elimination of a starch granule protein, SGP-1, of wheat generates an altered starch with apparent high amylose. Theor. Appl. Genet. 2000, 101, 21–29. [53] Hogg, A. C., Gause, K., Hofer, P., Martin, J. M. et al., Creation of a high-amylose durum wheat through mutagen-esis of starch synthase II (SSIIa). J. Cereal Sci. 2013, 57, 377–383. [54] Konik-Rose, C., Thistleton, J., Chanvrier, H., Tan, I. et al., Effects of starch synthase IIa gene dosage on grain, protein and starch in endosperm of wheat. Theor. Appl. Genet. 2007, 115, 1053–1065. [55] Ceballos, H., Sanchez, T., Denyer, K., Tofino, A. P. et al., Induction and identification of a small-granule, high-amylose mutant in cassava (Manihot esculenta Crantz). J. Agric. Food Chem. 2008, 56, 7215–7222. [56] Ball, S., Guan, H. P., James, M., Myers, A. et al., From glycogen to amylopectin: A model for the biogenesis of the plant starch granule. Cell 1996, 86, 349–352. [57] Gelders, G. G., Vanderstukken, T. C., Goesaert, H., Delcour, J. A., Amylose–lipid complexation: A new fractionation method. Carbohydr. Polym. 2004, 56, 447–458. [58] Billmeyer, F. W. J., Textbook of Polymer Science, John Wiley Sons, Inc., New York 1984. [59] Takeda, Y., Shirasaka, K., Hizukuri, S., Examination of the purity and structure of amylose by gel-permeation chroma-tography. Carbohydr. Res. 1984, 132, 83–92. [60] Ong,M.H.,Jumel, K.,Tokarczuk,P.F., Blanshard, J.M. V., Harding, S. E., Simultaneous determinations of the molecular weight distributions of amyloses and the fine structures of amylopectins of native starches. Carbohydr. Res. 1994, 260, 99–117. [61] Takeda, Y., Hizukuri, S., Takeda, C., Suzuki, A., Structures of branched molecules of amyloses of various origins and molar fractions of branched and unbranched molecules. Carbohydr. Res. 1987, 165, 139–145. [62] Shibanuma, K., Takeda, Y., Hizukuri, S., Shibata, S., Molecular structures of some wheat starches. Carbohydr. Polym. 1994, 25, 111–116. [63] Takeda, Y., Preiss, J., Structures of B90 (sugary) andW64a (normal) maize starches. Carbohydr. Res. 1993, 240, 265–275. [64] Takeda, Y., Hizukuri, S., Juliano, B. O., Purification and structure of amylose from rice starch. Carbohydr. Res. 1986, 148, 299–308. [65] Chen, M. H., Bergman, C. J., Method for determining the amylose content, molecular weights, and weight- and molar-based distributions of degree of polymerization of amylose and fine-structure of amylopectin. Carbohydr. Polym. 2007, 69, 562–578. [66] Takeda, Y., Shibahara, S., Hanashiro, I., Examination of the structure of amylopectin molecules by fluorescent labeling. Carbohydr. Res. 2003, 338, 471–475. [67] Shibanuma, Y., Takeda, Y., Hizukuri, S., Molecular and pasting properties of some wheat starches. Carbohydr. Polym. 1996, 29, 253–261. [68] Takeda, Y., Hizukuri, S., Juliano, B. O., Structures of rice amylopectins with low and high affinities for iodine. Carbohydr. Res. 1987, 168, 79–88. [69] Gomand, S. V., Verwimp, T., Goesaert, H., Delcour, J. A., Structural and physicochemical characterisation of rye starch. Carbohydr. Res. 2011, 346, 2727–2735. [70] Hizukuri, S., Relationship between the distribution of the chain length of amylopectin and the crystalline structure of starch granules. Carbohydr. Res. 1985, 141, 295–306. [71] Yoo, S. H., Perera, C., Shen, J., Ye, L. et al., Molecular structure of selected tuber and root starches and effect of amylopectin structure on their physical properties. J. Agric. Food Chem. 2009, 57, 1556–1564. [72] Hizukuri, S., Kaneko, T., Takeda, Y., Measurement of the chain length of amylopectin and its relevance to the origin of crystalline polymorphism of starch granules. Biochim. Biophys. Acta 1983, 760, 188–191. [73] Laohaphatanaleart, K., Piyachomkwan, K., Sriroth, K., Santisopasri, V., Bertoft, E.,Astudy of the internal structure in cassava and rice amylopectin. Starch/Stärke 2009, 61, 557–569. [74] Hizukuri, S., Takeda, Y., Maruta, N., Juliano, B. O., Molecular structures of rice starch. Carbohydr. Res. 1989, 189, 227–235. 2014 WILEY-VCH Verlag GmbH Co. KGaA, Weinheim www.starch-journal.com
  • 14. 14 J. Waterschoot et al. Starch/Stärke 2014, 66, 1–16 [75] Zhu, F., Bertoft, E., Kallman, A., Myers, A. M., Seethara-man, K., Molecular structure of starches from maize mutants deficient in starch synthase III. J. Agric. Food Chem. 2013, 61, 9899–9907. [76] Jobling, S. A., Schwall, G. P.,Westcott, R. J., Sidebottom, C. M. et al., A minor form of starch branching enzyme in potato (SolanumtuberosumL.) tubers has a major effect on starch structure: Cloning and characterisation of multiple forms of SBE A. Plant J. 1999, 18, 163–171. [77] Baba, T., Arai, Y., Structural features of amylomaize starch. 3. Structural characterization of amylopectin and intermedi-ate material in amylomaize starch granules. Agric. Biol. Chem. 1984, 48, 1763–1775. [78] Takeda, C., Takeda, Y., Hizukuri, S., Structure of the amylopectin fraction of amylomaize. Carbohydr. Res. 1993, 246, 273–281. [79] Song, Y., Jane, J., Characterization of barley starches of waxy, normal, and high amylose varieties. Carbohydr. Polym. 2000, 41, 365–377. [80] Vilaplana, F., Gilbert, R. G., Characterization of branched polysaccharides using multiple-detection size separation techniques. J. Sep. Sci. 2010, 33, 3537–3554. [81] Cave, R. A., Seabrook, S. A., Gidley, M. J., Gilbert, R. G., Characterization of starch by size-exclusion chromatogra-phy: The limitations imposed by shear scission. Biomacro-molecules 2009, 10, 2245–2253. [82] Thompson, D. B., On the non-random nature of amylopectin branching. Carbohydr. Polym. 2000, 43, 223–239. [83] Jane, J. L., Xu, A., Radosavljevic, M., Seib, P. A., Location of amylose in normal starch granules. I. Susceptibility of amylose and amylopectin to cross-linking reagents. Cereal Chem. 1992, 69, 405–409. [84] Kasemsuwan, T., Jane, J., Location of amylose in normal starch granules. II. Locations of phosphodiester cross-linking revealed by Phosphorus-31 nuclear magnetic resonance. Cereal Chem. 1994, 71, 282–287. [85] Zobel, H. F.,Molecules to granules:Acomprehensive starch review. Starch/Stärke 1988, 40, 44–50. [86] Cooke, D., Gidley, M. J., Loss of crystalline and molecular order during starch gelatinization – Origin of the enthalpic transition. Carbohydr. Res. 1992, 227, 103–112. [87] Colonna, P., Buléon, A., 9th International Cereal and Bread Congress, ICC, Paris 1992. [88] Vermeylen, R., Goderis, B., Reynaers, H., Delcour, J. A., Amylopectin molecular structure reflected in macromolecu-lar organization of granular starch. Biomacromolecules 2004, 5, 1775–1786. [89] Jane, J. L., Kasemsuwan, T., Leas, S., Zobel, H., Robyt, J. F., Anthology of starch granule morphology by scanning electron microscopy. Starch/Stärke 1994, 46, 121–129. [90] Jacquier, J. C., Kar, A., Lyng, J. G., Morgan, D. J., McKenna, B. M., Influence of granule size on the flow behaviour of heated rice starch dispersions in excesswater. Carbohydr. Polym. 2006, 66, 425–434. [91] Jiang, H., Horner, H. T., Pepper, T. M., Blanco, M. et al., Formation of elongated starch granules in high-amylose maize. Carbohydr. Polym. 2010, 80, 533–538. [92] Kim, H. S., Huber, K. C., Physicochemical properties and amylopectin fine structures of A- and B-type granules of waxy and normal soft wheat starch. J. Cereal Sci. 2010, 51, 256–264. [93] Kim, H. S., Huber, K. C., Channels within soft wheat starch A- and B-type granules. J. Cereal Sci. 2008, 48, 159–172. [94] Fannon, J. E.,Hauber, R. J., Bemiller, J. N., Surface pores of starch granules. Cereal Chem. 1992, 69, 284–288. [95] Juszczak, L., Fortuna, T., Krok, F., Non-contact atomic force microscopy of starch granules surface. Part II. Selected cereal starches. Starch/Stärke 2003, 55, 8–16. [96] Juszczak, L., Fortuna, T., Krok, F., Non-contact atomic force microscopy of starch granules surface. Part I. Potato and tapioca starches. Starch/Stärke 2003, 55, 1–7. [97] Atwell, W. A., Hood, L. F., Lineback, D. R., Varriano- Marston, E., Zobel, H. F., The terminology andmethodology associated with basic starch phenomena. Cereal Foods World 1988, 33, 306–311. [98] Teng, L. Y., Chin, N. L., Yusof, Y. A., Rheological and textural studies of fresh and freeze-thawed native sago starch-sugar gels. II. Comparisons with other starch sources and reheating effects. Food Hydrocolloids 2013, 31, 156–165. [99] Singh, J., Singh, N., Studies on the morphological, thermal and rheological properties of starch separated from some Indian potato cultivars. Food Chem. 2001, 75, 67–77. [100] Ao, Z., Jane, J. L., Characterization and modeling of the A-and B-granule starches of wheat, triticale, and barley. Carbohydr. Polym. 2007, 67, 46–55. [101] Kim, Y. S., Wiesenborn, D. P., Orr, P. H., Grant, L. A., Screening potato starch for novel properties using differen-tial scanning calorimetry. J. Food Sci. 1995, 60, 1060– 1065. [102] Kaur, L., Singh, N., Sodhi, N. S., Some properties of potatoes and their starches. II. Morphological, thermal and rheological properties of starches. Food Chem. 2002, 79, 183–192. [103] Anggraini, V., Sudarmonowati, E., Hartati, N. S., Suurs, L., Visser, R. G. F., Characterization of cassava starch attributes of different genotypes. Starch/Stärke 2009, 61, 472–481. [104] Li, J. Y., Yeh, A. I., Relationships between thermal, rheological characteristics and swelling power for various starches. J. Food Eng. 2001, 50, 141–148. [105] Singh, S., Singh, N., Isono, N., Noda, T., Relationship of granule size distribution and amylopectin structure with pasting, thermal, and retrogradation properties in wheat starch. J. Agric. Food Chem. 2010, 58, 1180–1188. [106] Sasaki, T., Yasui, T.,Matsuki, J., Effect of amylose content on gelatinization, retrogradation, and pasting properties of starches from waxy and nonwaxy wheat and their F1 seeds. Cereal Chem. 2000, 77, 58–63. [107] Liu, H., Lelièvre, J., Ayoungchee, W., A study of starch gelatinization using differential scanning calorimetry, X-ray and birefringence measurements. Carbohydr. Res. 1991, 210, 79–87. [108] Sasaki, T., Yasui, T., Matsuki, J., Satake, T., Rheological properties of mixed gels using waxy and non-waxy wheat starch. Starch/Stärke 2002, 54, 410–414. [109] Tester, R. F., Morrison, W. R., Swelling and gelatinization of cereal starches. 2.Waxy rice starches. Cereal Chem. 1990, 67, 558–563. [110] Gidley, M. J., Bulpin, P. V., Crystallisation of malto-oligosaccharides as models of the crystalline forms of starch:Minimum chain-length requirement for the formation of double helices. Carbohydr. Res. 1987, 161, 291–300. 2014 WILEY-VCH Verlag GmbH Co. KGaA, Weinheim www.starch-journal.com
  • 15. Starch/Stärke 2014, 66, 1–16 15 [111] Gomand, S. V., Lamberts, L., Gommes, C. J., Visser, R. G. F. et al., Molecular and morphological aspects of annealing-induced stabilization of starch crystallites. Biomacromole-cules 2012, 13, 1361–1370. [112] Kerr, R. W., Chemistry and Industry of Starch, Academic Press, New York 1950. [113] Hermansson, A. M., Svegmark, K., Developments in the understanding of starch functionality. Trends Food Sci. Technol. 1996, 7, 345–353. [114] Gomand, S. V., Lamberts, L., Visser, R. G. F.,Delcour, J.A., Physicochemical properties of potato and cassava starches and their mutants in relation to their structural properties. Food Hydrocolloids 2010, 24, 424–433. [115] Ring, S. G., Some studies on starch gelation. Starch/Stärke 1985, 37, 80–83. [116] Kaur, A., Singh, N., Ezekiel, R., Guraya, H. S., Physico-chemical, thermal and pasting properties of starches separated from different potato cultivars grown at different locations. Food Chem. 2007, 101, 643–651. [117] Vandeputte,G. E., Derycke, V.,Geeroms, J., Delcour, J.A., Rice starches. II. Structural aspects provide insight into swelling and pasting properties. J. Cereal Sci. 2003, 38, 53–59. [118] Tester, R. F., Morrison, W. R., Swelling and gelatinization of cereal starches. 1. Effects of amylopectin, amylose and lipids. Cereal Chem. 1990, 67, 551–557. [119] Ghiasi, K., Hoseney, R. C., Varrianomarston, E., Gelatiniza-tion of wheat starch. 1. Excess-water systems. Cereal Chem. 1982, 59, 81–85. [120] Van Steertegem, B., Pareyt, B., Brijs, K., Delcour, J. A., Combined impact of Bacillus stearothermophilus maltogenic alpha-amylase and surfactants on starch pasting and gelation properties. Food Chem. 2013, 139, 1113– 1120. [121] Roach, R. R., Hoseney, R. C., Effect of certain surfactants on the swelling, solubility and amylograph consistency of starch. Cereal Chem. 1995, 72, 571–577. [122] Takahashi, S., Seib, P. A., Paste and gel properties of prime corn and wheat starches with and without native lipids. Cereal Chem. 1988, 65, 474–483. [123] Doublier, J. L., Llamas, G., Lemeur, M., A rheological investigation of cereal starch pastes and gels – effect of pasting procedures. Carbohydr. Polym. 1987, 7, 251–275. [124] Srichuwong, S., Sunarti, T. C., Mishima, T., Isono, N., Hisamatsu,M., Starches from different botanical sources II: Contribution of starch structure to swelling and pasting properties. Carbohydr. Polym. 2005, 62, 25–34. [125] Zhu, F., Corke, H., Gelatinization, pasting, and gelling properties of sweetpotato and wheat starch blends. Cereal Chem. 2011, 88, 302–309. [126] Putseys, J. A., Gommes, C. J., Van Puyvelde, P., Delcour, J. A., Goderis, B., In situ SAXS under shear unveils the gelation of aqueous starch suspensions and the impact of added amylose–lipid complexes. Carbohydr. Polym. 2011, 84, 1141–1150. [127] Miles, M. J., Morris, V. J., Orford, P. D., Ring, S. G., The roles of amylose and amylopectin in the gelation and retrogradation of starch. Carbohydr. Res. 1985, 135, 271– 281. [128] Eliasson, A. C., Gudmundsson, M., in: Eliasson, A. C. (Ed.), Carbohydrates in food, Marcel Dekker, New York 1996, pp. 431–503. [129] Slade, L., Levine, H., in: Stivala, S. S., Crescenzi, V., Dea, I. C. M. (Eds.), Industrial polysaccharides, the impact of biotechnology and advanced methodologies, Gordon Breach, New York 1987, pp. 387–430. [130] Zeleznak, K. J., Hoseney, R. C., The role of water in the retrogradation ofwheat starch gels and bread crumb. Cereal Chem. 1986, 63, 407–411. [131] Lin, Y. S., Yeh, A. I., Lii, C. Y., Correlation between starch retrogradation and water mobility as determined by differential scanning calorimetry (DSC) and nuclear magnet-ic resonance (NMR). Cereal Chem. 2001, 78, 647–653. [132] Kalichevsky, M. T., Orford, P. D., Ring, S. G., The retrogradation and gelation of amylopectins from various botanical sources. Carbohydr. Res. 1990, 198, 49–55. [133] Gudmundsson, M., Eliasson, A. C., Retrogradation of amylopectin and the effects of amylose and added surfactants/emulsifiers. Carbohydr. Polym. 1990, 13, 295–315. [134] Klucinec, J. D., Thompson, D. B., Amylose and amylopectin interact in retrogradation of dispersed high-amylose starches. Cereal Chem. 1999, 76, 282–291. [135] Parovuori, P.,Manelius, R., Suortti, T., Bertoft, E., Autio, K., Effects of enzymically modified amylopectin on the rheological properties of amylose–amylopectin mixed gels. Food Hydrocolloids 1997, 11, 471–477. [136] Jane, J. L., Chen, J. F., Effect of amylosemolecular size and amylopectin branch chain length on paste properties of starch. Cereal Chem. 1992, 69, 60–65. [137] Biliaderis, C. G., in: BeMiller, J., Whistler, R. (Eds.), Starch: Chemistry and Technology, Academic Press, New York 2009, pp. 293–372. [138] Dhillon, S., Seetharaman, K., Rheology and texture of starch gels containing iodine. J. Cereal Sci. 2011, 54, 374–379. [139] Karam, L. B., Grossmann, M. V. E., Silva, R., Ferrero, C., Zaritzky, N. E., Gel textural characteristics of corn, cassava and yam starch blends: A mixture surface response methodology approach. Starch/Stärke 2005, 57, 62–70. [140] Vandeputte, G. E., Vermeylen, R., Geeroms, J., Delcour, J. A., Rice starches. III. Structural aspects provide insight in amylopectin retrogradation properties and gel texture. J. Cereal Sci. 2003, 38, 61–68. [141] Wang, L. Z., White, P. J., Functional properties of oat starches and relationships among functional and structural characteristics. Cereal Chem. 1994, 71, 451–458. [142] Hibi, Y., Roles of water-soluble and water-insoluble carbohydrates in the gelatinization and retrogradation of rice starch. Starch/Stärke 1998, 50, 474–478. [143] Luallen, T., in: Eliasson, A. C. (Ed.), Starch in Food: Structure, Function and Applications, Woodhead Publish-ing, Cambridge 2004, pp. 393–424. [144] Craig, S. A. S.,Maningat, C. C., Seib, P. A., Hoseney, R. C., Starch paste clarity. Cereal Chem. 1989, 66, 173–182. [145] Jane, J. L., Kasemsuwan, T., Chen, J. F., Juliano, B. O., Phosphorus in rice and other starches. Cereal Foods World 1996, 41, 827–832. [146] Jacobson,M. R., Obanni,M., BeMiller, J. N., Retrogradation of starches from different botanical sources. Cereal Chem. 1997, 74, 511–518. [147] Nuwamanya, E., Baguma, Y., Wembabazi, E., Rubaihayo, P., A comparative study of the physicochemical properties of starches from root, tuber and cereal crops. Afr. J. Biotechnol. 2011, 10, 12018–12030. 2014 WILEY-VCH Verlag GmbH Co. KGaA, Weinheim www.starch-journal.com
  • 16. 16 J. Waterschoot et al. Starch/Stärke 2014, 66, 1–16 [148] Aviara, N. A., Igbeka, J. C., Nwokocha, L. M., Effect of drying temperature on physicochemical properties of cassava starch. Int. Agrophys. 2010, 24, 219–225. [149] Bello-Perez, L. A., Ortiz-Maldonado, F., Villagomez-Mendez, J., Toro-Vazquez, J. F., Effect of fatty acids on clarity of starch pastes. Starch/Stärke 1998, 50, 383–386. [150] Sodhi, N. S., Singh, N., Morphological, thermal and rheological properties of starches separated from rice cultivars grown in India. Food Chem. 2003, 80, 99–108. [151] Goff, H. D., in: Eliasson, A. C. (Ed.) Starch in Food: Structure, Function and Applications, Woodhead Publish-ing, Cambridge 2004, pp. 425–440. [152] Ferrero, C., Zaritzky, N. E., Effect of freezing rate and frozen storage on starch–sucrose-hydrocolloid systems. J. Sci. Food Agric. 2000, 80, 2149–2158. [153] Srichuwong, S., Isono, N., Jiang, H., Mishima, T., Hisamatsu, M., Freeze–thaw stability of starches from different botanical sources: Correlation with structural features. Carbohydr. Polym. 2012, 87, 1275–1279. [154] Navarro, A. S., Martino, M. N., Zaritzky, N. E., Viscoelastic properties of frozen starch–triglycerides systems. J. Food Eng. 1997, 34, 411–427. [155] Zheng, G. H., Sosulski, F. W., Determination of water separation from cooked starch and flour pastes after refrigeration and freeze–thaw. J. Food Sci. 1998, 63, 134–139. [156] Li, L. L., Kim, Y., Huang, W. N., Jia, C. L., Xu, B. C., Effects of ice structuring proteins on freeze–thaw stability of corn and wheat starch gels. Cereal Chem. 2010, 87, 497–503. [157] Jobling, S. A., Westcott, R. J., Tayal, A., Jeffcoat, R., Schwall, G. P., Production of a freeze–thaw-stable potato starch by antisense inhibition of three starch synthase genes. Nat. Biotechnol. 2002, 20, 295–299. [158] Charoenrein, S., Tatirat, O., Rengsutthi, K., Thongngam, M., Effect of konjac glucomannan on syneresis, textural properties and the microstructure of frozen rice starch gels. Carbohydr. Polym. 2011, 83, 291–296. [159] Arunyanart, T., Charoenrein, S., Effect of sucrose on the freeze–thaw stability of rice starch gels: Correlation with microstructure and freezable water. Carbohydr. Polym. 2008, 74, 514–518. [160] Yuan, R. C., Thompson, D. B., Freeze–thaw stability of three waxymaize starch pastes measured by centrifugation and calorimetry. Cereal Chem. 1998, 75, 571–573. 2014 WILEY-VCH Verlag GmbH Co. KGaA, Weinheim www.starch-journal.com