You are on page 1of 7

Food Chemistry: X 18 (2023) 100672

Contents lists available at ScienceDirect

Food Chemistry: X
journal homepage: www.sciencedirect.com/journal/food-chemistry-x

Effects of particle size on the structure, cooking quality and anthocyanin


diffusion of purple sweet potato noodles
Han Hu a, Xiang-Yu Zhou b, Ya-Shu Wang a, Yu-xin Zhang a, Wen-Hua Zhou a, Lin Zhang a, *
a
Hunan Key Laboratory of Processed Food for Special Medical Purpose, National Engineering Research Center of Rice and Byproduct Deep Processing, School of Food
Science and Technology, Central South University of Forestry & Technology, Changsha 410004, PR China
b
Faculty of Medical Science, Division of Medicine, University College London, WC1E 6BT, The United Kingdom

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

Keywords: The effects of different particle sizes of purple sweet potato flour (PSPF) on the structure and quality of noodles
Purple sweet potato and the diffusion kinetics of anthocyanins during cooking were studied. As the particle size of the PSPF decreased
Noodle quality (from 269 to 66 μm), the adverse effects of the addition of PSPF on the quality of noodles were reduced. The
Anthocyanins
smaller particle size of PSPF was beneficial for the secondary structure orderliness and the tighter microstructure
Kinetics
Diffusivity
of PSP noodles. The diffusion of anthocyanins in noodles to the soup during cooking could be fitted well with
Fick’s second law, and diffusion coefficients were in the range of 8.3248–14.0893 × 10-9 m2/s. The noodles with
15% 66 μm PSPF showed the best cooking properties, the highest sensory score, the highest anthocyanin
retention ability and a compact and orderly microstructure. Thus, they could be considered as noodles rich in
anthocyanins for commercial application.

1. Introduction habits of 43,880 men aged 32 to 81 over 24 years and found that
habitual intake of anthocyanins reduced the risk of nonfatal myocardial
Noodles are a traditional staple food and the main source of energy infarction by 14%. According to the high content of anthocyanins, PSP
for many people around the world. Noodles prepared from refined was prepared as purple sweet potato flour (PSPF) and added to foods to
wheat flour offer better mouthfeel but are not rich in nutrition (Choo, & provide unique color, flavor and nutrition. Cui and Zhu (2021) found
Aziz, 2010). One important way to enhance the nutritional value of that PSPF improved the color, polyphenol content and antioxidant ac­
noodles is through the incorporation of natural substances. For example, tivity of Chinese steamed bread. Wang et al. (2021) found that the
ginkgo biloba powder was incorporated into noodles to increase the addition of purple sweet potato enhanced the structural properties and
content of flavonoids (Li, Zhou, & Wu, 2022), and grain bran was added reduced the digestibility of the rice product. However, few studies have
to increase the total phenolic content and antioxidant activity in noodles been conducted to investigate the influence of the particle size of PSPF
(Levent, Koyuncu, & Bilgicli, 2020). on the quality of noodles. There have been no studies on anthocyanin
Purple sweet potato (Ipomoea batatas) (PSP) is a special kind of sweet diffusion behavior during noodle cooking.
potato that not only contains the nutrients of a regular sweet potato but Thermal processes such as boiling and blanching of cooking food
is also rich in anthocyanins, a group of flavonoids with a benzopyran lead to the diffusion of naturally soluble components such as anthocy­
structure (Wang, Nie, & Zhu, 2016). Many studies have shown that anins in food into hot water (Igoumenidis, Lekka, & Karathanos, 2016),
anthocyanins contribute to various health effects, such as antioxidative, and diffusion may be a major cause of solute loss compared to thermal
antiobesity, immunomodulatory, antiaging, antihyperglycemic, hep­ oxidation and degradation (Selman, 1994; Delchier et al., 2012). By
atoprotective, antimicrobial and cardiovascular effects (Jokioja et al., fitting the Fick diffusion model, the diffusion rates of dissolved sub­
2020; Chen et al., 2020). A chronic placebo-controlled crossover study stances in food can be quantitatively calculated (Igoumenidis, Lekka, &
(Dohadwala et al. 2011) suggested that 4 weeks of consumption of juice Karathanos, 2016). The current application of this law in the food sector
containing anthocyanins improved endothelial function and reduced the has focused on the calculation of the diffusion rates of active ingredients
risk of cardiovascular disease. Cassidy et al. (2016) observed the dietary in single crops, such as the diffusion of active ingredients in apple and

* Corresponding author at: Hunan Key Laboratory of Processed Food For Special Medical Purpose, National Engineering Laboratory for Rice and Byproduct Deep
Processing, School of Food Science and Engineering, Center South University of Forestry and Technology, Changsha 410004, PR China.
E-mail address: zhanglin840514@126.com (L. Zhang).

https://doi.org/10.1016/j.fochx.2023.100672
Received 25 November 2022; Received in revised form 17 March 2023; Accepted 1 April 2023
Available online 3 April 2023
2590-1575/© 2023 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
H. Hu et al. Food Chemistry: X 18 (2023) 100672

chickpeas (Simpson, Ramírez, & Birchmeier, 2015; Yildirim, Öner, & For PSP dough (PSPD) preparation, wheat flour (85 g) was mixed
Bayram, 2011). This work creatively attempts to analyze the diffusion with PSPF (15 g) thoroughly. The mixed flour was placed in a Micro-
laws of polyphenols in the cooking processes of composite systems such dough LAB (Micro-dough LAB 2800, Dachol Technologies, Stockholm,
as PSP noodles (PSPN) and to study the effects of different formulas on Sweden) and the powder type was selected as “flour” with a water
the diffusion rates of polyphenols. content of 30 g per 100 g of flour. The apparatus was switched on for
In this study, PSPFs with different particle sizes were incorporated mixing. When the maximum viscosity was reached, the dough was
into PSPN (no salt) formulations. The cooking properties, textural considered to be prepared and removed immediately and wrapped in
properties, tensile properties, sensory evaluation, protein secondary plastic wrap.
structure, starch short-range order and microstructure of noodles were The PSPD was placed into a self-sealing bag in an intelligent artificial
investigated. The diffusion law of anthocyanins during the noodle climate incubator (35 ℃) and allowed to stand for 30 min (PDR-150 A,
cooking process was analyzed, and the diffusion coefficient was calcu­ Ningbo Xinzhi Biotechnology Co., Ltd., Ningbo, China). In this study,
lated. This study provides a reference for the development of noodle each formulated dough was prepared in duplicate. One of the doughs
products that are rich in polyphenols. was prepared for anthocyanin content analysis, and the other was rolled
with a small noodle roller machine (DHH-180 A, Yongkang Haiou
2. Materials and methods Electric Appliance Co., Ltd., Jinhua, China) five times to produce a
dough piece (thickness 1.5 mm). The dough piece was passed through
2.1. Materials the noodle machine to obtain PSPN strands with dimensions of 3.0 mm
in width and 200.0 mm in length.
PSP variety Mianzishu 9 was supplied by the Hunan Academy of
Agricultural Sciences (Hunan, China). Wheat flours were purchased 2.6. Cooking properties
from Kaixue Grain & Oil Food Co., Ltd. (Hunan, China). KBr was pur­
chased from Sinopharm Chemical Reagent Co., Ltd. (Beijing, China). KCl Cooking properties of formulated noodles were determined with the
and NaAc were purchased from Sigma–Aldrich (Shanghai) Trading Co., method.
Ltd. (Shanghai, China). Hydrochloric acid (HCl) solution was purchased reported by Li, Zhou, and Wu (2022) with slight modifications. First,
from Xilong Scientific Co., Ltd. (Guangdong, China). All other chemical PSPNs (20 pieces) were cooked in boiling water (500 mL). Noodles were
reagents used were of analytical grade. removed every 30 s during cooking, cut into short 1 cm strips and
squeezed with a transparent glass plate. The time when the white core
2.2. PSPF preparation disappeared was recorded as the optimum cooking time.
Then, a 10 g PSPN sample (20 cm in length) was placed into 400 mL
The PSPF was prepared with the method reported by Cui and Zhu of boiling water until the optimum cooking time. The well-cooked
(2021) with slight modifications. The fresh PSP tubers were cleaned and PSPNs were drained with filter paper and weighed with a microbal­
sliced into 2 mm thick slices. Then, the tubers were cooked, peeled, and ance. Water absorption was calculated as the percentage increase in the
dehydrated in a cabinet dryer (101-3EBS, Yongguang Medical In­ weight of the cooked noodles compared to the weight of uncooked
struments Co., Ltd., Beijing, China) at 60 ◦ C with hot air circulation of noodles. Noodle soup was collected in a volumetric flask and diluted to
6,900 m3/h until the moisture was maintained at 6–10%. The dried PSP 400 mL with DI water. Then, 100 mL of solution was placed into an
tubers were crushed into powder in a pulverizer (FW-400A, Zhongxing aluminum box and dried to a constant weight. Cooking loss was calcu­
Weiye Instrument Co., Ltd., Beijing, China) at a speed of 26,000 r/min. lated as the percentage of dry matter lost during cooking with respect to
The resulting flour was subjected to stainless steel sieves of mesh sizes of uncooked PSPN weight.
180, 140, and 100. After sieving, the PSPFs with different particle sizes
were collected in polyethylene bags, vacuum sealed with a heat sealer 2.7. Texture profile analysis and tensile strength analysis
and stored at 4 ◦ C.
Texture profile analysis (TPA) of cooked PSPN was performed using a
2.3. Particle size distributions texture analyzer (TA-XT2i, Stable Micro System Ltd., UK) with the
method described by Tan et al. (2018) with some modifications. The
Particle size distributions of PSPF and wheat flour were confirmed noodles were placed into 500 mL of boiling water until the optimum
through a particle size analyzer (LS 13320, Beckman Instruments, Inc., cooking time and then immediately rinsed with cold water for 1 min. Six
USA), and sizes were presented as Sauter mean diameter [D4,3 µm]. replicates of noodles were evaluated within 5 min after cooling. The
noodles were placed on a flat metal plate and measured with a cylin­
2.4. Chemical composition analyses drical probe (P/36R, 36 mm). TPA was set to a pretest speed of 2.0 mm/
s, test speed of 0.8 mm/s, posttest speed of 2.0 mm/s, strain of 70%, and
The moisture content of PSPF and wheat flour was determined using interval time of 1 s. The mean value of six replications was used.
a halogen rapid moisture analyzer (JH-HS, Yixind Instrument Co., Ltd., Tensile strength analysis (TSA) was performed with another Code A/
Jiangsu, China). The protein content was determined using the Kjeldahl SPR probe. Parameters were set as follows: pretest speed of 2.0 mm/s,
method with a Kjeltec K9840 analyzer (Haineng Instruments Co., Ltd, test speed of 2.0 mm/s, posttest speed of 10.0 mm/s, and distance of 80
Shandong, China) based on the Association of Official Analytical mm. The tensile strength and distance were detected. The tensile
Chemists (AOAC) method. The starch content, cellulose content and strength was expressed as the ratio of the force acting on the noodles to
total carbohydrate content were measured by a starch assay kit, cellu­ the cross-sectional area of the noodles.
lose assay kit and total carbohydrate assay kit, which were purchased
from Solarbio Technology Co., Ltd, Beijing, China. 2.8. Sensory evaluation

2.5. Preparation of PSPN The sensory evaluations of cooked PSPN were determined with the
method reported by Sun et al. (2019) with slight modifications. The
Depending on the results (Table S1, Figure S2a, Figure S3 and evaluations were carried out by a trained panel of 30 panelists (15 fe­
Table S2) of the effects of different addition levels of PSPF on the males and 15 males, 23–52 years old) selected from the School of Food
qualities of noodles, the optimum formulation of PSPN was determined Science and Technology. All panelists passed the basic taste test, the
to be 85 g wheat flours, 15 g PSPF and 30 g deionized (DI) water. odor test and the color vision test, and their evaluation ability was

2
H. Hu et al. Food Chemistry: X 18 (2023) 100672

routinely checked using individual control cards for each panelist. The Table 1
panel was particularly familiarized with the sensory assessment of Particle size distribution and nutrient distributions of WF and PSPF sieved
diverse wheat products. The PSPNs were cooked until the optimal through different sieves.
cooking time, and 10 g cooked noodles were tasted by each panelist. The Sample Size Moisture Total Starch Protein Cellulose
evaluation finished within 15 min, and the mouths of the sensory panel (D4,3 (%) carbohydrate (g/ (g/100 (g/100
panelists were rinsed with spring water between samples. All samples μm) (μmol/g) 100 g) g) g)

were coded in random order and evaluated on a 9-point hedonic scale (1: WF 80 ± 13.58 ± 9.20 ± 1.30c 67.24 17.24 1.51 ±
extremely disgusted, 5: neither liked nor disliked, 9: extremely liked). 2c 0.08a ± ± 0.04d
0.22a 0.56a
PSPF 269 6.62 ± 19.82 ± 43.44 3.44 ± 2.73 ±
2.9. Scanning electron microscopy mesh ± 0a 0.27d 0.22a ± 0.02b 0.01a
size 0.40b
100
The raw and cooked noodles were lyophilized and fractured by 140 123 7.31 ± 19.84 ± 41.19 3.17 ± 2.40 ±
folding. The noodles were then fixed on the sample stage by conductive ± 1b 0.17c 0.46a ± 0.17c 0.35b
double-sided tape, and a thin layer of gold was sputtered on the cross- 0.33c
180 66 ± 7.97 ± 19.58 ± 37.90 2.71 ± 2.09 ±
section of the sample. The 500 × cross-sectional images were
1d 0.28b 1.32b ± 0.09d 0.22c
observed with scanning electron microscopy (SEM) (TESCAN MIRA 0.08d
LMS, Tescan Co., Ltd., Czech Republic) at an accelerating voltage of 10
Values are the mean ± standard deviation (SD). WF, wheat flour; PSPF, purple
kV.
sweet potato flour. Different letters in the same column indicate significant
differences (p < 0.05).
2.10. Fourier transform infrared spectroscopy analysis
( )12
The hydrogen bonding, protein secondary structure and starch short- Mt D
= 4 (2)
range order in raw and cooked noodles were determined with an Fourier M∞ wt
transform infrared spectroscopy (FTIR) spectrometer (Nicolet iN10 MX,
Thermo Fisher Scientific, USA) according to the method of Zou et al. where Mt is the anthocyanin content in the noodle soup at time t, M∞ is
(2022). The samples and spectral pure KBr (analytical grade) were dried the anthocyanin content when all the anthocyanins in PSPN diffuse into
at 40 ℃ for 8–12 h and mixed at a ratio of 1:100. The background was the noodle soup, D is the diffusion coefficient, w is the width of the
scanned (64 times) in the range of 400 cm− 1-4,000 cm− 1 at a resolution noodles (3.0 mm), and t is the thickness of the noodles (1.5 mm).
of 4 cm− 1. Omnic (version 8.0, Thermo Nicolet Inc, Waltham, MA, USA) Equation (2) can be further simplified to:
and PeakFit (version 4.12, SPSS Inc., Chicago, IL, USA) were used to
k2 wt
analyze the Fourier infrared spectrum. D = (3)
16

where k is the slope of the line fitting the anthocyanin diffusivity


2.11. Diffusion behaviors of anthocyanins in noodles during cooking
(calculated from Mt/M∞) to the square root of the cooking time.
The PSPNs were cut into 100 mm lengths. Weighed 3 g portions of
noodles were submerged in 500 mL of boiling water and cooked for 2.12. Statistical analysis
different times (30 s, 60 s, 90 s, 120 s, 150 s, 180 s, 210 s, 240 s). The
cooked noodles were removed, lyophilized and ground into powder. The The results are expressed as the means ± standard deviations. Sta­
anthocyanin contents were determined using an anthocyanin content kit tistical differences of all the samples were calculated by one-way anal­
(Solarbio Technology Co., Ltd, Beijing, China). The details of the method ysis of variance (ANOVA) and Duncan’s multiple range test (p < 0.05)
for anthocyanin content determination are described in the supporting using SPSS 16.0 (SPSS Inc., Chicago, IL, USA). The graphs were made by
information (Section 1.1). Origin 2019b (Origin Lab Corporation, Northampton, MA, USA).
The noodle soup was collected and cooled in an ice-water bath. The
determination of anthocyanin content in noodle soup was performed
3. Results and discussion
according to the method reported by Lee, Durst, and Wrolstad (2005). A
total of 1.49 g of KCl was accurately weighed and diluted to 100 mL with
3.1. Particle size and nutrient distributions
DI water, and the pH value was adjusted to 1.0 with 0.2 mol/L HCl
solution. NaAc (1.64 g) was dissolved in 100 mL DI water, and the pH
Particle size distribution plays a crucial role in the physicochemical
value was adjusted to 4.5 with 0.2 mol/L HCl solution. Then, noodle
and functional properties of PSPF (Ma & Mu, 2016). As shown in
soup (1 mL) was mixed with 9 mL of buffer solution (pH = 1.0, pH =
Figure S1a and Table 1, the unscreened WF had a particle size of 80 μm.
4.5), and after 40 min in a water bath at 40 ◦ C, the absorbance values
PSPF was sieved through 180, 140 and 100 mesh sieves, leading to D4,3
were measured at wavelengths of 520 nm and 700 nm by a microplate
values of 66, 123 and 269 μm, respectively. PSPF sieved through 180
reader (SpectraMax i3, Molecular Devices, United States). The antho­
mesh sieve had the smaller particles. When it mixed with wheat flour, it
cyanin content was calculated according to equation (1):
provided a narrower particle size distribution than the ones in the
Anthocyanin content/(mg/g) = (A × Mr )/(ε × 1) × Df (1) mixtures, which mixed wheat flour and PSPF sieved through the larger
sieves (100 and 140 mesh) (Figure S1b). The nutrient distributions in
where A= (A520 nm pH 1.0-A700 nm pH 1.0) - (A520 nm pH 4.5-A700 nm pH 4.5), Mr PSPF with different particle sizes are different, which might contribute
is the relative molecular mass of cyanidin-3-glucoside (449.2), ε is the to the functional properties of the products (Ahmed, 2014). As shown in
molar extinction coefficient of cyanidin-3-glucoside (26,900/mol), and Table 1, the water content of PSPF increased with decreasing particle
Df is the dilution factor (10). size. The refined PSPF has a larger specific surface area, making it more
Taking the cooking time as the abscissa and the amount of antho­ capable of absorbing water than coarse PSPF (Chau, Wen, & Wang,
cyanins diffused per gram of PSPN in the noodle soup as the ordinate, a 2006). The contents of protein, cellulose, starch, and carbohydrate in
line graph was drawn to describe the diffusion law of anthocyanins. The PSPF decreased with decreasing PSPF particle size, as macromolecular
diffusion coefficient was calculated by Fick’s second law (equation (2). nutrients were damaged during mechanical crushing. The lower the

3
H. Hu et al. Food Chemistry: X 18 (2023) 100672

Table 2
Cooking properties, textural properties, and tensile properties of cooked PSPNs.
Sample Cooking properties Textural properties and tensile properties

Optimal Cooking Water Cooking Hardness (g) Adhesiveness Springiness Chewiness Tensile Elasticity
Time (s) Absorption (%) Loss (%) (g⋅s) (g⋅s) (g⋅s) Strength(g) Distance (mm)

WFN 190.00 ± 17.32b 3.79 ± 0.11b 7.05 ± 0.18c 2633.27 ± 64.28 ± 4.72d 90.46 ± 1536.67 ± 16.93 ± 26.77 ± 1.60a
17.60a 4.75a 20.63d 1.14a
PSPN- 230.00 ± 17.32a 8.13 ± 0.07a 9.49 ± 2049.13 ± 86.73 ± 4.61c 64.82 ± 1762.50 ± 14.33 ± 19.43 ± 2.61d
15l 0.29a 26.93d 2.20d 16.21c 0.98b
PSPN- 220.00 ± 8.28 ± 0.57a 8.83 ± 2267.91 ± 93.29 ± 1.82b 71.33 ± 3.03c 1886.64 ± 14.62 ± 22.19 ± 2.88c
15m 17.32ab 0.59a 30.17c 26.03b 0.37b
PSPN- 200.00 ± 8.63 ± 0.29a 7.86 ± 2449.40 ± 99.81 ± 2.23a 79.65 ± 1978.70 ± 16.87 ± 24.24 ± 2.31b
15s 17.32ab 0.36b 9.55b 6.14b 18.77a 1.05a

WFN, wheat flour noodle; PSPN-15 refers to formula flours with 15% PSPF; l, m, s indicate the large size (269 μm), medium size (123 μm), and small size (66 μm) PSPF;
different letters in the same column indicate significant differences (p < 0.05).

Fig. 1. SEM images of the cross-sections of raw wheat flour noodles (a), raw PSPNs with different particle sizes (269 μm, 123 μm and 66 μm) of PSPF (b, c, d), cooked
wheat flour noodles (e) and cooked PSPNs with different particle sizes (269 μm, 123 μm and 66 μm) of PSPF (f, g, h).

particle size is, the higher the degree of destruction. These findings were PSPN. However, with the decrease in the particle size, the optimal
consistent with the study of Ma, Wang, Li, and Wang (2020), who re­ cooking time was shortened from 230 s to 200 s. The water absorption of
ported a similar relationship between particle size and the protein PSPN was improved by PSPF addition, probably because PSPF contains
content of wheat flour. more hydrophilic groups (Zhu, & Sun, 2019). As the particle size of PSPF
decreased, the specific surface area of PSPF increased, and more hy­
drophilic groups were exposed, resulting in an increase in the water
3.2. Cooking properties absorption of PSPN. PSPF substitution significantly (p < 0.05) increased
the cooking loss of PSPN. The reason for this might be that the phos­
The optimal cooking time, water absorption, and cooking loss of phoric acid groups in PSPF could weaken the gluten network structure,
PSPN-15 l, PSPN-15 m, and PSPN-15s were measured to evaluate their and the starch granules were easily freed from the gluten network,
cooking quality, and WFN was tested as a control group. As shown in resulting in an increase in the cooking loss (Zhu, & Sun, 2019).
Table 2, the addition of PSPF prolonged the optimal cooking time of

Fig. 2. FTIR (Fourier Transform Infrared Spectroscopy) transmittance spectra of starch in raw (a) and cooked (b) noodles. The Y-axis represents the infrared
transmittance of each noodle sample in the range of 400 cm− 1-4,000 cm− 1.

4
H. Hu et al. Food Chemistry: X 18 (2023) 100672

Furthermore, PSP protein could be cross-linked with gluten protein, Table 3


which was not conducive to gluten protein wrapping of starch granules Secondary structures and short-range order of raw and cooked noodles.
and would also lead to cooking loss (Cui, & Zhu, 2022). PSPN-15s Sample Secondary structures Short-range order
noodles with 66 μm PSPF showed the best cooking quality.
α-helix β-sheet β-turn Random DO DD (995
(%) (%) (%) coil (%) (1047 cm− 1/
3.3. Texture profile and tensile strength analysis cm− 1/ 1022
1022 cm− 1)
The effects of PSPF on the hardness, adhesiveness, springiness, cm− 1)

chewiness, tensile strength and elasticity distance of noodles are shown Raw noodles
in Table 2. The hardness, adhesiveness and springiness values of PSPN WFN 24.86 30.51 24.08 20.52 ± 1.3859 ± 1.4784
± 0.01a ± 0.27a 0.21g 0.0021b
were lower than those of WFN. This is due to the lower gluten concen­
± ±
0.06g 0.0125c
tration in PSPN. The chewiness values of PSPN were higher than those of PSPF- 16.22 24.82 29.46 29.49 ± 1.1731 ± 1.3035
WFN, which is due to the higher cellulose concentration of PSPF 15l ± 0.62d ± 0.01c ± 0.07d 0.0064e ±
(Table 1). When the particle size of PSPF was decreased, all the textural 0.05d 0.0061g
properties of PSPN were enhanced. This may be because the smaller PSPF- 18.54 26.34 28.04 27.10 ± 1.2566 ± 1.3152
15m ± 0.75c ± 0.05b ± 0.03e 0.0004d ±
particles could be better packed into the gluten network, resulting in a 0.43e 0.0005g
compact noodle microstructure (Fig. 1) and better textural properties PSPF- 21.77 26.57 26.31 25.37 ± 1.3225 ± 1.3870
(Table 2). Similarly, Azeem et al. (2020) found that sweet potato-wheat 15s ± 0.34b ± 0.12b ± 0.03f 0.0007c ±
bread prepared from sweet potato flour with the smallest particle size 0.37f 0.0004e
Cooked noodles
showed outstanding resilience, cohesion, gumminess and elasticity
WFN 16.89 20.17 ± 31.74 31.14 ± 1.4426 ± 1.5401 ±
values compared to other breads. ± 0.04d 0.13d ± 0.49c 0.0039a 0.0057a
0.41c
3.4. Sensory evaluation PSPF- 9.49 ± 15.40 ± 39.22 35.89 ± 1.2700 ± 1.3723 ±
15l 0.19g 0.44g ± 0.50a 0.0003d 0.0009f
0.08a
Sensory evaluation of PSPN-15l, PSPN-15m, and PSPN-15s was PSPF- 10.83 16.90 ± 39.08 33.21 ± 1.3398 ± 1.4451 ±
carried out with respect to color, appearance, palatability, toughness, 15m ± 0.65f 0.31f ± 0.27b 0.0003c 0.0103d
viscosity, smoothness, flavor and total score, and WFN was also tested as 0.08a
a control group. As shown in Figure S2, PSPF improved the color, PSPF- 12.75 18.23 ± 36.67 32.30 ± 1.3733 ± 1.4992 ±
15s ± 0.33e 0.06e 0.21b 0.0005b 0.0081b
appearance, palatability, toughness and flavor of the noodles. In terms of
±
0.14b
viscosity and smoothness, PSPN-15l tasted less viscous (7.01 points) and
excessively rough (7.56 points) than the control group (7.67 and 8.07
points). However, with decreasing particle size of PSPF, PSPN-15m 3.6. Structural determination of PSPN with FTIR
(7.72 and 8.24 points) and PSPN-15s (7.75 and 8.30 points) showed
greater advantages than WFN in these two aspects. With the exception of To investigate the structure of PSPN, raw and cooked PSPN-15 and
flavor, the other sensory properties increased with decreasing PSPF WNF were detected with FTIR spectroscopy (Fig. 2). The absorption
particle size. This may be due to the finer PSPF being distributed more peak at approximately 3,600 cm− 1-3,500 cm− 1 is reported to be related
evenly in the noodles and being bound more tightly to the gluten to –OH in the free state. With the formation of intramolecular or inter­
network, so that the color-forming substances in PSPF, anthocyanins, molecular hydrogen bonds, the absorption peak of –OH moved to a
were better protected and the gluten network structure was more stable lower wavelength and overlapped with the vibration absorption peak of
(Azeem et al., 2020). PSPN-15s were considered to be more uniform in N–H, resulting in a wider deformation peak (Liu et al., 2021). As shown
color, more moderate in viscosity and smoother in taste and achieved in Fig. 2a and b, WNF showed a wide and strong absorption peak at
the highest total score (8.39 points). approximately 3,395 cm− 1, indicating that a large number of intra­
molecular or intermolecular hydrogen bonds may exist in the molecular
3.5. Microstructure of raw and cooked PSPN polymer of WNF. With the addition of PSPF, the absorption peak further
approached the lower wavelength (raw noodles, 3,395.03 cm− 1-
The microstructure of noodles greatly affects noodle quality and the 3,383.88 cm− 1; cooked noodles, 3,395.26 cm− 1-3,384.06 cm− 1), and the
retention rates of bioactive ingredients. In this study, scanning electron lower particle size of PSPF exposed more hydrogen bonds, which
microscopy (SEM) was used to investigate the effect of PSPF particle size enhanced the water absorption capacity of PSPF. Furthermore,
on the cross-sectional microstructure of noodles. As shown in Fig. 1a-d, hydrogen bonds are the primary force that maintains the α-helix struc­
the starch granules of the WFN were more closely combined with gluten ture (Jarpa-Parra et al., 2015).
protein, and the surface was relatively smooth and flat. With the addi­ The amide I region (1,700 cm− 1-1,600 cm− 1) is considered a char­
tion of PSPF, the noodle structure appeared to become looser, and holes acteristic band associated with the four types of secondary structures
appeared. However, the reduction in particle size resulted in a signifi­ (β-sheet, 1,600 cm− 1-1,637 cm− 1 and 1,682 cm− 1-1,700 cm− 1; random
cantly tighter noodle structure because the larger particle size of PSPF coil, 1,600 cm− 1-1,637 cm− 1; α-helix, 1,646 cm− 1-1,664 cm− 1; β-turn,
reduced the degree of starch exposure and disrupted the gluten network 1,665 cm− 1-1,681 cm− 1) (Byler & Susi, 1986). As shown in Fig. 2 and
in WFN, while the smaller particle size of PSPF could pack into the Table 3, the addition of PSPF resulted in a significant (p < 0.05) decrease
gluten structure, making the structure more stable (Zhang, et al., 2019). in the contents of α-helices and β-sheets and a significant (p < 0.05)
The microstructures of cooked WFN and PSPN are shown in Fig. 1e- increase in the contents of β-turns and random coils. It is generally
h. WFN had a typical cooked noodle structure with a continuous protein accepted that α-helix and β-sheet contents are positively correlated with
network and tangled structure. With the addition of PSPF, the gluten the orderliness of protein, while β-turn and random coil contents are
concentration was diluted, and larger holes were obtained in cooked related to disorderly structure (Pelton & Mclean, 2000). However, as the
PSPN-15 noodles. When the particle size of PSPF was decreased, part of particle size of PSPF decreased, the α-helix and β-sheet contents signif­
the protein network of the cooked PSPN was rebuilt, and the network icantly (p < 0.05) increased, and the β-turn and random coil contents
structure appeared to be relatively stable. This might be due to the significantly (p < 0.05) decreased. This suggests that the decrease in
greater water absorption and enhanced filling properties of PSPF with a PSPF particle size enhanced the structural orderliness and stability of the
smaller particle size. protein network in PSPN.

5
H. Hu et al. Food Chemistry: X 18 (2023) 100672

Fig. 3. Concentration of anthocyanins in PSPNs (a) and noodle soups (b) during the cooking time. The relationship between PSPN anthocyanin diffusivity and the
square root of cooking time (c).

The FTIR spectrum region of approximately 1,000 cm− 1 was used to affected the structural orderliness of the noodles by altering the number
characterize the short-range order of starch. The peaks at 995 cm− 1 and of hydrogen bonds, the composition of the protein secondary structure
1,047 cm− 1 are related to the crystallization region of starch, while the and the short-range order of starch. The smaller PSPF particle size led to
peak at 1,022 cm− 1 represents the amorphous region. By calculating the a tighter and more stable noodle structure. The anthocyanin content
ratio of these three peaks, the order degree (DO, 1,047 cm− 1/1,022 decreased dramatically during cooking, which was mainly caused by
cm− 1) and double helicity (DD, 995 cm− 1/1,022 cm− 1) can be obtained diffusion. The diffusion behavior of anthocyanins from noodles to soup
(Zou et al., 2022). As shown in Fig. 2 and Table 3, the DO and DD of is in accordance with Fick’s second law. The diffusion coefficients of
PSPN were significantly (p < 0.05) lower than those of WFN. As the anthocyanins ranged from 8.3248 × 10-9 to 14.0893 × 10-9 m2/s. PSPN
particle size of PSPF decreased, the DO and DD of PSPN significantly (p with 15% 66 μm PSPF showed the highest sensory score, most compact
< 0.05) increased, which indicated that the reduction in particle size microstructure and greatest anthocyanin retention rate during process­
could increase the intermolecular contact between amylose and short ing and cooking. This study suggested that PSP could be considered the
chain amylopectin, contributing to the formation of nascent double raw material for noodles due to the high concentration of processable
helices. anthocyanins, which could reduce the risk of cardiovascular disease.
The digestion, absorption and function of PSPN can be further investi­
3.7. Diffusion behaviors of anthocyanins during PSPN cooking gated in future studies.

As shown in Fig. 3, the concentrations of anthocyanins in different CRediT authorship contribution statement
PSPNs decreased during cooking, while the anthocyanin concentrations
in soups increased. These results indicated that the loss of anthocyanins Han Hu: Conceptualization, Methodology, Formal analysis, Inves­
in PSPNs was partly due to anthocyanin diffusion from noodles to soups tigation, Writing – original draft, Supervision. Xiang-Yu Zhou: Soft­
and thermal degradation (Igoumenidis, Lekka, & Karathanos, 2016; ware, Resources, Data curation, Visualization. Ya-Shu Wang:
Rocchetti, Lucini, & Chiodelli, 2017). Owing to the higher concentration Validation. Yu-xin Zhang: . Wen-Hua Zhou: Project administration,
of anthocyanins in PSPN-15s raw noodles (Table S2), the anthocyanin Funding acquisition. Lin Zhang: Conceptualization, Methodology,
content in PSPN-15s was higher than that in the other noodles during the Visualization, Funding acquisition.
cooking time (Fig. 3a).
To investigate the kinetics of anthocyanin diffusion during cooking Declaration of Competing Interest
time, the diffusion rate of anthocyanins and the square root of diffusion
time were fitted with Fick’s second law (Pinelo, Zornoza, & Meye, 2008) The authors declare that they have no known competing financial
(Fig. 3c). The diffusion coefficient of anthocyanins in the PSPN cooking interests or personal relationships that could have appeared to influence
process was calculated according to the kinetic parameters obtained by the work reported in this paper.
Fick’s second law. The diffusion coefficients (D) of PSPN-15l, PSPN-15m
and PSPN-15s were 14.0893 × 10-9 m2/s, 9.9998 × 10-9 m2/s, and Data availability
8.3248 × 10-9 m2/s, the kinetic parameters k were 2.2382 s− 1, 1.8856
s− 1, and 1.7204 s− 1, and the correlation coefficients R2 were 0.9958, Data will be made available on request.
0.9749, and 0.9885. In this study, PSPN-15s made with the finest PSPF
had the lowest D, indicating the slowest diffusion of anthocyanins in Acknowledgments
PSPN-15s. This might be due to the tighter structure of PSPN-15s
(Fig. 1d and h), which makes it less porous and inhibits the leaching This work was supported by the Science and Technology Innovation
of anthocyanins. On the other hand, finer PSPF may promote the non­ Program of Hunan Province (2022RC1148), the Innovation Leading
covalent interaction between starch and anthocyanins (Ferruzzi et al. Plan Project of Hunan province (2021GK4022), Natural Science Foun­
2020) to protect anthocyanins from loss during cooking. Thus, PSPN dation of Hunan Province (No. 2022JJ50260, 2022JJ31009), Natural
with 15% 66 μm PSPF was more beneficial to the prevention of the Science Foundation of Changsha (kq2014150, kq2202282), a program
diffusion of anthocyanins during cooking. for Special Project for the Construction of the Innovative Province in
Hunan Province (2019TP2011) and a program for Science and Tech­
4. Conclusion nology of Education Department of Hunan Province (20B620).

The influence of PSPF particle size on wheat flour noodles was Appendix A. Supplementary data
investigated in this study. PSPF addition could deteriorate the cooking
properties and textural properties of noodles, but this could be offset by Supplementary data to this article can be found online at https://doi.
the decrease in PSPF particle size. FTIR analysis showed that PSPF org/10.1016/j.fochx.2023.100672.

6
H. Hu et al. Food Chemistry: X 18 (2023) 100672

References by the pH differential method: Collaborative study[J]. Journal of AOAC international,


88(5), 1269–1278.
Levent, H., Koyuncu, M., Bilgicli, N., et al. (2020). Improvement of chemical properties
Ahmed, J. (2014). Effect of particle size and temperature on rheology and creep behavior
of noodle and pasta using dephytinized cereal brans[J]. LWT-Food Science and
of barley β-d-glucan concentrate dough. Carbohydrate Polymers, 111, 89–100.
Technology, 128, Article 109470.
Azeem, M., Mu, T. H., & Zhang, M. (2020). Influence of particle size distribution of
Li, L., Zhou, W., Wu, A., et al. (2022). Effect of Ginkgo Biloba Powder on the
orange-fleshed sweet potato flour on dough rheology and simulated gastrointestinal
Physicochemical Properties and Quality Characteristics of Wheat Dough and Fresh
digestion of sweet potato-wheat bread. LWT-Food Science and Technology, 131.
Wet Noodles[J]. Foods, 11(5), 698.
Byler, D. M., & Susi, H. (1986). Examination of the secondary structure of proteins by
Liu, Z., Zheng, Z., Zhu, G., Luo, S., Zhang, D., Liu, F., et al. (2021). Modification of the
deconvolved FTIR spectra. Biopolymers, 25(3), 469–487.
structural and functional properties of wheat gluten protein using a planetary ball
Cassidy, A., Bertoia, M., Chiuve, S., Flint, A., Forman, J., & Rimm, E. B. (2016). Habitual
mill. Food chemistry, 363, Article 130251.
intake of anthocyanins and flavanones and risk of cardiovascular disease in men. The
Ma, M. M., & Mu, T. H. (2016). Effects of extraction methods and particle size
American journal of clinical nutrition, 104(3), 587–594.
distribution on the structural, physicochemical, and functional properties of dietary
Chau, C. F., Wen, Y. L., & Wang, Y. T. (2006). Improvement of the functionality of a
fiber from deoiled cumin. Food Chemistry, 194, 237–246.
potential fruit insoluble fibre by micron technology. International journal of food
Pelton, J. T., & McLean, L. R. (2000). Spectroscopic methods for analysis of protein
science & technology, 41(9), 1054–1060.
secondary structure. Analytical Biochemistry, 277, 167–176.
Chen, K., Wei, X., Zhang, J., Pariyani, R., Jokioja, J., Kortesniemi, M., et al. (2020).
Pinelo, M., Zornoza, B., & Meyer, A. S. (2008). Selective release of phenols from apple
Effects of Anthocyanin Extracts from Bilberry (Vaccinium myrtillus L.) and Purple
skin: Mass transfer kinetics during solvent and enzyme-assisted extraction.
Potato (Solanum tuberosum L. var. ‘Synkea ¨Sakari’) on the Plasma Metabolomic
Separation and Purification Technology, 63(3), 620–627.
Profile of Zucker Diabetic Fatty Rats. Journal of Agricultural and Food Chemistry, 68
Rocchetti, G., Lucini, L., Chiodelli, G., et al. (2017). Impact of boiling on free and bound
(35), 9436–9450.
phenolic profile and antioxidant activity of commercial gluten-free pasta. Food
Choo, C. L., & Aziz, N. A. A. (2010). Effects of banana flour and β-glucan on the
Research International, 100, 69–77.
nutritional and sensory evaluation of noodles[J]. Food Chemistry, 119(1), 34–40.
Selman, J. D. (1994). Vitamin retention during blanching of vegetables. Food Chemistry,
Cui, R., & Zhu, F. (2021). Changes in structure and phenolic profiles during processing of
49(2), 137–147.
steamed bread enriched with purple sweetpotato flour. Food Chemistry, 369, Article
Simpson, R., Ramírez, C., Birchmeier, V., et al. (2015). Diffusion mechanisms during the
130578.
osmotic dehydration of Granny Smith apples subjected to a moderate electric field.
Delchier, N., Reich, M., & Renard, C. M. G. C. (2012). Impact of cooking methods on
Journal of Food Engineering, 166, 204–211.
folates, ascorbic acid and lutein in green beans (Phaseolus vulgaris) and spinach
Tan, H. L., Tan, T. C., & Easa, A. M. (2018). Comparative study of cooking quality,
(Spinacea oleracea). LWT-Food science and technology, 49(2), 197–201.
microstructure, and textural and sensory properties between fresh wheat noodles
Dohadwala, M. M., Holbrook, M., Hamburg, N. M., Shenouda, S. M., Chung, W. B.,
prepared using sodium chloride and salt substitutes. LWT-Food Science and
Titas, M., et al. (2011). Effects of cranberry juice consumption on vascular function
Technology, 97, 396–403.
in patients with coronary artery disease. The American journal of clinical nutrition, 93
Wang, J., Li, M., Wang, C., Dai, Y., Sun, Y., Li, X., et al. (2021). Effect of extrusion
(5), 934–940.
processing and addition of purple sweet potatoes on the structural properties and in
Ferruzzi, M. G., Hamaker, B. R., & Bordenave, N. (2020). Phenolic compounds are less
vitro digestibility of extruded rice. Food & Function, 12(2), 739–746.
degraded in presence of starch than in presence of proteins through processing in
Wang, S., Nie, S., & Zhu, F. (2016). Chemical constituents and health effects of sweet
model porridges. Food chemistry, 309, Article 125769.
potato. Food Research International, 89, 90–116.
Igoumenidis, P. E., Lekka, E. G., & Karathanos, V. T. (2016). Fortification of white milled
Yildirim, A., Öner, M. D., & Bayram, M. (2011). Fitting Fick’s model to analyze water
rice with phytochemicals during cooking in aqueous extract of Mentha spicata
diffusion into chickpeas during soaking with ultrasound treatment. Journal of Food
leaves. An adsorption study. LWT-Food Science & Technology, 65, 589–596.
Engineering, 104(1), 134–142.
Jarpa-Parra, M., Bamdad, F., Tian, Z., Zeng, H., Temelli, F., & Chen, L. (2015). Impact of
Zhang, J., Li, M., Li, C., & Liu, Y. (2019). Effect of wheat bran insoluble dietary fiber with
pH on molecular structure and surface properties of lentil legumin-like protein and
different particle size on the texture properties, protein secondary structure, and
its application as foam stabilizer. Colloids and Surfaces B: Biointerfaces, 132, 45–53.
microstructure of noodles. Grain & Oil Science and Technology, 2(4), 97–102.
Jokioja, J., Linderborg, K. M., Kortesniemi, M., Nuora, A., Heinonen, J., Sainio, T., et al.
Zhu, F., & Sun, J. (2019). Physicochemical and sensory properties of steamed bread
(2020). Anthocyanin-rich extract from purple potatoes decreases postprandial
fortified with purple sweet potato flour. Food Bioscience, 30, Article 100411.
glycemic response and affects inflammation markers in healthy men. Food Chemistry,
Zou, X., Wang, X., Zhang, M., Peng, P., Ma, Q., & Hu, X. (2022). Pre-baking-steaming of
310, Article 125797.
oat induces stronger macromolecular interactions and more resistant starch in oat-
Lee, J., Durst, R. W., Wrolstad, R. E., et al. (2005). Determination of total monomeric
buckwheat noodle. Food Chemistry, 400, Article 134045.
anthocyanin pigment content of fruit juices, beverages, natural colorants, and wines

You might also like