You are on page 1of 10

Food Hydrocolloids 103 (2020) 105707

Contents lists available at ScienceDirect

Food Hydrocolloids
journal homepage: http://www.elsevier.com/locate/foodhyd

Effects of cultivar and growth region on the structural, emulsifying and


rheological characteristic of mango peel pectin
Zhuodan Deng a, 1, Yonggui Pan a, 1, Wenxue Chen a, Weijun Chen a, b, Yonghuan Yun a,
Qiuping Zhong a, Weimin Zhang a, **, Haiming Chen a, b, *
a
College of Food Sciences & Engineering, Hainan University, 58 People Road, Haikou, 570228, PR China
b
Chunguang Agro-product Processing Institute, Wenchang, 571333, China

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

Keywords: Mango peel pectin (MPP) of eight cultivars obtained from different growth regions (China, Thailand, Vietnam,
Mango peel pectin Burma and Australia) exhibit significant different recovery yield and characteristics of structure, emulsifying and
Cultivar rheology. Among the 12 MPP samples, Golek cultivar from Bruma (GoBP) showed the highest galacturonic acid
Growing region
(GalA) content (80.63%), methyl-esterification degree (DM) (71.80%) and average molecular weight (AMW)
Emulsion
(1589 kDa). The thermal analysis showed that the endothermic peak of pectin may related to its degree of
Rheology
methoxylation (DM), and highly branched side chains may enhance its thermal stability. The emulsifying
properties of MPP were assessed by surface tension, ξ-potential, droplet size and microscope observation,
showing that GuCP-stabilized emulsion appears to be more stable. As for the rheological behavior, most pectin
samples exhibited shear-thinning behaviors and GoBP which had the highest AMW, had the highest viscosity
under different shear rate (0.1–100 1/s). GuCP exhibited more elastic-solid characteristics (G0 > G00 ), which could
be used as a functional ingredient in jelly food, while GrCP exhibited higher viscosity (G00 > G0 ) and could be
applied as a thickener. In all, the cultivar and growing region have significant effects on the properties of MPP
and this study established a foundation for their applications in food industry.

1. Introduction source can be used in the application of commercial pectin. An estimate


reported by the Food and Agriculture Organization (2016) noted a
Pectin is an important natural polymer and is largely used in foods, global volume of food wastage (approximately 1.6 billion tons per year),
pharmaceuticals and numbers of other industries. Because pectin has in which fruit waste accounts for the highest wastage rate. Therefore,
excellent characteristics in gelling, thickening and stabilizing, it is many researchers have attempted to recover pectin from various fruit
widely used as a food additive in the food industry. However, conven­ wastes and byproducts, such as kalamanzi (Citrus microcarpa) peel,
tional sources of pectin (citrus peel and apple pomace) have been unable pineapple waste, cocoa pods, mango and banana peels, jack fruit and
to meet the present or projected demand for human consumption, and passion fruit rind (Madhav & Pushpalatha, 2006; Petkowicz, Vriesmann,
the identification of different sources is of great interest due to the rising & Williams, 2017).
global demand for pectin (Müller-Maatsch et al., 2016). Mango is one of the most favored and valuable fruit throughout the
Although pectin is commonly observed in most plant tissues, the tropical and subtropical areas. Mango peel was reported to be an
number of sources that may be used for the commercial manufacture of excellent source of pectin with high DM and low protein content (Ajila,
pectin is limited. This limitation is observed because the pectin from Aalami, Leelavathi, & Prasada Rao, 2010). The physicochemical and
different sources showed different gelling abilities due to variations in functional properties of mango peel pectin are highly correlated with
numerous parameters (Fraeye et al., 2010). Therefore, a high level of their structures, including average molecular weight (AMW), degree of
pectin content in a certain source is not sufficient to determine that the methyl-esterification (DM), galacturonic acid (GalA) content, and

* Corresponding author. College of Food Sciences & Engineering, Hainan University, 58 People Road, Haikou, 570228, PR China.
** Corresponding author.
E-mail addresses: zhwm1979@163.com (W. Zhang), hmchen168@126.com (H. Chen).
1
These authors contributed equally to this study.

https://doi.org/10.1016/j.foodhyd.2020.105707
Received 5 October 2019; Received in revised form 30 December 2019; Accepted 22 January 2020
Available online 25 January 2020
0268-005X/© 2020 Elsevier Ltd. All rights reserved.
Z. Deng et al. Food Hydrocolloids 103 (2020) 105707

neutral sugar components. Several studies also indicated that pectin was flushed with pure nitrogen gas on an oil bath at 70 � C to remove the
yield, structure, and physicochemical properties were affected by the residual TFA. Then, the hydrolyzed samples were mixed with 1 mL of
cultivar, the degree of ripeness, the extraction method and other pro­ NaOH solution (0.3 mol/L) and 400 μL of 1-phenyl-3-methyl-5-pyrazo­
cesses (Geerkens et al., 2015; Yang, Mu & Ma, 2018; Yang, Nisar, Hou, lone methanolic solution (0.5 mol/L). Lactose was added as an inter­
Gou, Sun & Guo, 2018). Mango has been bred to hundreds of cultivars nal standard before derivatization. The resulting solution was extracted
around the world such as Okrong, Nuan Chan, Namdokmai No. 4 and twice by an equal volume of chloroform.
Narcissus in Thailand, Green and JinHwang in Southeast Asia, and Irwin The concentrations of neutral sugars and galacturonic acid were
mango in USA and China. However, comparative studies of the struc­ quantified on a HPLC system (2690, Waters Inc., Milford, MA, USA)
ture, emulsifying, and rheological properties have been rarely reported. equipped with a C18 column (250 mm � 4.6 mm, 4 μm, GraceSmart™,
The present study aimed to recover pectin from mango peel of eight Deerfield, IL, USA), with a flow rate of 1 mL/min. The mobile phase
cultivars from different growth regions (China, Burma, Australia, were composed of acetonitrile (A) and 100 mM of sodium phosphate
Thailand, and Vietnam) to determine the effects of cultivar and growth buffer (pH 6.7, B), using a gradient elution of 15% A and 85% B at 0–10
region on the basic compositions (AMW, compositional mono­ min, 17% A and 83% B at 10–30 min, 20% A and 80% B at 30–35 min,
saccharides, DM, protein content) and characteristics (FT-IR spectrum, and 15% A and 85% B at 35–40 min. The column temperature was 30 � C
thermodynamic characteristics, emulsifying and rheological behaviors) and injection volume of sample was 20 μL.
of MPP. Total protein content in pectin was determined by using a Bradford
protein assay kit (Pierce; Thermo Fisher Scientific Inc., USA). BSA was
2. Materials and methods used as the standard for calibration in the range of 0–750 mg/mL (Wang
et al., 2016).
2.1. Materials
2.4. Determination of AMW and DM
The twelve varieties of mangoes used in this study were purchased
from different countries as shown in Table 1. The AMW and its distribution of extracted pectin were determined by
high performance size exclusion chromatography (HPSEC) (Kermani
2.2. Extraction of pectin from mango peel et al., 2014). Pectin solution (0.1% w/v) was filtered through a 0.45-μm
filter and separated on an HPSEC system. Three Waters columns (Wa­
The pectin was extracted using the hot acid method (Wang et al., ters, Milford, MA) used in HPSEC studies were Ultrahydrogel 250, 1000,
2016). In brief, mango was harvested at preclimacteric mature green and 2000 with exclusion limits of 8 � 104, 4 � 106, and 1 � 107 g/mol,
stage washed with tap water to eliminate the bulk of sand and other respectively. Two detectors were present in line, a multi-angle light
inorganic materials and treated with ethylene (10 ppm) for 24 h. scattering detector (PN3621, Postnova analytics, Germany) and a
Thereafter, the mangoes were stored at 25 � C (4–7 days) until their refractive index detector (Shodex RI-101, Showa Denko, New York, NY,
maturity meeted the juicing requirement (identified by Prof. Weijun USA). The eluent was NaNO3 solution (0.05 M) together with NaN3 (0.5
Chen, Hainan University) and peeled. The obtained peel was washed g/L). The dn/dc values of all pectin samples were set at 0.138 mL/g
with ethanol (96%, v/v) for 20 min and with 70% ethanol for 5 min to when AMW was obtained. The AMW of pectin and Mw/Mn (Poly­
exclude low-molecular-weight sugars, amino acids, organic acids, inor­ dispersity index) were analyzed using ASTRA software (Wyatt Tech­
ganic salts, and inactivate enzymes. The residue was then oven dried at nology Corp., Santa Barbara, CA, USA; version 5.3.4.14).
40 � C, ground and passed through a 60-mesh sieve. A ratio of 1:30 (w/w) DM was calculated from the amount of Methanol (mol of Methanol
mango peels to HCl (pH 2) was selected to recover pectin at 85 � C for 4 h. per 100 mol of galacturonic acid) (Guillon & Thibault, 1987). Briefly,
The extraction solution was cooled in an ice bath for approximately 30 methoxy groups were removed from pectin by saponification in NaOH
min and precipitated by 4 vol of isopropyl alcohol. The pectin was dis­ solution (0.2 mol/L) in 50% isopropanol at 4 � C for 2 h. Methanol was
solved in the water and dialyzed against Milli-Q water at 25 � C using a determined by HPLC on an Aminex HPX 87H column (300 mm � 7.8
dialysis membrane with a 10 kDa cut-off until a constant conductivity mm; Bio-Rad Laboratories, Inc., Hercules, CA, USA) using succinic acid
value was reached. Then the pectin solution was freeze-dried. as the internal standard.

2.3. Determination of neutral sugars, galacturonic acid and protein 2.5. FT-IR analysis
contents
The pectin samples were blended with KBr powder (1:100, w/w) and
The neutral sugars and galacturonic acid contents of mango pectin pressed into tablets. FT-IR spectra were recorded on a FT-IR spectrom­
were quantified by using a HPLC method (Zhang et al., 2013). Ten eter (TENSOR 27, Bruker Optics, Ettlingen, Germany) at room temper­
milligrams of pectin powder were hydrolyzed firstly in trifluoroacetic ature over a range of 4000–400 cm 1 with a resolution of 4 cm 1.
acid (TFA) solution (5 mL, 2 mol/L) at 110 � C for 6 h. The hydrolysate
2.6. Determination of thermal properties
Table 1
Twelve cultivars of mango from different growth regions. The thermal properties of the samples were determined on a differ­
NO Mango cultivar Regions Countries
ential scanning calorimeter (DSC) (Q100, TA Instrument, USA) accord­
ing to the method of Khatkar et al. with minor modifications (Khatkar,
SuBP Sundiro Pathein Burma
Barak, & Mudgil, 2013). First, pectin (6.0 mg, dry basis) was weighed in
GoBP Golek Pathein Burma
GrVP Green Nha Trang Vietnam a standard aluminum pan and sealed immediately. An empty standard
JHVP JinHwang Nha Trang Vietnam aluminum pan was used as a reference. All measurements were per­
AuAP Australian Perth Australia formed in a dynamic inert nitrogen atmosphere (50 mL/min) at a
JHTP JinHwang Chiengmai Thailand
heating rate of 10 � C/min over a temperature range of 20� C–300 � C.
GrTP Green Chiengmai Thailand
NaTP Narcissus Chiengmai Thailand
IrCP Irwin Taiwan China 2.7. Emulsifying properties
GuCP Guifei Hainan China
GrCP Green Hainan China 2.7.1. Preparation of emulsions
JHCP JinHwang Hainan China
Mango pectin solution was prepared by dissolving 0.8 g of mango

2
Z. Deng et al. Food Hydrocolloids 103 (2020) 105707

pectin in 80 mL of demineralized water at room temperature with 3. Results and discussion


continuous stirring. To prevent solutions from deteriorating during
storage, sodium benzoate (0.1 g) and citric acid (0.1 g) were added, and 3.1. Yield, composition and protein contect of MPPs
the pH was adjusted to 3.5 using HCl solution (1 M). Then, demineral­
ized water and 10 g of soybean oil were added to achieve a final mass of As shown in Table 2, cultivar and growth environment play an
100 g. The mixture was then homogenized using an Ultra-Turrax device important role in the yield and composition of MPP. Among the 12
at a speed of 24,000 rpm for 30 min to form emulsion. MPPs, the yields are in the range of 7.63%–14.75%. JHVP and GrVP,
both of which were obtained from Vietnam, showed relatively higher
2.7.2. Surface tension determination yields (14.75% and 14.28%, respectively), indicating that the high
The surface tension was measured at 25 � C using a DCAT21 surface temperature and rainy climate and the black soil clay of Vietnam may be
tensiometer (Dataphysics Co., Germany). For surface tension measure­ favorable for the formation of pectin in mango peel (Table S1). GrVP,
ments, pectin solutions (1.0%, w/v) containing 0.02% sodium azide GrTP and GrCP obtained from Green cultivar showed higher yields than
were prepared by dissolving dried pectins in fresh Milli-Q water at pH that from other cultivars, simultaneously, the environment in which
3.5. The resistivity of distilled water which was used as a reference in these Green varieties grow is a tropical monsoon climate, and perhaps
this study was 10.3–18.5 MΩ. The measuring cup was placed in a climate factor has some obviously influence. However, the yield of AuAP
stainless steel jacket on a perpendicularly movable platform. A platinum cultivar is lower than that of Calypso cultivar mango which is also from
plate (24 mm � 10 mm � 0.1 mm) was gently placed at the bottom of a Australia (Oliveira et al., 2016). Comparing with other resources, such
sensitive electronic balance and immersed in the measured solutions by as pomegranate peel (8.5%) (Yang et al., 2018b), passion fruit rind
slowly raising the platform. The force (mN/m) required to detach the (2.9%–11.3%) (Yapo & Koffi, 2006) and banana peel (5.2%–12.2%)
plate from the surface was recorded as the dynamic surface tension. (Oliveira et al., 2016), MPP showed higher yeilds, which is close to apple
Finally, the surface tension was calculated based on balance-force pectin production (10.05%–16.68%) (Wang et al., 2014), suggesting
measurements by applying the appropriate correction factors (Chen, that mango peel is a potential and ideal pectin resource.
Fu, & Luo, 2016). Pectin is a heteropolysaccharide consisting of a chain structure of (1
→ 4)-linked α-D-galacturonic acid (GalA) units interrupted by the
2.7.3. ζ-potential determination insertion of (1 → 2)-linked L-rhamnopyranosyl residues and other
The ζ-potential of the emulsion was measured on a dynamic light neutral sugars (D-galactose and L-arabinose dominate) (Chen et al.,
scattering electrophoresis equipment (Zetasizer Nano S90, Malvern In­ 2016). The composition of pectin recovered from different plants are
struments, Malvern, UK). The pectin-stabilized emulsion was diluted different (Hua, Wang, Yang, Kang, & Zhang, 2015). GalA is the main
100 times with distilled water. Then, a pipette was used to capture 1 mL constituent unit of homogalacturonan (HG) and rhamnogalacturonan
of the resulting emulsion into a four-sided glass cuvette. Measurements (RG), and all MPP samples showed high GalA contents. As shown in
were performed in triplicate. Table 2, the GalA contents of most MPPs are close to or greater than 65%
(w/w), and only GrCP (49.6%) and GrTP (54.41%) are slightly lower,
2.7.4. Emulsion droplet size distribution which may be attributed to the high rate of side chain and the impurity
A laser diffraction instrument (Zetasizer Nano S90, Malvern In­ of arabinogalactan (Geerkens et al., 2015). Mannose (Man), ribose (Rib)
struments, Japan) was employed to determine the particle size distri­ and glucose (Glc) may originate from the remnant of soluble sugars
bution and PdI (Polydispersity Index) after the pectin-stabilized during the extraction process or non-pectic polysaccharides, such as
emulsion was diluted with distilled water at 1:100 ratios. The operation hemicellulose and cellulose (Wang et al., 2016). Continuous homo­
was performed in triplicate, and the average value was reported. galacturonan (HG) regions are ~100 residues long have found by Ralet
et al. (2008). In our study, HG ranges from 49.19%–80.12%, suggesting
2.7.5. Microscopic observation that most pectins are characterized by the predominance of long
To observe the microscopic droplets of emulsions, an inverted mi­ homogalacturonan regions. Furthermore, the molar ratio of Rha/GalA
croscope DCGV-20N (Aote Optical Instrument Co., Ltd., Chongqing reveals the contribution of Rha to the RG, and arabinan or arabinoga­
China) was used with a 10 � objective lens. lactan side chains formed RG-I, while (Gal þ Ara)/Rha exhibits the
length of branches attached to RG-I. As reported by Schols and Voragen
2.8. Rheological characterization (1996), the main constituent of pectin is the RG-I region if the molar
ratio of Rha/GalA is in the range of 0.05–1. As shown in Table 2, all
To compare the rheological properties of different pectin solutions at cultivars of mango pectins exhibited low Rha/GalA ratios
a concentration of 1.0% (m/v), a CMT rheometer (AR 550, TA, US) with (0.007–0.044), illustrating that pectin molecules have long HG domains,
a 60-mm parallel steel plate configuration was used (Wang, Chen, & Lü, especially GoBP and AuAP. A high value of (Ara þ Gal)/Rha in AuAP
2014). Flow curves with increasing shear rate (0.01–100 s 1) were reflects the branching extent of the RG-I domain. The protein content of
measured at 25 � C. Oscillatory measurements were used to determine GuCP (4.33 � 0.08%) was higher than that of other mango cultivars
the storage modulus (G0 ) and loss modulus (G00 ) of the samples. Strain (Table 2), whereas JHVP and JHCP have relative low protein contents
sweep (0.01–100% at 1 Hz) was applied to measure the linear visco­ (<2%) but higher than that of the commercial apple pectin (1.4%–1.6%)
elastic regions. The frequency dependence of G0 and G00 was conducted (Kravtchenko, Voragen, & Pilnik, 1992). The results demonstrated that
by a frequency sweep (1–100 Hz at 1% stain). the composition of MPPs were also cultivar- and growth
region-dependent.
2.9. Statistical analysis
3.2. DM and AMW analysis
All experimental results were expressed as the means � standard
deviation. Deviations were statistically analyzed using one-way ANOVA Methyl-esterification reaction mainly reacts with galacturonic acid
with SPSS 17.0 (SPSS Inc., Chicago, Illinois, USA) and performed using on the main chain of pectin. DM corresponds to the percentage of
Tukey’s Test (p < 0.05). The figures were plotted using Origin 9.0 carboxyl group esterified with methanol and plays an important role in
(OriginLab Inc., Northampton, Massachusetts, USA). the physicochemical properties, especially gelation behavior of pectin
(Chen, Fu, Abbasi, & Luo, 2015). As shown in Table 2, all the MPPs
belong to high - Methoxyled pectin (DM > 50%), and SuBP showed the
highest DM value (71.80%). Meanwhile, the DM value of GuCP was

3
Z. Deng et al. Food Hydrocolloids 103 (2020) 105707

Table 2
Yields and chemical composition of mango peel pectin.
Sample Yield GalA Rha (%) Ara (%) Gal (%) Glc (%) Man (%) Fuc (%) Rha/GalA RG-I HG DM (%) Protein
(%)a (%)b (%)

SuBP 9.23 � 63.00 � 0.60 � 1.65 � 18.70 � 14.1 � 0.00 � 0.00 � 0.010 � 21.55 � 62.40 � 66.16 � 3.06 �
0.20bc 0.06h 0.01d 0.03g 0.07d 0.05d 0.00c 0.00d 0.000c 0.12e 0.23h 2.13b 0.05c
GoBP 11.0 � 80.63 � 0.51 � 1.20 � 16.08 � 0.00 � 0.07 � 0.00 � 0.044 � 18.30 � 80.12 � 71.80 � 3.10 �
0.54b 0.07a 0.02e 0.02i 0.12h 0.00j 0.01b 0.00d 0.002a 0.04g 0.36a 0.47a 0.86c
GrVP 14.3 � 66.25 � 0.99 � 3.35 � 27.47 � 0.00 � 0.08 � 0.00 � 0.006 � 32.92 � 65.20 � 63.81 � 4.17 �
0.24a 0.12f 0.02a 0.02c 0.15a 0.00j 0.00b 0.00d 0.000e 0.18a 0.12f 1.81c 0.08ab
JHVP 14.6 � 64.46 � 0.74 � 2.47 � 22.06 � 8.15 � 0.07 � 0.21 � 0.016 � 26.01 � 63.72 � 61.79 � 3.78 �
0.35a 0.05g 0.03b 0.08d 0.06b 0.04e 0.01b 0.01b 0.001b 0.09c 0.10g 1.48f 0.11bc
AuAP 8.99 � 76.97 � 0.45 � 2.25 � 17.00 � 1.05 � 0.09 � 0.18 � 0.006 � 20.15 � 76.52 � 71.07 � 2.23 �
0.34bc 0.07b 0.01f 0.09e 0.02f 0.01i 0.00b 0.01c 0.000e 0.05b 0.11b 1.58ab 0.03e
JHTP 7.63 � 73.33 � 0.65 � 3.55 � 17.27 � 3.18 � 0.12 � 0.00 � 0.009 � 22.12 � 72.68 � 65.85 � 2.84 �
0.54c 0.04c 0.01c 0.07b 0.07e 0.02c 0.01a 0.00d 0.000c 0.17d 0.18d 1.85b 0.05d
GrTP 13.4 � 54.41 � 0.43 � 1.41 � 12.53 � 29.8 � 0.08 � 0.15 � 0.008 � 14.80 � 53.98 � 60.38 � 3.76 �
0.47ab 0.09c 0.02f 0.00h 0.05i 0.12b 0.01b 0.01b 0.000d 0.05c 0.09i 1.44a 0.19bc
NaTP 12.4 � 62.95 � 0.46 � 1.22 � 15.71 � 17.7 � 0.08 � 0.18 � 0.007 � 17.85 � 62.49 � 70.60 � 1.95 �
0.45ab 0.13i 0.01c 0.01i 0.05bc 0.02c 0.00b 0.01b 0.001de 0.10h 0.19h 2.25de 0.09f
IrCP 12.2 � 76.18 � 0.55 � 1.67 � 12.49 � 7.48 � 0.00 � 0.00 � 0.007 � 15.26 � 75.63 � 68.72 � 2.83 �
0.24ab 0.11a 0.02e 0.01g 0.03i 0.00f 0.00c 0.00d 0.000e 0.08i 0.25c 1.53b 0.26d
GuCP 11.4 � 72.90 � 0.75 � 2.27 � 16.47 � 5.17 � 0.12 � 0.26 � 0.010 � 20.24 � 72.15 � 53.64 � 4.33 �
0.42b 0.10d 0.07b 0.03e 0.02g 0.01g 0.01a 0.02a 0.001c 0.07f 0.24b 1.61e 0.08a
GrCP 14.7 � 49.60 � 0.41 � 2.21 � 11.95 � 34.8 � 0.00 � 0.00 � 0.008 � 14.98 � 49.19 � 59.70 � 3.95 �
0.22a 0.07j 0.01f 0.02ef 0.01j 0.14a 0.00c 0.00d 0.000d 0.03j 0.10j 1.04d 0.10b
JHCP 11.5 � 70.88 � 0.75 � 5.08 � 20.00 � 1.56 � 0.10 � 0.07 � 0.010 � 26.58 � 70.13 � 69.04 � 1.86 �
0.27b 0.15e 0.03b 0.08a 0.08c 0.03h 0.01ab 0.00c 0.000c 0.06b 0.22e 0.60ab 0.05f
a
The data are averages of three measurements with standard deviation. Data with different letters (~
aj) in a same column were significantly different (p < 0.05).
b
GalA, galacturonic acid; Rha, rhamnose; Ara, arabinose; Gal, galactose; Glc, glucose; Man, mannose; Fuc, fructose; HG, homogalacturonan; RG, rhamnoga­
lacturonan; DM, degree of methyl-esterification.

approximately 53.64%, much lower than that of JHCP (69.04%), indi­ AMW of most mango pectins is larger than that of other fruits, such as
cating that the impact of cultivar is obviously. However, JHTP, JHVP grape pomace (110–205 kDa) (Minjares-Fuentes et al., 2014). The
and JHCP, all of which were extracted from cultivar JinHwang mango, different distributions of pectins could be compared through the poly­
showed no significant (p < 0.05) difference in DM. In agreement with dispersity index (AMW/Mn), and higher AMW/Mn values represent a
our results, previous studies have shown that MPPs from Am�elior�ee and wider molar mass distribution (Rogo�si�c, Mencer, & Gomzi, 1996). In our
Keitt cultivars are both high DM pectin (Wang et al., 2016). Pectin with study, the AMW/Mn of all the MPPs ranged from 1.324 to 2.357, which
high DM value appears to be more stable against endo-α-1,4-poly­ shows that the distribution of molecular weight was relative narrow.
galacturonase and high temperature (Liu, Cao, Huang, Cai, & Yao, These values indicated that MPPs are highly homogeneous poly­
2010), which is suitable for high-sugar food products. saccharides with concentrated molecular-weight distributions (Yang,
The AMW of pectin relies on both its main chain and side chain Mu, & Ma, 2018a).
(Sousa, Nielsen, Armagan, Larsen, & Sørensen, 2015). As shown in
Table 3, the significant differences exhibit in the molecular-weight dis­
tribution of the pectins and the AMW of GoBP is higher than those of the 3.3. FTIR characterization
other MPPs. Correspondingly, the Mn of GoBP was also the highest,
showed significant (p < 0.05) difference than other cultivars. The Mn The FTIR spectra of the 12 MPPs are illustrated in Fig. 1. It can be
values of MPPs are in the range of 200–800 kDa, which were lower than observed that these mango pectins contain almost the same functional
that extracted from Keitt mango (Kermani, Shpigelman, Pham, Loey, & groups and structures. Fig. 1A shows that the band at approximately
Hendrickx, 2015) and Kent mango (Kermani et al., 2014). However, the 3400 cm 1 was attributed to the hydroxyl groups. An absorption at
2930 cm 1 was caused by C–H stretching involving CH, CH2 and CH3
bending vibrations. The signal at 1750 cm 1 shows the esterified
Table 3
carboxyl groups (COCH3) and is overlapping with the free carboxyl
The weight-average molecular weight (AMw), AMw/Mn of different varieties of
mango peel pectins.a.
groups (COOH), the signal at 1640 cm 1 shows the carboxylate group.
The DM value of pectin can be determined according to the ratio of the
Samples AMW (kDa)b Mn (kDa) AMW/Mn
peak area at 1750 cm 1 to the overall peak areas of 1750 and 1640 cm 1
SuBP 266�2h
115�1 h
2.31 � 0.04a (Vriesmann & Petkowicz, 2009). Fig. 1 indicates that all MMPs are high
GoBP 1589�5a 911�4a 1.74 � 0.01d methoxyl pectins, which agrees with the results of DM determination
GrVP 569�3c 362�2c 1.57 � 0.00f
JHVP 382�1e 233�2e 1.63 � 0.01b
(Table 2). In addition, most of the peaks below 1100 cm 1 are related to
AuAP 438�3d 257�2d 1.70 � 0.01f the monosaccharides. The peaks in the range of 1010–1100 cm 1 indi­
JHTP 201�1i 116�1h 1.74 � 0.02d cate the presence of pyranose and the furanose ring. Furthermore, it has
GrTP 271�2g 172�2f 1.58 � 0.01g been reported that the structure of grapefruit peel pectin was similar to
NaTP 265�1h 133�1cd 2.00 � 0.02b
mango pectins with similar functional groups and monosaccharide types
IrCP 760�4b 572�3b 1.33 � 0.00g
GuCP 143�1j 108�1i 1.33 � 0.01g (Wang et al., 2017).
GrCP 297�1f 126�1g 2.37 � 0.02a FTIR spectroscopy is also a common method for investigating
JHCP 509�4b 271 � 1.65c 1.88 � 0.01c protein-carbohydrate systemk, as there are several readily identifiable
a
The data are averages of three measurements with standard deviation. Data regions of the mid-infrared spectrum where the chemical fingerprints of
with different letters (~
aj) in a same column were significantly different (p < carbohydrates and proteins do not overlap significantly. The charac­
0.05). teristic bands found in the infrared spectra of proteins include the amide
b
AMW, average molecular weight. I and II bands at 1700–1600 cm 1 and 1600–1500 cm 1, respectively

4
Z. Deng et al. Food Hydrocolloids 103 (2020) 105707

Fig. 1. Fourier transform infrared spectra (FT-IR) of mango peel pectin obtained from different varieties of mango.

(Wang, Bao, & Chen, 2013). The amide I band is due to CO stretching conformation of the galacturonan ring. The intensity of the heat flow of
vibrations, while the amide II band is due to C–N stretching vibration as JHVP was slightly higher than that of other pectin samples, revealing
well as N–H bending vibration. Some studies have reported that that JHVP had lower thermal stability (Osorio, Carriazo, & Barbosa,
protein-carbohydrate conjugates had strong absorption peaks at 2011). A sharper exothermic peak means that the sample has a narrow
1065.43, 1066.22, 1520.02, 1581.41, 1382.86 and 1388.59 cm 1 melting range, a concentrated distribution of molecular weight and an
(Rozenberg, Lansky, Shoham, & Shoham, 2019). As shown in Fig. 1, ordered molecular arrangement. As shown in Table 3, the low ΔHg
there were strong peaks in these several wavelength bands, and there values of GrCP and GrTP also revealed that their molecular arrangement
were obvious absorption peaks of covalent bonding in the vicinity of was more orderly. In addition, the high Tg and ΔHg values indicated that
3400 and 2930 cm 1, which is related to the telescopic changes in –NH, GrVP and AuAP have broad melting ranges and good thermal stability,
–OH, and –CH. which showed that highly branched side chains such as (Gal þ Ara)/Rha
and RG-1, may enhance the thermal stability of MPPs. The differences
3.4. Thermal analysis among the Green variety of MPPs also verified the growing environment
had a great impact on the stucture and properties of MPP.
DSC was performed to evaluate the thermal behaviors of the MPPs in
the temperature range of 20� C–300 � C. As shown in Fig. 2A, an endo­ 3.5. Physical stability of MPPs-stabilized MCT-water emulsions
thermic peak and an exothermic peak were observed in the DSC ther­
mograms of all MPPs. As shown in Table 4, the parameters including The physical stability of the emulsions stabilized by MPPs were
onset temperature (To), melting temperature (Tm), melting enthalpy evaluated by surface tension (Fig. 2B), ξ-potential (Fig. 2C), particle size
(ΔHm), degradation temperature (Tg) and degradation enthalpy (ΔHg) of (Table 3) and microscopic observation (Fig. 3). To compare the surface
the two peaks were evaluated. The Tm and Tg values of MPPs varied behavious of MPPs, the surface tension was determined at the air/water
from 93.47 � C to 126.88 � C and 248.08 � C–263.53 � C, respectively. This interface at room temperature with pectin solutions (0.5%, w/v). As a
result is consistent with the findings of previous studies (Wang et al., reference, the surface tension of the distilled water was measured. As
2017). Since the samples was not fully dried before DSC, therefore, the shown in Fig. 2B, the surface tension of MPPs ranged from 51.82 to
peak around 100 � C appeared is likely because release of water from the 58.50 mN/m, which is much lower than that of the reference (71.85
sample, and increasing the temperature will accelerate water uptake mN/m), indicating that MPPs were all surface-active. In addition, MPP
from the environment due to higher mobility of water molecules and was more effective to reduce the surface tension than the commercial
easier transport to the particle surface (Einhorn-Stoll, 2018). The pectins (citrus peel and apple pomace pectin). According to the study
melting temperature of SuBP and AuAP was above 120 � C, indicating reported by Lutz, Aserin, Wicker, and Garti (2009), the surface tension
that the larger transition factor maybe the more stable 4C1 chair of apple pectin solution (0.5 wt%, MW 30–100 kDa, 70–75% random

5
Z. Deng et al. Food Hydrocolloids 103 (2020) 105707

Fig. 2. (A) Represents DSC thermograms of different varieties of mango peel pectins, (B) represents surface tension and (C) represents ξ-potential of different va­
rieties of mango peel pectin emulsions, (D) represents flow behaviors of different varieties of mango peel pectin solution.

methylesterification) was 63 mN/m. And the surface tension of citrus


Table 4
peel pectin (galacturonic acid > 74.0% and methoxy group > 6.7%) as
The parameters determined by DSC.
reported by Wu, Eskin, Cui, and Pokharel (2015) was approximate 67
Samplesa Tm (� C)b ΔHm (J/g) Tg (� C) ΔHg (J/g) To (� C) mN/m under the concentration of 0.5% (w/w). While, as a typical
SuBP 124.47 � 154.4 � 257.84 � 99.50 � 245.59 � amphoteric substance, SPI solution (2%, w/w) with a surface tension of
2.23a 1.7cd 2.65ab 1.27c 3.24b 44.78 � 1.38 mN/m was more surface-active than all MPP samples
GoBP 93.470 � 477.9 � 263.53 � 55.78 � 253.03 � tested in this study (Fig. 2B). Pectin is a group of polysaccharides rich in
1.44c 4.2a 3.06a 0.55g 4.04a
GrVP 125.60 � 286.4 � 258.92 � 112.1 � 244.92 �
galacturonic acid and are mostly hydrophilic polymers, which do not
3.86a 2.5bc 1.23ab 0.92a 2.23b exhibit significant surface activity. In general, it could be observed that
JHVP 116.35 � 433.5 � 252.23 � 59.08 � 242.29 � the reduction of surface tension caused by pectin was related to the
3.13ab 3.2ab 1.47b 0.53f 1.65bc proteinaceous moiety. While, the protein content was not the only factor
AuAP 126.88 201.9 262.10 101.6 � 247.72 �
� � �
determining the surface activity. The research reported by Huang,
1.67a 2.7c 2.51a 0.74b 1.98b
JHTP 107.04 � 221.9 � 248.08 � 73.75 � 236.45 � Kakuda, and Cui (2001) showed that the surface activity of many
2.78bc 3.6c 1.77bc 1.28d 2.38d polysaccharides included pectin does not correlate with their overall
GrTP 112.51 � 304.1 � 255.62 � 22.48 � 243.63 � protein content, but instead with the content of amphiphilic
1.55b 3.4bc 2.48b 0.63i 1.87b protein-polysaccharides entities. Specific pectins such as sugar beet
NaTP 117.01 346.6 258.10 38.95 � 245.96 �
pectin have a pronounced surface-tension-lowering ability than com­
� � �
1.31ab 2.9b 2.84ab 1.56h 1.77ab
IrCP 109.20 � 425.1 � 259.43 � 59.67 � 246.73 � mercial pectins because of high protein content and high DM value
2.34b 2.1ab 1.49ab 1.76f 1.49b (Chen et al., 2016; Dea & Madden, 1986). Modified pectins with octenyl
GuCP 109.08 � 372.0 � 250.75 � 58.87 � 239.91 � succinic anhydride (Chen, Fu, & Luo, 2015), bovine serum album (Chen
1.56b 3.8b 2.30bc 1.21f 2.01c
et al., 2018) and depolymerization (Akhtar, Dickinson, Mazoyer, &
GrCP 119.06 � 289.7 � 255.54 � 20.94 � 240.5 �
3.09ab 2.0bc 2.03b 0.38j 1.27c Langendorff, 2002) also showed more surface activity than the native
JHCP 104.14 � 460.1 � 251.18 � 70.08 � 240.28 � pectins. Probably due to the high protein content (4.33 � 0.08%) and
1.40bc 3.1a 2.45b 2.35e 1.31c low DM value (53.64 � 1.61%), the surface tension of the GuCP solution
a
The data are averages of three measurements with standard deviation. Data (51.82 � 0.34%) was the lowest among the MPP samples, and its value
with different letters (~
ai) in a same column were significantly different (p < was lower than that of sugar beet pectin solution (53.4 mN/m in con­
0.05). centrations of 0.5% w/v) (Yapo, Wathelet, & Paquot, 2007). Meanwhile,
b
To, onset temperature; Tm, melting temperature; ΔHm, melting enthalpy; Tg, the surface tension of the IrCP solution was 58.50 mN/m, which was
degradation temperature; ΔHg, degradation enthalpy. significantly higher than that of the other samples. The proportion of the
side chains (Rha/GalA) also palyed a positive role in reducing the

6
Z. Deng et al. Food Hydrocolloids 103 (2020) 105707

Fig. 3. Microscopic observation of different varieties of mango peel pectin emulsions.

surface tension of GuCP solution (0.007) when compared with that of


the IrCP sample (0.010). And it seems that the surface activity of pectin Table 5
The particle size distribution and PdI of emusions.a.
has a direct relation with its AWM (143 and 760 kDa for GuCP and IrCP,
b
respectively), which was consistent with the conclusion obtained by Samples Particle size (μm) PdI
Akhtar et al. (2002). However, the GoBP with the highest contents of SuBP 3.95 � 0.24a
0.39 � 0.10a
GalA (80.63%), HG (80.12%), DM (71.80%) and AMW (1589 kDa) gave GoBP 2.53 � 0.23e 0.26 � 0.06ab
rise to a higher surface tension than GuCP, showing that protein fraction, GrVP 2.79 � 0.18e 0.20 � 0.04b
JHVP 2.15 � 0.22f 0.09 � 0.03c
molecular weight, DM and internal charge distribution might all AuAP 2.07 � 0.08f 0.23 � 0.03b
contribute to the surface tension of emulsions (Baldino, Mileti, Lupi, & JHTP 2.56 � 0.10e 0.20 � 0.04b
Gabriele, 2018). GrTP 3.32 � 0.20c 0.15 � 0.05bc
Because the electrostatic repulsion is one of several factors contrib­ NaTP 3.32 � 0.04c 0.29 � 0.03ab
IrCP 3.78 � 0.19b 0.16 � 0.03bc
uting to emulsion stabilization against coalescence, ξ-potential was
GuCP 3.05 � 0.12cd 0.18 � 0.06bc
measured to evaluate the electrostatic effect on the stability of the GrCP 3.29 � 0.25c 0.20 � 0.07b
emulsion. As shown in Fig. 2C, all the pectin-stabilized emulsions pre­ JHCP 2.86 � 0.28e 0.23 � 0.05b
sented negative charges at pH 3.5. It was reported that acidic a
The data are averages of three measurements with standard deviation. Data
polysaccharides-stabilized emulsions with low zeta-potentials (þ30 ~ with different letters (~
af) in a same column were significantly different (p <
30 mV) tend to coagulate or flocculate and may make it difficult to 0.05).
maintain a stable emulsion system (Jiao et al., 2019; Nakauma et al., b
PdI, polydispersity index.
2008). In this study the zeta-potentials of all MPP samples varying from
5.58 to - 13.26 mV, are within �30 mV because of high DM values, but addition, JHVP-stabilized emulsion has a low PdI value (0.09), sug­
relatively stable emulsions were also obtained, indicating that the gesting that the emulsion prepared by JHVP may be more stable. It can
electrostatic repulsion was not a critical factor on stabilizing the be seen from the microscope images, the AuAP-stabilized emulsion
emulsion. exhibited the smallest droplet size and most uniform distribution among
The particle size distribution is an important feature defining the all the emulsions. Simultaneously, AuAP and JHVP either has a longer
stability of emulsion (Kermani et al., 2015). The PdI value gives an (Gal þ Ara)/Rha domain or RG-1 domain, suggesting that the lateral
important parameter to better explain the particle sizes of an emulsion. chains of pectin, especially arabinose and galactose, mainly affecting the
As shown in Table 5, the PdI values of all MPPs-stabilized emulsions aggregation or coalescence of the oil droplet by steric effects.
were lower than 0.2, suggesting that all emulsion samples are homo­
geneous (Silva et al., 2011). Combined particle size with microscopic
3.6. Rheological analysis
observation, the emulsions prepared with SuBP (3.95 � 0.24 μm) and
IrCP (3.78 � 0.19 μm) showed larger droplet sizes. This result was in line
Fig. 2D showed the flowing curves of MPP solutions (1%, w/v) under
with the general trend of surface tension as discussed above. Meanwhile,
steady-shear conditions, which was characterized by a plot of apparent
the highest PdI value (0.39) was also observed in the SuBP-stabilized
viscosity versus shear rate. The apparent viscosity of GoBP, JHTP, JHVP,
emulsion, indicating that the SuBP-stabilized emulsion system was
GrVP, IrCP, GuCP and NaTP decreased with the increase of shear stress
fragile and unstable compared with other emulsions. JHVP- and
(0.01–100 s 1). The characteristic of shear thinning suggested that the
AuAP-stabilized emulsions have a slight smaller particle size. In
seven pectins mentioned above had non-Newtonian pseudoplastic

7
Z. Deng et al. Food Hydrocolloids 103 (2020) 105707

behavior. Similar behavior was observed on pectins from apple pomace stain to determine the dependence of G0 and G00 on the frequency (Fig. 5).
(Zhang et al., 2013), sugar beet pulp (Chen, Fu, & Luo, 2015) and Most MPPs (except GrVP, AuAP and GrCP) showed typical behaviour of a
grapefruit peel (Wang et al., 2016). While, the solutions of AuAP, GrTP, polysaccharide solution at low frequency where the G0 is greater than
GuCP and GrCP were close to Newton fluids because their apparent the G00 . The viscous behavior of most pectins gradually dominated the
viscosity almost unchanged during the increase of shear rate. As shown elastic behavior (G00 > G0 ) as the frequency increased. However, storage
in Fig. 2D, the solutions of GoBP, JHTP, JHVP and GrVP showed relative modulus of SuBP, GuCP, JHCP and NaTP was predominant, which
higher viscosity at the beginning than other samples (<0.5 Pa s), which revealed that their elastic properties were dominant, possibly because
mainly related to the AMW (Kim, Teng, & Wicker, 2005). According to the inter chain entanglements of pectin molecular (Piermaria, Canal, &
the correlation analysis (Fig. S1), AMW, GalA and RG-I were beneficial Abraham, 2008). And it may also be related to their breeds, which grow
to increase the viscosity of mango pectins. While, Glc and Fuc in the in higher altitudes. The crossover point of G0 and G00 of GrVP and GrCP
lateral chain had a negative impact on viscosity of mango pectins. appeared to be the fastest. It can attributed to their geographical envi­
Compared with the viscosity of commercial pectins (<1.0 Pa s at a ronment of the coast and tight structural organization, and the pectins
concentration of 2.0 wt%) as reported by Wang et al. (2014), there was with viscous liquid behavior could be used as a thickener in, for
potential values for GoBP (2.19 Pa s at a concentration of 1.0 wt%) and example, baking jams. The DM value of GuCP was lower than that of the
JHTP (1.51 Pa s at a concentration of 1.0 wt%) to be used as a thickener other pectin solutions, but the elastic characteristic of the GuCP solution
or stabilizer in the food industry by providing higher viscosity. was dominant, similar to the previous study (Min, Lim, Ko, Lee, Lee, &
Strain sweeps at a constant frequency of 1 Hz were carried out with Lee, 2011). On the other hand, the homogalacturonan (HG) region is
deformation values of 0.1–1000% to determine the linear viscoelastic believed to be mainly responsible for the gelling properties (Fig. S1).
region of MPPs. As shown in Fig. 4, the G0 of all pectins was slightly Therefore, lower gelation capability of pectin poor in HG can be ex­
higher than G00 at the beginning of strain sweeping with frequency of 1 pected, which was consistent with GrCP in our study. According to the
Hz, which indicated that there were more intermolecular entanglement above results, it was believed that the viscoelastic characteristics of
and polymer chain interactions. A good indication of the viscoelastic pectin might be determined by the high GalA content and low Glc
behavior that can be used to confirm the beginning of the elastic content. The elastic characteristics of JHTP, NaTP, and GuCP are all
behavior (Piermaria, de la Canal, & Abraham, 2008). The crossover prominent but have lower AMW in our study, which was also confirmed
points of G0 and G00 of NaTP appeared to be the fastest (0.5%), while by Kermani et al. (2014).
GrTP appeared to be the slowest (approximately 200%). According to
the correlation analysis, the main chain structures such as GalA, HG and 4. Conclusions
RG-I, were more conducive to gel-forming of mango pectins. While, the
gel-forming ability of mango pectin decreased with the increase of the Among the 12 MPPs, the yields are in the range of 7.63%–14.75%.
Glc content. JHVP and GrVP, both of which were obtained from Vietnam, showed
Frequency sweeps were performed over a range of 1–100 Hz at 1% relatively higher yields. GrVP, GrTP and GrCP obtained from Green

Fig. 4. Strain sweeps at 1 HZ frequency of modulus G0 and G00 of different varieties of mango peel pectin.

8
Z. Deng et al. Food Hydrocolloids 103 (2020) 105707

Fig. 5. Frequency sweeps at 1% strain of modulus G0 and G00 of different varieties of mango peel pectin.

cultivar also showed relative higher yields than that from other culti­ Declaration of competing interest
vars. These results showed that cultivar and growth environment played
an important role in the yield and composition of MPP. In addition, all The authors declare that they have no conflict of interest.
the MPPs belong to high - methoxyled pectin (DM > 50%), and SuBP
showed the highest DM value (71.80%), which was also confirmed by CRediT authorship contribution statement
FT-IR characterization. Meanwhile, the DM value of GuCP was approx­
imately 53.64%, much lower than that of JHCP (69.04%), indicating Zhuodan Deng: Writing - original draft. Yonggui Pan: Writing -
that the impact of cultivar is obviously. On the other hand, MPPs are original draft. Wenxue Chen: Data curation. Weijun Chen: Investiga­
highly homogeneous polysaccharides with concentrated molecular- tion. Yonghuan Yun: Software. Qiuping Zhong: Methodology. Wei­
weight distributions according to the results of AMW/Mn min Zhang: Funding acquisition. Haiming Chen: Writing - review &
(1.324–2.357). The high Tg and ΔHg values indicated that GrVP and editing.
AuAP have broad melting ranges and good thermal stability. The
physical stability of the emulsions stabilized by MPPs were evaluated by Acknowledgments
surface tension, ξ-potential, particle size and microscopic observation.
The surface tension of MPPs ranged from 51.82 to 58.50 mN/m, which is This study was funded by grants from the Natural Science Founda­
much lower than that of the reference (71.85 mN/m), indicating that tion of Hainan Province of China (2019RC120), the National Natural
MPPs were all surface-active. In addition, MPP was more effective to Science Foundation of China (31801949), and the Hainan University
reduce the surface tension than the commercial pectins (citrus peel and Start-up Scientific Research Projects of China (kyqd1630; kyqd1551).
apple pomace pectin). Combined droplet size with microscopic obser­
vation, the emulsions prepared with SuBP (3.95 � 0.24 μm) and IrCP Appendix A. Supplementary data
(3.78 � 0.19 μm) showed larger droplet size. The characteristic of shear
thinning of GoBP, JHTP, JHVP, GrVP, IrCP, GuCP and NaTP suggested Supplementary data to this article can be found online at https://doi.
that the seven pectins had non-Newtonian pseudoplastic behavior. Ac­ org/10.1016/j.foodhyd.2020.105707.
cording to the correlation analysis, AMW, GalA and RG-I were beneficial
to increase the viscosity of mango pectins. While, Glc and Fuc in the References
lateral chain had a negative impact on viscosity of mango pectins.
Compared with the viscosity of commercial pectins, there was potential Ajila, C. M., Aalami, M., Leelavathi, K., & Prasada Rao, U. J. S. (2010). Mango peel
powder: A potential source of antioxidant and dietary fiber in macaroni
values for GoBP and JHTP to be used as a thickener or stabilizer in the preparations. Innovative Food Science & Emerging Technologies, 11, 219–224.
food industry by providing higher viscosity. Akhtar, M., Dickinson, E., Mazoyer, J., & Langendorff, V. (2002). Emulsion stabilizing
properties of depolymerized pectin. Food Hydrocolloids, 16, 249–256.

9
Z. Deng et al. Food Hydrocolloids 103 (2020) 105707

Baldino, N., Mileti, O., Lupi, F. R., & Gabriele, D. (2018). Rheological surface properties Müller-Maatsch, J., Bencivenni, M., Caligiani, A., Caligiani, A., Tedeschi, T.,
of commercial citrus pectins at different pH and concentration. Lebensmittel- Bruggeman, G., et al. (2016). Pectin content and composition from different food
Wissenschaft & Technologie, 93, 124–130. waste streams. Food Chemistry, 201, 37–45.
Chen, H. M., Fu, X., Abbasi, A. M., & Luo, Z. G. (2015a). Preparation of environment- Nakauma, M., Funami, T., Noda, S., Ishihara, S., Al-Assaf, S., Nishinari, K., et al. (2008).
friendly pectin from sugar beet pulp and assessment of its emulsifying capacity. Comparison of sugar beet pectin, soybean soluble polysaccharide, and gum Arabic as
International Journal of Food Science and Technology, 50(6), 1324–1330. food emulsifiers. 1. Effect of concentration, pH, and salts on the emulsifying
Chen, H. M., Fu, X., & Luo, Z. G. (2015b). Properties and extraction of pectin-enriched properties. Food Hydrocolloids, 22, 1254–1267.
materials from sugar beet pulp by ultrasonic-assisted treatment combined with Oliveira, T.�I. S., Rosa, M. F., Cavalcante, F. L., Pereira, P. H., Moates, G. K., Wellner, N.,
subcritical water. Food Chemistry, 168, 302–310. et al. (2016). Optimization of pectin extraction from banana peels with citric acid by
Chen, H. M., Fu, X., & Luo, Z. G. (2016). Effect of molecular structure on emulsifying using response surface methodology. Food Chemistry, 198, 113–118.
properties of sugar beet pulp pectin. Food Hydrocolloids, 54, 99–106. Osorio, C., Carriazo, J. G., & Barbosa, H. (2011). Thermal and structural study of guava
Chen, H., Ji, A., Qiu, S., Liu, Y., Zhu, Q., & Yin, L. (2018). Covalent conjugation of bovine (Psidium guajava L) powders obtained by two dehydration methods. Quimica Nova,
serum album and sugar beet pectin through Maillard reaction/laccase catalysis to 34, 636–U684.
improve the emulsifying properties. Food Hydrocolloids, 76, 173–183. Petkowicz, C. L. O., Vriesmann, L. C., & Williams, P. A. (2017). Pectins from food waste:
Dea, I. C. M., & Madden, J. K. (1986). Acetylated pectic polysaccharides of sugar beet. Extraction, characterization and properties of watermelon rind pectin. Food
Food Hydrocolloids, 1, 71–88. Hydrocolloids, 65, 57–67.
Einhorn-Stoll, U. (2018). Pectin-water interactions in foods–From powder to gel. Food Piermaria, J. A., Canal, M. L. D. L., & Abraham, A. G. (2008). Gelling properties of
Hydrocolloids, 78, 109–119. kefiran, a food-grade polysaccharide obtained from kefir grain. Food Hydrocolloids,
Fraeye, I., Colle, I., Vandevenne, E., Duvetter, T., Buggenhout, S. V., Moldenaers, P., et al. 22(8), 1520–1527.
(2010). Influence of pectin structure on texture of pectin–calcium gels. Innovative Rogo�si�c, M., Mencer, H. J., & Gomzi, Z. (1996). Polydispersity index and molecular
Food Science & Emerging Technologies, 11, 0-409. weight distributions of polymers. European Polymer Journal, 32, 1337–1344.
Geerkens, C. H., Nagel, A., Just, K. M., Miller-Rostek, P., Kammerer, D. R., Rozenberg, M., Lansky, S., Shoham, Y., & Shoham, G. (2019). Spectroscopic FTIR and
Schweiggert, R. M., et al. (2015). Mango pectin quality as influenced by cultivar, NMR study of the interactions of sugars with proteins. Spectrochimica Acta Part A:
ripeness, peel particle size, blanching, drying, and irradiation. Food Hydrocolloids, Molecular and Biomolecular Spectroscopy, 222, 116861.
51, 241–251. Schols, H. A., & Voragen, A. G. J. (1996). Complex Pectins: Structure elucidation using
Guillon, F., & Thibault, J. F. (1987). Characterization and oxidative cross-linking of sugar enzymes. Progress in Biotechnology, 14, 3–19.
beet pectins after mild acid hydrolysis and arabanases and galactanases degradation. Silva, M. D. S., Cocenza, D. S., Grillo, R., Melo, N. F. S., Tonello, P. S., Oliveira, L. C., et al.
Food Hydrocolloids, 1, 547–549. (2011). Paraquat-loaded alginate/chitosan nanoparticles: Preparation,
Huang, X., Kakuda, Y., & Cui, W. (2001). Hydrocolloids in emulsions: Particle size characterization and soil sorption studies. Journal of Hazardous Materials, 190,
distribution and interfacial activity. Food Hydrocolloids, 15, 533–542. 366–374.
Hua, X., Wang, K., Yang, R., Kang, J., & Zhang, J. (2015). Rheological properties of Sousa, A. G., Nielsen, H. L., Armagan, I., Larsen, J., & Sørensen, S. O. (2015). The impact
natural low-methoxyl pectin extracted from sunflower head. Food Hydrocolloids, 44, of rhamnogalacturonan-I side chain monosaccharides on the rheological properties
122–128. of citrus pectin. Food Hydrocolloids, 47, 130–139.
Jiao, W., Chen, W., Mei, Y., Yun, Y., Wang, B., Zhong, Q., et al. (2019). Effects of Vriesmann, L. C., & Petkowicz, G. L. O. (2009). Polysaccharides from the pulp of
molecular weight and guluronic acid/mannuronic acid ratio on the rheological cupuassu (Theobroma grandiflorum): Structural characterization of a pectic fraction.
behavior and stabilizing property of sodium alginate. Molecules, 24(23), 4374. Carbohydrate Polymers, 77, 72–79.
Kermani, Z. J., Shpigelman, A., Kyomugasho, C., Buggenhout, S. V., Ramezani, M., Wang, W. Q., Bao, Y. H., & Chen, Y. (2013). Characteristics and antioxidant activity of
Loey, A. M. V., et al. (2014). The impact of extraction with a chelating agent under water-soluble Maillard reaction products from interactions in a whey protein isolate
acidic conditions on the cell wall polymers of mango peel. Food Chemistry, 161, and sugars system. Food Chemistry, 139(1–4), 355–361.
199–207. Wang, X., Chen, Q., & Lü, X. (2014). Pectin extracted from apple pomace and citrus peel
Kermani, Z. J., Shpigelman, A., Pham, H. T. T., Loey, A. M. V., & Hendrickx, M. E. (2015). by subcritical water. Food Hydrocolloids, 38, 129–137.
Functional properties of citric acid extracted mango peel pectin as related to its Wang, W., Ma, X., Jiang, P., Hu, L., Chen, Z. Z. J., Ding, T., et al. (2016). Characterization
chemical structure. Food Hydrocolloids, 44, 424–434. of pectin from grapefruit peel: A comparison of ultrasound-assisted and conventional
Khatkar, B. S., Barak, S., & Mudgil, D. (2013). Effects of gliadin addition on the heating extractions. Food Hydrocolloids, 61, 730–739.
rheological, microscopic and thermal characteristics of wheat gluten. International Wang, W., Wu, X., Chantapakul, T., Wang, D., Zhang, S., Ma, X., et al. (2017). Acoustic
Journal of Biological Macromolecules, 53, 38–41. cavitation assisted extraction of pectin from waste grapefruit peels: A green two-
Kim, Y., Teng, Q., & Wicker, L. (2005). Action pattern of Valencia orange PME de- stage approach and its general mechanism. Food Research International, 102,
esterification of high methoxyl pectin and characterization of modified pectins. 101–110.
Carbohydrate Research, 340, 2620–2629. Wu, Y., Eskin, N. A. M., Cui, W., & Pokharel, B. (2015). Emulsifying properties of water
Kravtchenko, T. P., Voragen, A. G. J., & Pilnik, W. (1992). Analytical comparison of three soluble yellow mustard mucilage: A comparative study with gum Arabic and citrus
industrial pectin preparations. Carbohydrate Polymers, 18, 17–25. pectin. Food Hydrocolloids, 47, 191–196.
Liu, L., Cao, J., Huang, J., Cai, Y., & Yao, J. (2010). Extraction of pectins with different Yang, J. S., Mu, T. H., & Ma, M. M. (2018a). Extraction, structure, and emulsifying
degrees of esterification from mulberry branch bar. Bioresource Technology, 101, properties of pectin from potato pulp. Food Chemistry, 244, 197–205.
3268–3273. Yang, X., Nisar, T., Hou, Y., Gou, X., Sun, L., & Guo, Y. (2018b). Pomegranate peel pectin
Lutz, R., Aserin, A., Wicker, L., & Garti, N. (2009). Structure and physical properties of can be used as an effective emulsifier. Food Hydrocolloids, 85, 30–38.
pectins with block-wise distribution of carboxylic acid groups. Food Hydrocolloids, Yapo, B. M., & Koffi, K. L. (2006). Yellow passion fruit RindA potential source of low-
23, 786–794. methoxyl pectin. Journal of Agricultural and Food Chemistry, 54, 2738–2744.
Madhav, A., & Pushpalatha, P. B. (2006). Characterization of pectin extracted from Yapo, B. M., Wathelet, B., & Paquot, M. (2007). Comparison of alcohol precipitation and
different fruit wastes. Journal of Tropical Agriculture, 40, 53–55. membrane filtration effects on sugar beet pulp pectin chemical features and surface
Minjares-Fuentes, R., Femenia, A., Garau, M. C., Meza-Vel� azquez, J. A., Simal, S., & properties. Food Hydrocolloids, 21, 245–255.
Rossell�o, S. (2014). Ultrasound-assisted extraction of pectins from grape pomace Zhang, L., Ye, X., Ding, T., Sun, X., Xu, Y., & Liu, D. (2013). Ultrasound effects on the
using citric acid: A response surface methodology approach. Carbohydrate Polymers, degradation kinetics, structure and rheological properties of apple pectin. Ultrasonics
106, 179–189. Sonochemistry, 20(1), 222–231.
Min, B., Lim, J., Ko, S., Lee, G. K., Lee, S. H., & Lee, S. (2011). Environmentally friendly
preparation of pectins from agricultural byproducts and their structural/rheological
characterization. Bioresource Technology, 102, 3855–3860.

10

You might also like