You are on page 1of 10

Journal of Cereal Science 91 (2020) 102897

Contents lists available at ScienceDirect

Journal of Cereal Science


journal homepage: http://www.elsevier.com/locate/jcs

Structural, rheological and gelatinization properties of wheat starch


granules separated from different noodle-making process
Jing Hong a, Di An a, Limin Li a, Chong Liu a, Mingfei Li a, Roman Buckow b, Xueling Zheng a, *,
Ke Bian a
a
School of Grain, Oil and Food Science, Henan University of Technology, Zhengzhou, 450001, China
b
CSIRO, Agriculture and Food, Werribee, VIC, 3030, Australia

A R T I C L E I N F O A B S T R A C T

Keywords: In this study, three typical wheat cultivars (ZM366, AK58, and ZM103) with high, medium, and low gluten
Starch strength, respectively, were selected as the raw material. The starch granules separated from different stages of
Structure the noodle-making process, including kneading, resting, sheeting, cutting, and drying, were used to explore the
Dynamic rheology
structure, dynamic rheology, and quality of the noodles. The D50 (median diameter) of the starch granules
Noodle quality
decreased during the noodle-making process, and the reduction was enhanced by an increase in the gluten
strength of the flour. Between steps 4 and 5 of the noodle-making process, the solubility of ZM103 variety
increased from 4.3% to 5.0% at 80 � C, while the peak viscosity decreased from 3626 to 3386 mPa s, which
resulted in a decrease in the cooking loss of noodles. Similar trend was observed in the ZM366 and AK58 va­
rieties. The gelatinization enthalpy was reduced, suggesting that the crystalline regions of the starch granules
were destroyed during the kneading process. Between steps 4 and 5 of the noodle-making process, the elastic
modulus of the starch granules significantly increased, while the temperature at which maximum elastic modulus
was decreased, indicating an increase in the crystalline stability of starch during the drying process. Correlation
analysis indicated that the changes occurred to the gelatinization property was primarily due to the change in the
particle size.

1. Introduction leach-out during cooking, form a continuous phase in white salted


noodles, thus providing higher elasticity. The swelling and gelatiniza­
Starch, composed of amylose and amylopectin, is the main carbo­ tion of the starch granules in noodles can differ during cooking. Lower
hydrate (~75%) in wheat flour. Wheat starch has a bimodal size dis­ swelling and gelatinization of the starch granules in noodles provides it
tribution, with large A-type and small B-type starch granules. A-type with higher elasticity and chewiness (Li et al., 2017).
starch granules have a lenticular shape with a diameter of 10–40 μm, Generally, noodle manufacture involves several processing steps
whereas B-type starch granules have a spherical shape with a smaller including mixing, kneading, resting, compounding, sheeting, cutting,
diameter, which is less than 10 μm (Meredith, 2010). Among the and drying. The amount of water added in the mixing step has a marked
physiochemical properties of starch, the ratio of amylose to amylo­ impact on the noodle quality, in terms of its hardness and elasticity. The
pectin, particle size distribution, paste viscosity, solubility, and swelling production of dried noodles has been industrialized, and the amount of
power are highly correlated with the texture and sensory qualities of water added during its manufacture has been reduced to prevent the
noodles. Notably, wheat flour with low amylose content, high peak formation of gluten network and optimize energy and water use. How­
viscosity, and low gelatinization temperature has resulted in higher ever, insufficient formation of gluten network during the noodle-making
water absorption, lower cooking loss, and better sensory quality of process can reduce the sensory quality of noodles (Hu et al., 2007).
noodles (Kaur et al., 2016). The noodle quality is affected by the prop­ Zhang et al. (2012) found that water addition, mixing time, and resting
erty of starch in the raw material and during cooking. Heo et al. (2012) time were the main factors affecting the springiness and hardness of
reported that the amylose and short chain of amylopectin, which noodles. The hardness of noodles decreased with an increase in water

* Corresponding author.
E-mail address: xlzhenghaut@126.com (X. Zheng).

https://doi.org/10.1016/j.jcs.2019.102897
Received 5 August 2019; Received in revised form 13 December 2019; Accepted 15 December 2019
Available online 17 December 2019
0733-5210/© 2019 Elsevier Ltd. All rights reserved.
J. Hong et al. Journal of Cereal Science 91 (2020) 102897

addition, while its springiness first increased and then decreased with mesh sieve. In this study, the starch yield was ~55%. The basic com­
prolonged mixing and resting times. Drying temperature and drying ponents of the separated starches are shown in S3 (Supplementary ma­
time can also influence the textural properties and cooking quality of terials). Starch separated from step 1 was labeled as ZM366s-1, AK58s-1,
dried noodles. It has been proved that higher drying temperature in and ZM103s-1 and the corresponding flour was labeled as ZM366-1,
noodle-making process can generate higher swelling power but lower AK58-1, and ZM103-1. Similarly, starch separated from steps 2–5 was
solubility and peak viscosity of the starch granules leading to an increase labeled as ZM366s-2, AK58s-2, ZM103s-2; ZM366s-3, AK58s-3, ZM103s-
in the firmness of pasta (Yue et al., 1999). 3; ZM366s-4, AK58s-4, ZM103s-4; and ZM366s-5, AK58s-5, ZM103s-5,
Although the influence of diverse raw materials on the properties of respectively. The corresponding flour grinded from the dough/flour/
starch and the cooking quality of noodle have been observed in previous noodle during steps 2–5 was labeled as ZM366-2, AK58-2, ZM103-2;
studies, few study has focused on the changes in starch properties when ZM366-3, AK58-3, ZM103-3; ZM366-4, AK58-4, ZM103-4; and
separated from different stages of the noodle-making process. In this ZM366-5, AK58-5, ZM103-5, respectively.
study, the physicochemical properties of the starch granules were
investigated during the mixing, kneading, resting, rolling, sheeting,
2.3. Morphological property
cutting, and drying processes to reveal the influence of changes in starch
properties on noodle quality.
The morphology of the samples obtained at different steps was
observed using a scanning electron microscope (S–3700N, HITACHI,
2. Materials and methods
Japan). The experiment was performed as described by Pu et al. (2017).
A small quantity of each sample was directly coated onto the surface of
2.1. Materials
the stub. All samples were coated with a thin gold layer before investi­
gation, and then examined under the scanning electron microscope at an
Three typical wheat cultivars of Zhengmai 366 (ZM366, high gluten
acceleration voltage of 10 kV.
strength), Aikang 58 (AK58, medium gluten strength), and Zhengmai
103 (ZM103, low gluten strength) were selected as the raw materials.
These cultivars were kindly supplied by the Henan Academy of Agri­ 2.4. Amylose content
cultural Sciences (Zhengzhou, China). Wheat flour was obtained by an
experimental mill of MLU-202 Bühler mill (Bühler Equipment Engi­ Amylose content was determined according to the AACC 61-03
neering Co. Ltd., Switzerland), according to the AACC 26-20 method method.
(AACC, 2000). The basic components of these three flour types are
shown in S1 (Supplementary materials). The starch content of the flour
2.5. Damaged starch
from ZM366, AK58, and ZM103 varieties was 75.4%, 79.6%, and 81.1%,
respectively, while the protein content of the flour from these varieties
Damaged starch content was determined using an SDmatic (Chopin
was 13.2%, 12.7%, and 11.0%, respectively. All chemicals used in the
Technologies, France), according to the AACC 76–33.01 method (AACC,
study were of analytical grade.
2000).

2.2. Preparation of samples from different stages of the noodle-making


process 2.6. X-ray diffraction (XRD)

Wheat flour, prior to mixing with water, was defined as step 1. The The structure of the starch granules was observed using a D8
noodle-making process comprised mixing and kneading (step 2), resting ADVANCE X-ray diffractometer (Siemens, Munich, Germany). Samples
(step 3), rolling, sheeting, and cutting (step 4), and drying (step 5) were scanned at the rate of 0.2 s under the following conditions: Cu-Kα
processes. Wheat starch was separated from five stages in the noodle- radiation (λ ¼ 0.15418 nm), 5–40� angular range, at 0.04� step intervals,
making process, as seen in S2 (Supplementary materials). Wheat flour, 40 kV voltage, and 40 mA current. The crystallinity (%) was described as
prior to mixing with water, was defined as step 1. After being mixed with the diffraction ratio of the peak to total area.
50% distilled water, which was based on the water absorption (as shown
in S1) and detected by the farinograph (810152Farinograph-AT, Ger­
2.7. Particle size distribution
many), the crumbly dough was formed from wheat flour. This process
was considered as step 2. Following removal from the container, the
Dry powder laser particle size analyzer (BT9300H, Dandong, China)
crumbly dough was sealed and rested at room temperature (25 � C) for
was used to determine the particle size of starch granules. This instru­
20 min, which was considered as step 3. The dough was subsequently
ment was used to analyze the distribution of starch particle size ac­
rolled and sheeted in a noodle machine (Model JMTD-168/140; Beijing;
cording to the principle of optics, which measures the spatial frequency
China) for 5 times, with a roller gap of 2 mm, and 6 more times with
spectrum of the particles (Ke and Lu, 2017).
lower roller gap, until the thickness of the dough sheets was reduced to
1 mm. Next, the dough sheet was cut into strips of fresh noodles with a
width of 2 mm. This was defined as step 4. Following drying at a con­ 2.8. Solubility and swelling power
stant temperature of 40 � C and relative humidity of 75% (SP–18S;
Jiangsu; China) for 10 h, it was allowed to stand at room temperature for The solubility and swelling power analysis was performed using the
10 h. The dried noodles were prepared, and this was regarded as step 5. method described by Hong et al. (2018). About 0.4 g (dry basis) of each
From each noodle-making step, 300 g (dry basis) of flour was mixed sample was mixed with distilled water to obtain a 2% (w/w) starch
with 150 mL distilled water in a pin-mixer, and rested for 20 min at room slurry in a centrifuge tube. The suspensions were heated at 80 � C or 90
temperature. Gluten was aggregated by washing 3–4 times with 500 mL �
C for 30 min. The cooked samples were cooled to room temperature and
distilled water. The slurry was sieved through 100 meshes and made to centrifuged at 1350 g for 15 min. The supernatants were placed in glass
stand for 6 h. Following centrifugation at 1350g for 15 min, the super­ Petri dishes and dried at 105 � C until constant weight was attained.
natant and the upper sediment layer of protein were discarded, while the Solubility was expressed as the percentage of dry solid weight based on
remaining sediment was collected as wheat starch. All starch samples the weight of the dried sample. The swelling power was calculated as the
were freeze-dried by vacuum filtration (FD-1A-50; Beijing; China), to a weight ratio of the wet sediment to the initial dry matter excluding the
moisture content of ~10%, and subsequently passed through a 120- water-soluble starch. The experiments were carried out in triplicate.

2
J. Hong et al. Journal of Cereal Science 91 (2020) 102897

2.9. Pasting properties The pre-test, test, and post-test speeds were 2.0 mm/s, 0.8 mm/s, 0.8
mm/s. The strain was 75% and the trigger force was 5.0 g. From the
The pasting properties of wheat starch samples were determined resulting curves, hardness (g), springiness, cohesiveness, and chewiness
according to the AACC International Method 76–21, using a Rapid Visco were determined. For tensile strength analysis, the instrument was
Analyser (RVA-4; Newport Scientific, Warriewood, Australia). Approx­ equipped with spaghetti/noodle tensile grips (A/SPR). The distance
imately 3.5 g starch sample (corrected to 14% wet basis) was weighed between the parallel rollers was 4.0 cm. The pre-test, test and post-test
into the RVA canister and 25 mL distilled water was added. The sample speeds were 2.0 mm/s, 2.0 mm/s and 10 mm/s, respectively. The dis­
was heated from 50 to 95 � C at the rate of 9 � C/min within 5 min, tance was 140 mm and the trigger force was 5.0 g. Tensile strength and
maintained at 95 � C for 3 min, and then cooled to 50 � C. The rotation break length were obtained from the tensile curves (Zhu et al., 2010).
speed was set at 160 rpm. Peak, trough, and final viscosities were ob­
tained from the pasting curve, and the breakdown value was the dif­ 2.14. Statistical analysis
ference between the peak viscosity and trough viscosity, while the
setback value was the difference between the final and trough viscos­ All experiments were performed in triplicate and the standard de­
ities. The experiments were carried out in triplicate. viation was determined. Analysis of variance (ANOVA) was performed
to compare the treatment means. The differences were analyzed, and the
2.10. Thermal stability difference was considered significant at a 95% level (p<0.05) using
Duncan’s multiple range test. Correlation coefficients were analyzed
Thermal stability was determined by differential scanning calorim­ using Pearson’s correlation test, and correlations were considered sig­
etry (DSC) (Q20, TA, USA) as described in our previous study (Hong nificant at the P0.05 and P0.01 levels. SPSS Statistics 17.0 and Origin 8.0
et al., 2016). Starch samples (2.5 mg, dry basis) were weighed directly software were used for data analysis.
into an aluminum pan, and distilled water (7.5 μL) was added to obtain a
starch-water ratio of 1:3 (w/w). Prior to analysis, the pan was hermet­ 3. Results and discussion
ically sealed and allowed to equilibrate for 2 h. The pan was then heated
from 20 to 120 � C, at the rate of 10 C/min. A sealed empty pan was used 3.1. Morphology of the dough/noodle obtained from different stages of

as the reference. The onset temperature (To), endothermic peak tem­ the noodle-making process
perature (Tp), final temperature (Tc) and enthalpy of gelatinization
(△H, J/g) were determined. The experiments were carried out in The SEM images of the dough/noodle from the different stages of the
triplicate. noodle-making process are shown in Fig. 1. The dispersibility of the
ZM366-1 and AK58-1 varieties obtained in step 1, was more evenly
2.11. Rheological properties compared to that of the ZM103-1 variety. During kneading, wheat flour
was mixed with water to form dough crumbles, which induces a
A rheometer (Haake RS600, Germany), equipped with parallel-plate connection between the starch and gluten networks from steps 1 to 2. An
geometry (P35Ti), was used to analyze the rheological properties. The incomplete gluten network was observed in ZM366-2, AK58-2, and
gap, strain, and frequency were 1.0 mm, 1% and 1.0 Hz, respectively. ZM103-2, due to the small amount of water added, which resulted in the
These values were within the linear viscoelastic range. Starch suspen­ exposure of most of the starch granules from the gluten network. During
sions (10%) were stirred for 1 h, in sealed vials, at room temperature. the resting phase (steps 2 to 3), the internal structure of the dough
Subsequently, 1 mL of the starch suspension was immediately trans­ crumble was relaxed and transformed into a low-energy state, and
ferred between the plates of the rheometer. Silicone oil and a protective simultaneously, the gluten was aggregated and formed adequately. The
cover were used to prevent a decrease in the moisture content of the formation of the gluten network was not noticeable when the moisture
sample while heating. The elastic modulus of the starch suspension was content of the dough crumble was lower than 35% (Don et al., 2005).
determined by Temperature Climb Mode. The starch suspension was Following multiple sheeting from steps 3 to 4, the gluten fragments
heated from 25 to 90 � C at the rate of 3.5 � C/min, and maintained at 90 gradually aggregated to form a continuous gluten network. As seen from

C for 10 min. Finally, the sample was cooled from 90 to 25 � C at the rate Fig. 1, the gluten network of ZM366-4 and AK58-4 was more directional
of 3.5 � C/min. The storage modulus (G’) of the starch suspension was and completely formed, compared to that of the ZM103-4. The gluten
recorded. network agglomerated during drying (steps 4 to 5), and the cross section
of the noodle became neat and firm (Güler et al., 2002). Pores were
2.12. Cooking properties observed in the noodle cross section in step 5, and the porosity of the
sample was reduced with decreasing gluten strength of the raw material,
Optimal cooking time and cooking loss were determined according indicating that the agglomeration of the gluten network in the drying
to the protocol of Pu et al. (2017), with a slight modification. Twenty stage was associated with the gluten strength of the raw material.
noodle strands were cooked in 500 mL of boiling distilled water. The
optimum cooking time (OCT) was evaluated by squeezing the noodles 3.2. Basic components
between two glass slides every 5 s, and observing the time when the
white core of the noodles disappeared. Next, 25 g of noodles was cooked The basic components of the wheat starch separated at 5 steps of the
in 500 mL of boiling distilled water for 4 min. After being rinsed in water noodle-making process are listed in S3 (Supplementary materials). At
at 25 � C for 30 s, and drained for 5 min, the cooked noodles were step 1, the protein content of the separated starch granules was 0.58%
weighed and labeled as A. The water was evaporated to a constant (ZM366s-1), 0.62% (AK58s-1), and 0.68% (ZM103s-1). This indicates
weight (M), at 105 � C, in an oven (FXB 101–2, Shanghai). Cooking loss that starch granules with low protein content are separated from high-
(CL) and cooking yield (CY) were calculated as follows: CL ¼ M/25 � gluten flour, thereby demonstrating the weak binding capacity of
100% and CY ¼ [(A-25)/25] � 100%. starch and protein, caused by high levels of glutenin aggregation (Borght
et al., 2005). The total starch and protein content of the wheat starch
2.13. Textural properties grains did not show any significant difference during the different steps
of the noodle-making process. Furthermore, S3 (Supplementary mate­
Textural properties of the cooked noodles were determined using a rials) shows that damaged starch was higher in samples from step 2
TA-XT Plus texture analyzer (Stable Micro Systems, UK). For texture compared to step 1, suggesting that kneading and mixing increased the
analysis, a cylindrical probe with a diameter of 35 mm (P/35) was used. damaged starch content. Similarly, a significant increase in the damaged

3
J. Hong et al. Journal of Cereal Science 91 (2020) 102897

Fig. 1. SEM images of dough/noodle samples in different stages of the noodle-making process.

starch content was observed in step 4 compared to step 3, indicating that from step 1, the starches separated from step 2 had lower crystallinity,
multiple sheeting and cutting (step 4) processes increased the damaged which can be explained by the exerted mechanical force on starch
starch content. granules during kneading (steps 1 to 2) (Martínez-Bustos et al., 2007).
Huang et al. (2008) pointed out that mechanical force exertion damages
the crystalline structure of starch, and causes the amorphous region to
3.3. Structural analysis become larger. However, the drying stage (steps 4 to 5) in the
noodle-making process had a different influence on crystallinity.
Wheat starch consists of semi-crystalline granules composed of Compared to the crystallinity of starches separated from step 4, the
alternating amorphous and crystalline lamellae, with a height of 9–10 crystallinity of the ZM366 starch granules from step 5 was increased to
nm. The crystalline lamellae consist of double helices of amylopectin, 0.98%, while that of AK58 and ZM103 was decreased to 0.89% and
while the amorphous lamella consist of branching regions of amylo­ 0.49%, respectively. Liu et al. (2015) confirmed that starches with
pectin and some amylose chains. Due to the different packing arrange­ higher damaged content, ZM366>ZM103>AK58 (shown in S3), were
ments of double helices, diverse crystalline structures, including A-type, more sensitive to the drying process from steps 4 to 5. It has been re­
B-type, and C-type, have been detected in wheat starches (Borght et al., ported that hydrothermal treatment, for example, the drying process,
2005). As shown in Fig. 2, all samples possessed the characteristic exerted a different effect on the crystallinity of starch granules (Hoover
monoclinic A-type crystalline packing with strong peaks at 15� and 23� , and Manuel, 1996; Vieira and Sarmento, 2010). Hoover and Manuel
and an unsolved double peak at 17� and 18� (Borght et al., 2005). (1996) found an increase in the crystallinity of starch, while the findings
Similar diffraction peaks were observed in starches separated from of Vieira and Sarmento (2010) were the opposite. Additionally, subse­
ZM366, AK58, and ZM103 varieties (from steps 1 to 5), indicating that quent analysis of enthalpy, detected by DSC, reflected the crystalline
no significant effect was detected on the crystalline structure of starch, structure of starch. In this study, as seen in Table 2, the ΔH of starch
during the noodle-making process. Compared to the starches separated

4
J. Hong et al. Journal of Cereal Science 91 (2020) 102897

Generally, D50 was used to describe the median diameter. The D50 of
starch granules from step 1 was in the following order: AK58>ZM366>
ZM103, indicating that the particle size distribution of the starches had
no regular relationship with gluten strength. The D10, D50, and D90
values of starches separated from ZM366, AK58, and ZM103, decreased
from steps 1 to 5. The D50 of starch granules separated from ZM366
decreased from 17.79 μm to 12.51 μm, displaying a reduction of 5.28 μm
from steps 1 to 5, while the D50 of starch granules separated from
ZM103 reduced from 17.70 μm to 14.96 μm, displaying a reduction of
2.74 μm from steps 1 to 5. It is speculated that wheat flour with high
gluten strength can enhance the stress of starch from steps 1 to 5. The
D50 of starch granules separated from ZM366 reduced from 17.79 μm to
15.75 μm from steps 1 to 2, whereas the D50 of starch granules sepa­
rated from ZM103 displayed no significant difference. This may be
ascribed to the stable gluten network of ZM366 formed during the
kneading process, which resulted in increased mechanical force, thereby
causing an increase in starch granules with smaller particle size (Mar­
tínez-Bustos et al., 2007).

3.5. Solubility and swelling power

Solubility and swelling power of starch granules separated from the


different stages of the noodle-making process are shown in Fig. 3. The
solubility of starch granules separated from step 1 was as follows:
ZM103 > AK58 > ZM366, indicating the solubility increased with the
increase of amylose content (as shown in S3). It was found that the
solubility and swelling power increased at higher temperatures (Fig. 3a
and b), owing to the intense fusion of amylose and lipids at high tem­
peratures. The solubility and swelling power are related to the interac­
tion between amylose-amylose and amylopectin-amylopectin chains
(Olayinka et al., 2008). It can be inferred that the amylose-amylose in­
teractions were destroyed during steps 1 to 5, thereby causing an in­
crease in solubility. The solubility of starch granules separated from
ZM366, AK58, and ZM103 increased from 6.76% to 10.41% (3.65%
increased), 8.25%–11.45% (3.20% increased), and 11.13%–14.01%
(2.88% increased) respectively, at 90 � C. A higher increase in solubility
(from steps 1 to 5) was observed in starch granules separated from wheat
flour containing higher gluten strength. This phenomenon, related to the
reduced D50 of starch granules during the noodle-making process,
demonstrated that the solubility increased with an increase in the small
starch granule content. This can be explained by the “close-packing”
property of starch granules, which implies that the starch granules have
reached their maximum swelling capacity and formed a compact
structure, resulting in closer arrangement within a specific volume, prior
to granule disruption (Bagley and Christianson, 2010). Therefore, the
swelling of large particles could confirm the breakage of starch granules
and prevent the leaching out of amylose.

3.6. Pasting property of starch granules

Table 1 shows the pasting properties of starch granules as deter­


mined by the rapid visco-analyzer (RVA) through programmed heating
and cooling cycles. Peak viscosity was deemed as the equilibrium state at
Fig. 2. XRD diffractograms of wheat starches in different stages of the noodle- the edge of intact and ruptured starch granules, during heating. The
making process. peak viscosity of starch granules separated from ZM366, AK58, and
ZM103 at step 1 was 4306 mPa s, 4174 mPa s, and 3831 mPa s,
from ZM366s increased from steps 4 to 5, while that of AK58s and respectively. The peak, trough, breakdown, and final viscosities of
ZM103s decreased (p˂0.05), revealing a variation in crystallinity among starch granules decreased from steps 1 to 5. The peak viscosity of starch
the three starch varieties, which can be confirmed by XRD analysis. granules separated from ZM366 decreased to 751 mPa s from steps 1 to
5, and displayed maximum change compared to the other two cultivars,
which might be due to an increase in smaller starch granules. Blanchard
3.4. Particle size distribution et al. (2012) found that wheat starch granules with larger particle size
exhibited higher peak viscosity, since the swelling of large starch
The particle size distribution of starch granules is listed in S4 (Sup­ granules occupied a greater volume fraction than the swelling of small
plementary materials). D10, D50, and D90 indicate that 10%, 50%, and granules, thereby leading to higher friction and viscosity. The peak
90% of the starch granules have a smaller diameter than this value. viscosity of starch granules separated from ZM366 and AK58 were 291

5
J. Hong et al. Journal of Cereal Science 91 (2020) 102897

Fig. 3. Solubility and swelling power of wheat starches in different stages of the noodle-making process.

mPa s and 210 mPa s, respectively, from steps 1 to 2, while ZM103 3.7. Thermal stability of starch granules
showed no apparent difference, indicating that the reduction in peak
viscosity increased with an increase in gluten strength and damaged Heating starch with sufficient water led to the gelatinization of
starch content, during the kneading and mixing processes. Additionally, starch granules, accompanied by the structural conversion from ordered
the drying process had a different influence on the setback viscosity. crystalline areas to amorphous areas, along with energy conversion. The
Compared to the starch granules separated from step 4, the peak vis­ thermal property of starch granules are shown in Table 2. Tester and
cosity of starch separated from ZM366 at step 5 was increased to 22 mPa Morrison (1990) noted that gelatinization enthalpy was related to the
s, while that of the AK58 and ZM103 starches was reduced to 124 mPa s crystallinity of starch granules, while gelatinization temperatures were
and 42 mPa s, respectively, which may be attributed to the increase in associated with the degree of crystallinity and stability. The gelatiniza­
the crystallinity, as demonstrated by the XRD analysis. tion enthalpy of starch granules separated from ZM366, AK58, and
ZM103 decreased to 1.01 J/g, 1.28 J/g, and 0.61 J/g respectively,

6
J. Hong et al. Journal of Cereal Science 91 (2020) 102897

Table 1 granules.
Pasting properties (RVA) of the wheat starches in different noodle-making units.
Sample Peak Trough Breakdown Final Setback
viscosity/ viscosity/ viscosity viscosity/ viscosity/ 3.8. Rheological property of starch granules
mPa⋅s mPa⋅s /mPa⋅s mPa⋅s mPa⋅s

ZM366s- 4306 � 3357�4a 949 � 42a 5219 � 1861 � Elastic modulus (G0 ) refers to the energy required by the starch
1 67a 22a 26c granules to recover after deformation during a sinusoidal vibration
ZM366s- 4015 � 3127 � 888 � 43b 5089 � 1962�1a period. It has been considered as an important parameter to analyze the
2 71b 27b 28b
rheological properties of starch granules (Chiou et al., 2005). Based on
ZM366s- 4012 � 3167 � 845�6b 5061�1b 1894 �
3 17b 11b 12bc the temperature scanning mode, the G0 of the starch granules separated
ZM366s- 3672 � 2972 � 805 � 19 b
4913 � 1941 � from the different stages of the noodle-making process are shown in
4 34c 24c 43c 19ab Fig. 4. In the beginning, the elastic modulus of starch stayed low for a
ZM366s- 3555 � 2754�5d 751 � 47 b
4717 � 1963 � while, and later increased sharply with increasing temperature (Fig. 4a,
5 17c 38d 32a
AK58s-1 4174�3a 3250 � 924 � 58ab 5245 � 1995 �
b, 4c). The temperature leading to a sharp increase in the elastic
62a 13a 48b modulus was considered as the transition temperature, and above this
AK58s-2 3964 � 2990 � 974 � 38a 5146 � 2156 �
53b 41b 13b 54a
AK58s-3 3931 � 2978 � 953 � 39ab 5093 � 2115�8a
10b 50b 58b
AK58s-4 3872 � 3011�2b 861 � 48b 5180 � 2169 �
45b 36ab 38a
AK58s-5 3809 � 2904 � 905�9b 4949 � 2045 �
38c 29b 40c 12ab
ZM103s- 3831 � 2985�5ab 845 � 23a 4737 � 1752�7ab
1 28a 12b
ZM103s- 3856�5a 3011 � 845 � 19a 4858 � 1846 �
2 14ab 24a 38a
ZM103s- 3852�1a 3070 � 782 � 37a 4788 � 1718 �
3 36a 16b 20abc
ZM103s- 3626 � 2930 � 696 � 28b 4585 � 1655 �
4 10b 18b 16c 34bc
ZM103s- 3386 � 2716 � 670 � 32b 4325 � 1609 �
5 32c 65c 36d 41c

Results were means � standard deviation. Different letters superscripted on the


results indicated significant difference (P < 0.05) at each column between each
parameter.

Table 2
Thermal stability (DSC) of the wheat starches in different noodle-making units.
Sample To/oC TP/oC Tc/oC △H/(J/g)
a a a
ZM366s-1 58.36 � 0.15 63.51 � 0.44 77.22 � 1.12 9.59 � 0.14a
ZM366s-2 58.28 � 0.24a 63.06 � 0.00ab 73.47 � 3.56ab 8.58 � 1.60b
ZM366s-3 58.36 � 0.04a 63.34 � 0.02ab 74.17 � 0.29ab 7.28 � 0.82c
ZM366s-4 57.34 � 0.42b 61.68 � 0.58c 70.52 � 3.06b 6.32 � 0.31d
ZM366s-5 57.65 � 0.36ab 62.31 � 0.50bc 71.58 � 1.13ab 6.87 � 0.24c
AK58s-1 57.73 � 0.24a 62.25 � 0.10b 75.61 � 0.39a 9.28 � 0.26a
AK58s-2 57.42 � 0.01a 62.4 � 0.14ab 75.27 � 0.21a 8.05 � 0.06b
AK58s-3 57.72 � 0.58a 63.32 � 0.14b 72.29 � 0.52a 8.06 � 0.44b
AK58s-4 57.79 � 0.27a 62.99 � 0.38a 75.47 � 1.32a 8.37 � 0.13b
AK58s-5 57.94 � 0.11a 62.47 � 0.29ab 69.56 � 0.06a 6.74 � 0.44c
ZM103s-1 58.61 � 0.48a 63.34 � 0.06a 76.54 � 0.47a 9.66 � 0.08a
ZM103s-2 57.76 � 0.76a 63.96 � 1.39a 75.44 � 0.06b 9.05 � 0.47b
ZM103s-3 58.13 � 0.21a 63.16 � 0.32a 75.28 � 0.03b 8.68 � 0.04b
ZM103s-4 57.69 � 0.05a 62.88 � 0.16a 74.43 � 0.25b 8.58 � 0.04b
ZM103s-5 57.96 � 0.06a 63.20 � 0.08a 72.67 � 0.76c 8.06 � 0.03c

To, TP, Tc,ΔH are represented by onset temperature, peak temperature,


conclusion temperature and gelatinization enthalpy, respectively. Results were
means � standard deviation. Different letters superscripted on the results indi­
cated significant difference (P < 0.05) at each column between each parameter.

during kneading and mixing, which can be attributed to increase in


damaged starch content from steps 1 to 2 (Table 2). These results are
consistent with the findings of Leo
�n et al. (2006), who demonstrated that
damaged starch was hydrated and dispersed spontaneously in cold
water. To, Tp, Tc, and ΔH of starch granules separated from ZM366 were
higher following step 5 (drying), as compared to step 4. This can be
explained by the increase in crystallinity, which requires more energy
for gelatinization. Compared to the starch granules separated from step
4, the starch granules from step 5 (drying stage) had greater thermal Fig. 4. Rheological property of wheat starches in different stages of the noodle-
stability, owing to the enhanced crystalline stability of the starch making process (temperature scanning).

7
J. Hong et al. Journal of Cereal Science 91 (2020) 102897

Table 3
Cooking and textural properties of noodles.
Samples Optimal cooking Cooking Cooking loss/ Hardness/g Springiness Cohesiveness Chewiness Tensile/g Break length/
time/s yield/% % mm

ZM366- 215�7d 130.36 � 10.08 � 4980.76 � 0.93 � 0.01a 0.76 � 0.01a 3506.47 � 13.33 � 135.04 �
4 1.67d 0.77d 100.32a 70.71a 0.11a 4.21a
e c bc
AK58-4 178�4 135.25 � 13.02 � 4794.91 � 0.84 � 0.01 0.65 � 0.01 2698.37 � 11.33 � 52.44 � 2.18b
2.77cd 0.25cd 121.34ab 105.45b 0.03c
ZM103- 160�7f 155.23 � 15.66 � 4378.39 � 95.80c 0.76 � 0.61 � 0.01d 2224.26 � 10.82 � 48.70 � 0.65bc
4 1.72b 0.60b 0.02d 177.01d 0.21d
ZM366- 348�4a 145.67 � 13.18 � 4512.47 � 0.89 � 0.66 � 0.00b 2560.16 � 13.32 � 122.75 �
5 2.06bc 1.61bc 180.36bc 0.01b 51.08bc 0.22a 3.58a
b
AK58-5 300�7 154.70 � 16.17 � 4627.26 � 0.85 � 0.01c 0.64 � 0.00 c
2556.28 � 12.55 � 48.26 � 0.74bc
1.66b 0.84a 103.42abc 23.05bc 0.20b
ZM103- 255�7c 184.64 � 18.24 � 4420.16 � 0.85 � 0.01c 0.62 � 0.00d 2472.52 � 11.35 � 41.81 � 2.23c
5 2.96a 1.41a 110.43c 69.79c 0.31c

Results were means � standard deviation. Different letters superscripted on the results indicate significant difference (P < 0.05) at each column between each
parameter.

temperature the associated starch molecules were found to dissociate starch granules increased. However, the increase in cooking loss was
(Tako et al., 2008). The transition temperature of starch granules mainly attributed to the weakening or cracking of the gluten-starch
separated from ZM366, AK58, and ZM103 from steps 1 to 5 was in the structure (Izydorczyk et al., 2010). With increasing optimal cooking
range of 76–80 � C. With an increase in temperature, the starch granules time from steps 4 to 5, the surface structure of noodles was broken by the
continued to swell, and the crystalline structure was destroyed, causing erosion of water, thereby inducing an increase in the dissolution of
the amylose content to leach out. The leached-out amylose and swollen starch during cooking. The increase in cooking loss through drying was
granules were arranged in a three-dimensional conformation, owing to positively correlated with the increase in solubility of the starch granules
the entanglement of amylose chains, resulting in a rapid rise in the (r ¼ 0.900*) at 90 � C, implying that the solubility of starch granules is an
elastic modulus (Chiou et al., 2005). TG’MAX represents the temperature indicator of the cooking loss of noodles. The cooking loss did not have
at which G0 reaches the highest value. The TG’MAX of starch granules any association with the gluten strength of raw materials.
separated from steps 1 to 4 was in the range of 80–83 � C, while the Following the drying step, the hardness and chewiness of the ZM366
drying process (step 5) caused a decrease in the TG’MAX. Compared to and AK58 starch varieties decreased from steps 4 to 5. This was found to
the TG’MAX of starch granules separated from ZM366, AK58, and ZM103 be associated with an increase in cooking yield. It can be seen from
at step 4, the TG’MAX of starch granules separated from step 5 was Table 3 that the cooking yield of ZM366 and AK58 increased 15.31%
reduced further to 32.16 � C, 5.43 � C, and 3.21 � C, respectively, indi­ and 19.45%, respectively, following drying, but the hardness decreased
cating that the reduction in the TG’MAX of starch granules increased with 468.29 g and 167.65 g, respectively. The degree of swelling and gela­
increasing gluten strength of raw materials, from steps 4 to 5. It is tinization of starch granules in noodle strands varied during the cooking
speculated that higher crystalline stability prevented the gelatinization process. A heterogeneous distribution of starch was observed, that is, the
of starch granules, thus requiring longer time and lower temperature for starch in the peripheral parts was more gelatinized and swollen
G0 to reach its highest value. This was also identified by our analysis of compared to that in the inner part. The variation in the degree of
pasting viscosity and TO, by RVA and DSC. Additionally, the significant swelling and gelatinization of noodles resulted in moderate elasticity,
increase in the G0 of starch granules from steps 4 to 5 might be associated cohesiveness, and chewiness (Li et al., 2017). The degree of swollen
with the increase in solubility, which in turn enhanced the starch granules in the inner part of the noodles was increased with
three-dimensional structure formation during cooling. higher water absorption, which can reduce the variation in the periph­
eral and inner parts of the noodles, thus causing a decrease in the
3.9. Cooking and textural properties of noodles hardness. The cooking yield of ZM103 increased to 29.41%, from steps 4
to 5, with an increase in the hardness to 42 g. Polymeric starch granules
To understand the effects of altered starch properties on end-product with high swelling power are closely stacked within a limited space,
quality, the cooking properties of fresh noodles (from step 4) and dried owing to the surrounding protein network. This results in the increased
noodles (from step 5) were measured. Cooking loss, cooking yield, and hardness of cooked noodles with increasing cooking yield (Li et al.,
optimal cooking time are identified as the primary predictors of overall 2017). The tensile strength of noodles was increased by the drying
cooking performance, by both the consumers, as well as the industry. In process from steps 4 to 5. Moderate drying can exacerbate the movement
general, noodles with high cooking quality have been found to possess of the macromolecular polymer, making it orderly with stable
higher cooking yield, lower broken ratio, and lower cooking loss. arrangement, presented here as a macroscopic increase in tensile
The cooking quality of noodles is shown in Table 3. With increase in strength (Kayserilio et al., 2003).
the protein content, an increase was also observed in the optimum
cooking time of fresh noodles from step 4 and dried noodles from step 5, 3.10. Correlation analysis
while a decrease was seen in the cooking yield and cooking loss. Higher
protein content strengthened the gluten network and reduced the loss of The correlation analysis of the physicochemical indicators of starch
amylose during cooking, thereby reducing the cooking loss effectively granules separated from steps 1 to 5 are shown in Table 4. Protein
(Pu et al., 2017). Following drying from steps 4 to 5, the optimum content was found to negatively correlate with swelling power (r ¼
cooking time, cooking yield, and cooking loss of noodles increased. The 0.880**). Higher protein content resulted in increased protein-starch
increase in cooking yield was caused by the drying process in step 5, interactions. The proteins covered the channels on the surface of the
which reduces the moisture content of the noodles, resulting in limited starch granules, thereby inhibiting the penetration of moisture into the
water absorption by fresh noodles during cooking (Pu et al., 2017). The interior of the starch granules, and reducing the swelling power of the
primary factor contributing to the cooking yield of noodles is the water starch granules (Liu et al., 2007). Damaged starch content showed a
absorbed by the starch granules during gelatinization. Therefore, the negative correlation with D50 (r ¼ 0.638*) and D90 (r ¼ 0.800**),
cooking yield increased from steps 4 to 5, as the swelling power of the suggesting that damaged starch granules produced by mechanical force

8
J. Hong et al. Journal of Cereal Science 91 (2020) 102897
and friction (step 1 to step 2), and extrusion and cutting (step 3 to step 4)
DS-damaged starch content, TRC-total relative crystallinity; S-solubility; SP-swelling power, PV-peak viscosity; TV-trough viscosity; BV-breakdown viscosity; FV-final viscosity; SV-setback viscosity; TO-onset temperature;
TC

during the noodle-making process, reduced the particle size of starch


0.572* 0.304 0.280 1

granules to some extent. Additionally, the smaller the particle size of


△H

wheat starch granules, the more severe the stress, which in turn causes
0.753** 0.891* 1

an increase in the damaged starch content (Tester et al., 1994). Damaged


starch content had a negative correlation with the To (r ¼ 0.562*) and
TP

0.616* 1

Tp (r ¼ 0.520*), indicating that an increase in damaged starch reduces


the thermal stability of starch granules, thus decreasing the gelatiniza­
TP- peak temperature; TC-conclusion temperature; △H-gelatinization enthalpy. * represent statistical significance at the P0.05 level and ** represent statistical significance at the P0.01 level.
TO

tion onset temperature. This may be associated with the destroyed


1

crystalline structure of starch granules thus causing the changes during


0.939**
0.829**
0.875**
0.590*

gelatinization. D50 showed a significant positive correlation with To (r


SV

¼ 0.600*), Tp (r ¼ 0.613*), Tc (r ¼ 0.603*) and ΔH (r ¼ 0.545*),


0.665**
0.621*

demonstrating that the greater the small starch granule content, the
0.473
0.338

0.243

lower the gelatinization temperature and the gelatinization enthalpy.


FV

With the increase of small size starch granules, the specific area of the
0.640*

granules increased thus making the starch granules liable to hydration


0.504

0.260
0.308
0.415
0.083 0.283

and gelatinization (Li et al., 2017). Swelling power had a very significant
BV

0.475 1

positive correlation with Tp (r ¼ 0.742**) and ΔH (r ¼ 0.774**), indi­


0.886**
0.545*

cating that the tendency of the starch granules to absorb water and swell
0.493

0.167
0.137
TV

reduces the energy required for gelatinization.


1
0.147

0.301
0.212
0.528*

0.570*
0.390
0.048
0.173

4. Conclusion
PV

1
0.899**

0.166

0.178
0.124
0.628*

0.623*

In the present study, the starch separated from ZM366 possessed a


0.440
0.193
0.278
90 � C

higher damaged starch content and lower D50 particle size compared to
1

the starch separated from AK58 and ZM103. The increase in solubility,
0.671**

0.742**
0.774**
0.454
0.483
0.319

0.131

as well as the overall decrease in peak viscosity and gelatinization


0.155

0.408
0.166
80 � C

enthalpy from steps 1 to 5, was observed in all three wheat cultivars. The
SP

drying process increases the elastic modulus and decreases the tem­
0.814**
0.790**
0.664**
0.821**
0.142

0.338

perature at which G’ reaches its highest value. Correlation analysis


0.47
0.027
0.100

0.354
0.497
90 � C

indicated that the decrease in particle size was the primary reason
1

behind the altered gelatinization property and thermal stability during


0.649**
0.562*

0.568*
0.738**

the noodle-making process.


0.254
0.018
0.085
0.235

0.319

0.376
0.547*
0.503
80 � C

1
S

Declaration of competing interest


0.583*
0.749**

0.722**
0.652**

0.751**
0.323
0.421

0.348

0.643*
0.066
0.306
0.220

0.463
D90

The authors declare no conflict of interest.


1
Correlation analysis between characteristics of wheat starches in different noodle-making units.

0.660**
0.587*
0.828**

0.657**

0.688**

CRediT authorship contribution statement


0.600*
0.613*
0.545*
0.603*
0.46
0.35
0.181
0.026
0.013
D50

Jing Hong: Conceptualization, Methodology, Writing - review &


0.731**
0.542*
0.904**

0.448
0.190

0.246
0.167

editing. Di An: Writing - original draft, Formal analysis, Investigation.


0.568*

0.276

0.432

0.428
0.400
0.458
0.441
0.373
D10

Limin Li: Investigation. Chong Liu: Investigation, Data curation. Min­


1

gfei Li: Formal analysis. Roman Buckow: Writing - review & editing.
0.143
0.125

0.349
0.040
0.174

0.072
0.571*

0.267
0.253
0.352
0.059

0.399
0.123

0.113

0.297
Xueling Zheng: Supervision. Ke Bian: Resources.
0.44
TRC

0.800**

Acknowledgements
0.638*

0.581*

0.604*
0.562*

0.520*
0.254
0.335

0.225
0.123

0.508
0.263
0.058
0.288
0.147

0.375
0.022
moisture starch content protein content DS

This work was supported by the National Natural Science Foundation


1

of China (31671892, U1604235, 31671810), Fundamental Research


Funds for the Henan Provincial Colleges and Universities in Henan
0.880**

0.823**
0.831**
0.592*
University of Technology (2018QNJH08), Province Key Laboratory of
0.066
0.311
0.316
0.232

0.378

0.041
0.522*

0.591*
0.125

0.288

0.260
0.283
0.149
0.080
Cereal Resource Transformation and Utilization (PL2018008), and
1

Doctoral Scientific Research Start-up Foundation from Henan University


of Technology (2018BS082).
0.021

0.030
0.110
0.226
0.391
0.020
0.076
0.149

0.157

0.079

0.281
0.065
0.163
0.138

0.005
0.126

0.040
0.005

0.055
Appendix A. Supplementary data
1

0.549*

0.554*
0.630*
Supplementary data to this article can be found online at https://doi.
0.281

0.083

0.150

0.429

0.523*
0.116

0.089
0.180

0.017
0.061
0.066

0.079
0.380

0.118
0.235
0.175
org/10.1016/j.jcs.2019.102897.
0.01
1 References
protein content
starch content

S 80 � C
90 � C
SP 80 � C
90 � C
AACC, 2000. Approved Methods of American Association of Cereal Chemists, tenth ed.

moisture
Table 4

AACC International, St. Paul. MN.

△H
TRC
D10
D50
D90

BV
TV
PV
DS

FV
SV
TO

TC
TP
9
J. Hong et al. Journal of Cereal Science 91 (2020) 102897

Bagley, E.B., Christianson, D.D., 2010. Swelling capacity of starch and its relationship to Ke, Z., Lu, Q.Y., 2017. Physicochemical properties of A- and B-type granules of wheat
suspension vicosity-effect of cooking time, temperature and concentration. starch and effects on the quality of wheat-based noodle. Int. J. Food Eng. 13,
J. Texture Stud. 13, 115–126. 253–263.
Blanchard, C., Labour�e, H., Verel, A., Champion, D., 2012. Study of the impact of wheat Le�on, A.E., Barrera, G.N., P�erez, G.T., Ribotta, P.D., Rosell, C.M., 2006. Effect of damaged
flour type, flour particle size and protein content in a cake-like dough: proton starch levels on flour-thermal behaviour and bread staling. Eur. Food Res. Technol.
mobility and rheological properties assessment. J. Cereal Sci. 56, 691–698. 224, 187–192.
Borght, A.V.D., Goesaert, H., Veraverbeke, W.S., Delcour, J.A., 2005. Fractionation of Li, M., Dhital, S., Wei, Y., 2017. Multilevel structure of wheat starch and its relationship
wheat and wheat flour into starch and gluten: overview of the main processes and to noodle eating qualities: starch structure with noodle quality. Compr. Rev. Food
the factors involved. J. Cereal Sci. 41, 221–237. Sci. Food Saf. 16, 1042–1055.
Chiou, B.S., Yee, E., Glenn, G.M., Orts, W.J., 2005. Rheology of starch-clay Liu, Q., Gu, Z., Donner, E., Tetlow, I., Emes, M., 2007. Investigation of digestibility in
nanocomposites. Carbohydr. Polym. 59, 467–475. vitro and physicochemical properties of A-and B-type starch from soft and hard
Don, C., Lichtendonk, W.J., Plijter, J.J., Vliet, T.V., Hamer, R.J., 2005. The effect of wheat flour. Cereal Chem. 84, 15–21.
mixing on glutenin particle properties: aggregation factors that affect gluten function Liu, C., Hong, J., Zheng, X.L., 2015. Effect of heat-moisture treatment on morphological,
in dough. J. Cereal Sci. 41, 69–83. structural and functional characteristics of ball-milled wheat starches. Starch - St€arke
Güler, S., K€
oksel, H., Ng, P.K.W., 2002. Effects of industrial pasta drying temperatures on 69 (5–6), 1–9.
starch properties and pasta quality. Food Res. Int. 35, 421–427. Martínez-Bustos, F., L� opez-Soto, M., Martín-Martínez, E.S., Zazueta-Morales, J.J., Velez-
Heo, H., Baik, B.K., Kang, C.S., Choo, B.K., Park, C.S., 2012. Influence of amylose content Medina, J.J., 2007. Effects of high energy milling on some functional properties of
on cooking time and textural properties of white salted noodles. Food Sci. jicama starch (Pachyrrhizus erosus L. Urban) and cassava starch (Manihot esculenta
Biotechnol. 21, 345–353. Crantz). J. Food Eng. 78, 1212–1220.
Hong, J., Zeng, X.A., Buckow, R., Han, Z., Wang, M.S., 2016. Nanostructure, morphology Meredith, P., 2010. Large and small starch granules in wheat-are they really different?
and functionality of cassava starch after pulsed electric fields assisted acetylation. Starch -St€
arke 33, 40–44.
Food Hydrocolloids 54, 139–150. Olayinka, O.O., Adebowale, K.O., Olu-Owolabi, B.I., 2008. Effect of heat-moisture
Hong, J., Zeng, X.A., Buckow, R., Han, Z., 2018. Structural, thermodynamic and treatment on physicochemical properties of white sorghum starch. Food
digestible properties of maize starches esterified by conventional and dual methods: Hydrocolloids 22, 225–230.
differentiation of amylose contents. Food Hydrocolloids 83, 419–429. Pu, H., Wei, J., Wang, L., Huang, J., Chen, X., Luo, C., Liu, S., Zhang, H., 2017. Effects of
Hoover, R., Manuel, H., 1996. The effect of heat-moisture treatment on the structure and potato/wheat flours ratio on mixing properties of dough and quality of noodles.
physicochemical properties of normal maize, waxy Maize, dull waxy maize and J. Cereal Sci. 76, 236–242.
amylomaize V starches. J. Cereal Sci. 23 (2), 153–162. Tako, M., Tamaki, Y., Konishi, T., Shibanuma, K., Hanashiro, I., Takeda, Y., 2008.
Hu, X.Z., Wei, Y.M., Wang, C., Kovacs, M.I.P., 2007. Quantitative assessment of protein Gelatinization and retrogradation characteristics of wheat (Rosella) starch. Food Res.
fractions of Chinese wheat flours and their contribution to white salted noodle Int. 41, 797–802.
quality. Food Res. Int. 40 (1), 1–6. Tester, R.F., Morrison, W.R., 1990. Swelling and gelatinization of cereal starches. I. effect
Huang, Z.Q., Xie, X.L., Chen, Y., Lu, J.P., Tong, Z.F., 2008. Ball-milling treatment effect of amylopectin, amylose and lipids. Cereal Chem. 67, 551–557.
on physicochemical properties and features for cassava and maize starches. Compt. Tester, R.F., Morrison, W.R., Gidley, M.J., Kirkland, M., Karkalas, J., 1994. Properties of
Rendus Chem. 11, 73–79. Damaged Starch Granules. III. Microscopy and Particle Size Analysis of Undamaged
Izydorczyk, M.S., Lagass� e, S.L., Hatcher, D.W., Dexter, J.E., Rossnagel, B.G., 2010. The Granules and Remnants. J. Cereal Sci. 20 (1), 0–67.
enrichment of Asian noodles with fiber-rich fractions derived from roller milling of Vieira, F.C., Sarmento, S.B.S., 2010. Heat-moisture treatment and enzymatic digestibility
hull-less barley. J. Sci. Food Agric. 85, 2094–2104. of peruvian carrot, sweet potato and ginger starches. Starch - St€ arke 60, 223–232.
Kaur, A., Shevkani, K., Katyal, M., Singh, N., Ahlawat, A.K., Singh, A.M., 2016. Yue, P., Rayas-Duarte, P., Elias, E., 1999. Effect of drying temperature on
Physicochemical and rheological properties of starch and flour from different durum physicochemical properties of starch isolated from pasta 1. Cereal Chem. 76,
wheat varieties and their relationships with noodle quality. J. Food Sci. Technol. 53, 541–547.
2127–2138. Zhang, B., Wei, Y.M., Wei-Jin, L.I., 2012. Comparison of factors influencing noodle
Kayserilio, Lu, B.S., Ufuk, B., Levent, Y., Nuri, A., 2003. Drying temperature and relative sensory qualities. Sci. Agric. Sin. 45, 2447–2454.
humidity effects on wheat gluten film properties. J. Agric. Food Chem. 51, 964–968. Zhu, F., Cai, Y.Z., Corke, H., 2010. Evaluation of Asian salted noodles in the presence of
Amaranthus betacyanin pigments. Food Chem. 118, 663–669.

10

You might also like