You are on page 1of 12

Food Hydrocolloids 121 (2021) 106845

Contents lists available at ScienceDirect

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

Extraction, characterization and gelling ability of pectins from Araçá


(Psidium cattleianum Sabine) fruits
Sarah da Costa Amaral a, b, Denis Roux b, François Caton b, Marguerite Rinaudo c, *, Shayla
Fernanda Barbieri a, Joana Léa Meira Silveira a
a
Postgraduate Program in Biochemistry Sciences, Sector of Biological Sciences, Federal University of Paraná, Curitiba, PR, 81531-980, Brazil
b
University Grenoble Alpes, CNRS, Grenoble INP, LRP, F-38000, Grenoble, France
c
Biomaterials Applications, 6 Rue Lesdiguières, 38000, Grenoble, France

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

Keywords: Pectins from Araçá (Psidium cattleianum Sabine) fruit pulp were extracted, purified and isolated, obtaining the
Araçá Araçá Purified Pectin (APP) under its sodium form. Yield was found 3.5% w/w of the initial dried material. From
Pectin titration, monosaccharide analysis after total hydrolysis and 1H and HSQC NMR analyses, it was established that
Molecular structure
the pectins are formed of long galacturonic acid blocks (HG) partially methylated (degree of methyl-esterification
De-esterification
– DM = 55.9 mol%) and acetylated (2 mol%) and RG-I blocks with side chains containing mainly arabinose and
Gelation
Rheology galactose. No gelation was observed in presence of calcium for the unmodified APP. Then, pectins were treated in
alkaline conditions to de-esterify progressively the initial sample down to DM = 0 after 60 min of treatment. Each
sample obtained was characterized by SEC and NMR demonstrating the stability of the main chains, where Rha,
GalA and Mw remained unchanged independently of the length of alkaline treatment. One of this sample (APP-
15) was selected to test the gelling ability and rheological behavior. The gel was formed by dialysis against 1 mol
L− 1 CaCl2 and its rheological study allows to conclude that a homogeneous physical gel is formed stabilized by
cooperative GalA− -Ca+2 junctions.

1. Introduction economic use of this natural resource of the Atlantic Forest if the actual
properties mentioned on laboratory scale are considered as mentioned
The fruit by-products have been demonstrating potential for use in in the following.
the development of functional ingredients. In contrast, many fruits from Studies report that Araçá fruit extracts have several pharmacological
native biodiversity remain an under-explored source of bioactive com­ properties, such as antimicrobial (Lima et al., 2020), antioxidant
pounds, unknown to the agri-food industry and to consumers (Biazotto (Denardin et al., 2015) and antihyperglycemic (Vinholes, Lemos, Bar­
et al., 2019; Picot-Allain, Ramasawmy, & Emmambux, 2020). bieri, Franzon, & Vizzotto, 2017) effects. These bioactivities are mainly
The Psidium cattleianum Sabine (Myrtaceae family), known as Araçá attributed to the high content of phenolic compounds, vitamin C, car­
in Brazil and as strawberry guava or cattley guava in other countries, is a otenes, anthocyanins and flavonoids (Cardoso et al., 2018; Medina et al.,
Brazilian native species appreciated by local populations for the con­ 2011; Pereira et al., 2018) which are well known as secondary metab­
sumption of fresh fruit and the manufacture of artisanal sweets produced olites with high antioxidant capacity (Pereira et al., 2018). These studies
in small family-based units (Patel, 2012; Pereira et al., 2018). Nowadays reporting the biological activities of Araçá demonstrated its potential in
in Brazil, there are no Araçá plantations destined to large-scale trade or vitro and in vivo. However, there are still no pharmacological products
industrial applications (Bittencourt et al., 2019; Haminiuk, Sierakowski, with Araçá extract available on the market. Commercially, in Brazil,
Vidal, & Masson, 2006). However, Araçá orchard can produce around only the essential oil from the Araçá leaves is sold for aromatherapeutic
10 ton of fruit per ha, considering the production of 2 kg per plant purposes by Grupo Laszlo © (Emporio Laszlo, 2021).
(Franzon, de Campos, Proença, & Sousa-Silva, 2009). Thus, there is In relation to the investigation of polysaccharides present in the
great interest in seeking applications and in enabling the sustainable Araçá fruits, until the moment, only one study was found, where

* Corresponding author.
E-mail address: marguerite.rinaudo38@gmail.com (M. Rinaudo).

https://doi.org/10.1016/j.foodhyd.2021.106845
Received 11 December 2020; Received in revised form 16 April 2021; Accepted 19 April 2021
Available online 10 May 2021
0268-005X/© 2021 Elsevier Ltd. All rights reserved.
S. da Costa Amaral et al. Food Hydrocolloids 121 (2021) 106845

Vriesmann, Petkowicz, Carneiro, Costa, and Beleski-Carneiro (2009) formation in the presence of excess acid or calcium, after the de-
analyzed the pectic polysaccharides present in the Araçá pulp. However, esterification process, is highlighted.
only parameters such as monosaccharide composition and extraction
yield were investigated, while the fine chemical structure of the pectins 2. Materials and methods
of this fruit remains unknown, as well as reports on the purification,
molar mass, and the gelling behavior. 2.1. Materials
Pectins are polysaccharides with a wide range of applications due to
their functional properties, being widely used in the food industry as a Ripe Araçá (Psidium cattleianum Sabine) fruits were collected from a
gelling, thickening, texturizing, emulsifying and stabilizing agent (Cao, conservation area in Irati-Paraná/Brazil, located at geographic co­
Lu, Mata, Nishinari, & Fang, 2020; Ciriminna, Chavarría-Hernández, ordinates of 25◦ 25′ South latitude, 50◦ 36′ West longitude, and 25◦ 17′
Hernández, & Pagliaro, 2015; Naqash, Masoodi, Rather, Wani, & Gani, South latitude, 50◦ 30′ West longitude. The fruits were selected, washed,
2017). In addition, pectins are also low toxic, biocompatible and and the seeds were removed manually. Then, the seedless fruits were
biodegradable, and can be applied in different areas beyond the food homogenized using a mixer (Black & Decker - SB60-BR, 300 W), giving
appliance, such as the pharmaceutical industry and biomedical appli­ rise to Araçá pulp (AP).
cations (Khotimchenko, 2020; Munarin, Tanzi, & Petrini, 2012; Noreen Methanol (≥99.8%), H2SO4 (95.0–98.0%), 3,5-dinitrosalicilic acid
et al., 2017). (DNS), and sodium tetraborate decahydrate (≥99.5%) were obtained
Structurally, pectins form a group of complex polysaccharides and, from VETEC (Duque de Caxias, Brazil). NaN3 (>99%), absolute grade
despite the wide variation in structure depending on the type of source, ethanol (P.A., ≥99.9%), and HCl 1 mol L− 1(1 N) Titripur® were pur­
the location on the plant and the extraction method, they can be clas­ chased from Merck (Darmstadt, Germany). Inositol (myo-Inositol,
sified into three main types of blocks, according to common character­ ≥99%), m-hydroxydiphenyl (phenyl phenol, ≥90%), NaBH4 (≥98.0%),
istics: homogalacturonans (HG), rhamnogalacturonans type I (RG-I) and NaCl (≥99%), NaOAc (≥99%), NaOH (≥97%), NaNO2 (≥99%), pyri­
rhamnogalacturonans type II (RG-II) (Dranca & Oroian, 2018; Gaw­ dine (≥99%), %), the standards glucose (Glc - purity ≥99%), galactur­
kowska, Cybulska, & Zdunek, 2018; Kravtchenko, Voragen, & Pilnik, onic acid (GalA - purity ≥97%), glucuronic acid (GlcA - purity ≥99%),
1992; Naqash et al., 2017; Voragen, Coenen, Verhoef, & Schols, 2009). arabinose (Ara - purity ≥99%), rhamnose (Rha - purity ≥99%), xylose
Most of the structure of pectins (~65%) usually consists of HG blocks (Xyl - purity ≥99%), mannose (Man - purity ≥99%), and trifluoroacetic
which are homopolymers of (1 → 4)-linked α-D-galacturonic acid units acid (TFA, ≥99.0%), were purchased from Sigma-Aldrich (St. Louis,
that are partially methyl-esterified at C-6. The degree of methyl- USA).
esterification (DM) is commonly used to classify pectins into: high Cellulose MWCO 3500 D cut off membrane from acquired Spectrum
methoxyl pectin (HM), when DM is larger than 50% and low methoxyl Labs™ and ultrapure water (18 MΩ, pH 6.9) was obtained through the
pectin (LM) when DM is less than 50% (Chan, Choo, Young, & Loh, Master system-MS2000/Gehaka.
2017; Picot-Allain et al., 2020; Voragen et al., 2009). The HG blocks are
interrupted by RG-I, which usually represents about 20–35% of the 2.2. Extraction, fractionation, and purification of polysaccharides
structure, formed by the repeating disaccharide [→2)-α-L-Rhap-(1 →
4)-α-D-GalAp-(1→]n ≥ 1, with neutral oligosaccharide side chains The Araçá pulp (AP, 200 g) was submitted to an enzymatic inacti­
mainly composed of arabinans, galactans, and arabinogalactans (AG) vation and depigmentation treatment with 99.5% ethanol in a boiling
(Khotimchenko, 2020; Voragen et al., 2009; Yapo, 2011). water bath under reflux for 30 min, giving the alcohol insoluble residues
The chemical structure of pectins and the degree of methyl- (Residue I, in Fig. 1). The aqueous extract from the Residue I in mild
esterification (DM) affects the gelling properties of these poly­ conditions was obtained by extraction with ultrapure water in a water
saccharides. While the physical-chemical conditions for the formation of bath at 98 ◦ C, under reflux for 2 h. Then, the pH of the solution was
gels with HM pectins usually are the presence of co-solutes (such as adjusted to 7 with 1 mol L− 1 NaOH solution. The volume was reduced to
sucrose) in high concentrations (60–65%) and in acidic pH (pH < 3.5), a quarter of the initial volume under vacuum, followed by precipitation
LM pectins gels are formed in the presence of Ca2+ ions through asso­ of the water soluble components with ethanol/water (up to around 1:1,
ciations between sequences of charged carboxylic groups of two v/v), washing the precipitate with increasing concentrations of ethanol/
different chains (Rinaudo, 1996; Stephen, Phillips, & Williams, 2006, p. water (70%, 80% v/v) and at end with ethanol P.A. and then dried at
702). atmospheric conditions (~1 atm, 25 ◦ C). The isolated polymers are
Native pectins are commonly HM type. Thus, studies have been identified as Araçá Crude Polysaccharide (ACP) (Fig. 1).
dedicated to producing LM pectins and to investigate the calcium The ACP fraction was further purified according to Rinaudo (2005)
binding properties with different levels and patterns of controlled de- with some adjustments: 1 g of the obtained crude fraction ACP was
esterification (Chan et al., 2017; Ngouémazong et al., 2012; Ralet, solubilized in 100 mL of deionized water. This solution was left under
Dronnet, Buchholt, & Thibault, 2001; Ventura, Jammal, & Bianco-Peled, magnetic stirring overnight (16 h). Then, the solution was centrifuged
2013), given that Ca2+- dependent gelation is one of the important (4600×g, 1 h, at 20 ◦ C) to remove water-insoluble impurities and get
functional properties of pectins to enable widespread use in the industry ACP-S. After centrifugation, the precipitate was discarded and 0.45 g of
(Cao et al., 2020). Indeed, anionic hydrogels with hydrophilic properties NaCl was added to the supernatant (0.08 mol L− 1), under stirring and
and lower sensitivity towards pH variation are widely used in various left to stir for 30 min to get the sodium form of charged polymers. The
fields of biomedicine including bondages, highly effective absorbents, pH was adjusted to 7 with 0.1 mol L− 1 NaOH solution. Then, ethanol
and targeted drug delivery systems (Khotimchenko, 2020). In addition, precipitation was carried out with increasing concentrations of ethanol.
LM pectins are important ingredients in the food industry to form low The polysaccharide precipitation was observed for a final ethanol/water
sugar jams, since it does not require the addition of sucrose for gelation mixture of approximately 60% (v/v). After, the precipitate was washed
(Noreen et al., 2017). with increasing concentrations of ethanol/water (70%, 80% v/v) and at
Currently, pectins are being sought in sources other than those ob­ end with ethanol P.A. The sample was dried at atmospheric conditions
tained from the citrus peel. Thus, polysaccharides with new structure (~1 atm, 25 ◦ C) until constant weight, obtaining the Araçá Purified
and new physical-chemical characteristics should allow a larger range of Pectin (APP) (Fig. 1).
industrial applications. In view of this, the present work aims to eluci­
date the chemical and molecular structure of the pectic polysaccharides 2.3. Protein content
extracted from the fruits of Araçá using an ecologically sustainable
method. In addition, the potential for application of these pectins for gel Protein content of the Araçá pectin samples was determined using

2
S. da Costa Amaral et al. Food Hydrocolloids 121 (2021) 106845

Fig. 1. Scheme of extraction and purification of pectins from the pulp of Araçá (Psidium cattleianum Sabine) fruits.

the Bradford method (Bradford, 1976). A calibration curve of bovine analyzed by the colorimetric m-hydroxybiphenyl method (Blu­
serum albumin (BSA) was built as a standard. Absorbance was measured menkrantz & Asboe-Hansen, 1973). Measurements were done in tripli­
by spectrophotometry (Epoch, BioTek, USA) at 595 nm, and results were cate and the values represent the averages ± SD.
expressed in g protein/100 g of sample. The measurements were done in The identity of the uronic acid was determined by anion exchange
triplicate and the values represent the averages ± SD (standard chromatography with pulse amperometric detection (HPAEC-PAD).
deviations). Samples were prepared as described by Amaral et al. (2019) and injected
into a Thermo Scientific Dionex ICS-5000 chromatograph (Thermo
Fisher Scientific, USA) with CarboPac PA20 column (3 × 150 mm), with
2.4. Uronic acid content and degree of methyl-esterification (DM) a gradient of 0.5 mol L− 1 NaOH and 1 mol L− 1 NaOAc as eluent (Nagel,
Sirisakulwat, Carle, & Neidhart, 2014) in an N2 atmosphere in a flow of
The uronic acid content and the degree of methyl-esterification (DM) 0.2 mL min− 1 at 24 ◦ C. The data was collected and analyzed using the
were quantified using a conductometric titration technique described Chromeleon TM 7.2 Chromatography Data System software. The gal­
previously (Cárdenas, Goycoolea, & Rinaudo, 2008; Thibault & acturonic acid was used as the standard and analysis was carried out in
Rinaudo, 1985). 20 mg of APP sample was dissolved in 50 mL of triplicate.
deionized water and solubilized overnight. Then, to determine the
uronic acid content in free carboxylic acid, titration was performed with
2.5. APP de-esterification
0.02 mol L− 1 HCl solution followed by titration with 0.02 mol L− 1
NaOH, using a conductometer (Thermo Scientific, Orion Star A215). The
To prepare pectins with different degrees of methyl-esterification
same titration was also used after total de-esterification (as described in
(DM), the de-esterification reaction was carried out with 0.02 mol L− 1
the next section 2-5). The difference between the uronic acid content
NaOH for 15 min (APP-15), 30 min (APP-30), 60 min (APP-60), and 105
after total de-esterification (total GalA) and the initial carboxylic groups
min (APP-105). Briefly, 150 mg of sample (APP) was solubilized in 24
content (GalA− Naþ) allowed to get the value of methyl-esterified uronic
mL of deionized water under stirring for 1 h and kept refrigerated
acid yield (GalA-Met). The degree of methyl-esterification (DM) was
overnight (~16 h). After, 6 mL of 0.1 mol L− 1 NaOH was added under
calculated through the ratio between GalA-Met and total GalA expressed
stirring. After 15, 30, 60 or 105 min of reaction in alkaline conditions, 1
in molarity x 100 as usually adopted.
mol L− 1 HCl was added gradually until reaching pH ~ 7–7.5. Then,
Additionally, from those determinations, the fraction of neutral
ethanol precipitation and drying of APP de-esterified samples were
carbohydrates (corresponding to the neutral side chains of the pectins in
carried out as described in section 2-2.
the sample) can be deduced by the difference between the total mass of
sample and total GalA content. And finally, to convert the percentages
obtained from g/g to mol/g, the values were normalized considering the 2.6. Neutral monosaccharide composition
molar mass of 198 g mol− 1, 190 g mol− 1, and 162 g mol− 1 for Gal­
A− Naþ, GalA-Met and neutral monosaccharides, respectively. Neutral monosaccharides involved in the polysaccharide structure
The content of total uronic acids of the Araçá pectin samples was also may be determined after total hydrolysis of the initial sample and

3
S. da Costa Amaral et al. Food Hydrocolloids 121 (2021) 106845

fractionation. However, the kinetic and conditions of hydrolysis are by 1H NMR spectroscopy integrating the hydrogen areas corresponding
important (Nagel et al., 2014). Then, in order to determine the best to H-1, H-5 of unesterified α-D-GalAp units and H-5 of esterified α-D-
hydrolysis condition to obtain the neutral monosaccharide rate, a hy­ GalAp units as discussed by Grasdalen, Bakøy, and Larsen (1988). The
drolysis kinetics was performed with the APP sample. hydrogen from –CH3 around 3.76 ppm gives DM using H-1 integral as
Briefly, 10 mg of APP were solubilized with 3.5 mL of 2 mol L− 1 TFA reference (Patova et al., 2019).
and this solution was divided into 7 tubes for hydrolysis at 100 ◦ C at
different times: 0.5, 1, 2, 3, 4, 6, and 8 h. After hydrolysis, the acid was 2.9. Gelation ability and characterization
removed by evaporation at 25 ◦ C under air current and washed with
methanol (1 mL x 3). Then, 1 mL of ultrapure water was added to each Depending on the degree of methyl-esterification (DM), pectins are
tube. able to form gel in acidic conditions (excess of HCl) and/or in presence
At each time (0.5, 1, 2, 3, 4, 6, and 8 h) of hydrolysis kinetics, the of calcium. Different amounts of the each samples were solubilized in 1
total carbohydrates (phenol-sulfuric acid method, Dubois, Gilles, Ham­ mL of deionized water. Then, 100 μL of 1 mol L− 1 HCl or 100 μL of 1 mol
ilton, Rebers, & Smith, 1956) and reducing carbohydrates (dini­ L− 1 CaCl2 was added to determine the critical concentration necessary
trosalicylic acid - DNS method, Miller, 1959) were measured. Both for the formation of homogeneous gel. Then, the critical concentration
determinations were calibrated using a standard solution containing for gelation was determined at different DM.
GalA and Ara standards (70:30, w/w) in triplicate and the values To prepare the pectin gels, a 10 g L− 1 solution previously dissolved in
represent the averages ± SD. deionized water were placed into a dialysis cellulose bag (Spectrum™
The neutral monosaccharides composition was determined using Labs Spectra/Por™, MWCO 3500 D) and placed to dialysis against 1
Gas–Liquid Chromatography (GLC - Thermo Scientific Trace GC Ultra mol L− 1 CaCl2 solution for 24 h at 4 ◦ C (Rahelivao, Andriamanantoa­
gas chromatograph) after acid hydrolysis. For that purpose, 10 mg of nina, Heyraud, & Rinaudo, 2013). The resulting formed gels, stabilized
samples, added of 50 μL of an inositol solution at 10 mg mL− 1 used as at 25 ◦ C, were recovered by cutting off the dialysis membrane with a
internal standard, were hydrolysed with 2 mol L− 1 TFA for 1 h or 8 h at scalpel and punched into 2 mm thickness disks for the rheological
100 ◦ C, followed by conversion to alditol acetates by NaBH4 reduction measurements.
(100 ◦ C for 10 min) (Wolfrom & Thompson, 1963a) and acetylation with
acetic anhydride (Ac2O)-pyridine (1:1, v/v, 1 mL) at 100 ◦ C for 30 min 2.10. Rheological characterization
(Wolfrom & Thompson, 1963b). The resulting alditol acetates were
extracted with CHCl3, and identified by comparing their profiles and Rheological gel characterization was performed on an ARES-G2
retention times with standards after GLC. A mixture of He and N2 with rheometer from TA Instruments Company. The APP-15 gel at 10 g L− 1
compressed air was applied as a carrier gas at 1 mL min− 1, and a formed in 1 mol L− 1 CaCl2 was loaded between the two rough plates
DB-225-MS column programmed from 100 ◦ C to 230 ◦ C at a heating rate separated by a gap of 2 mm and of 10 mm diameter. All experiments
of 60 ◦ C min− 1 was utilized. To determine the total neutral mono­ were performed at 28 ◦ C, thanks to the forced convection oven. Three
saccharide (TNM) content, the peak areas of all the detected mono­ different types of experiments were conducted: (1) a classical oscillatory
saccharides were summed, quantified using inositol as internal standard, shear frequency sweep, (2) an orthogonal oscillatory frequency sweep
and the values represent the averages ± SD of a duplicate experiment. and (3) one compression test from 2 to 1.5 mm gap at constant velocity
0.05 mm s− 1. Experiments (1) and (2) were performed under respec­
2.7. Size exclusion chromatography tively 0.1% and 0.05% strain well under the upper limit of the linear
regime determined at 0.5% strain.
The weight-average molecular weights (Mw) of each sample were
determined through Size Exclusion Chromatography (SEC) measure­ 3. Results and discussion
ments using two columns in series from Shodex SB 806 M HQ associated
with a pre-column SB-G 6B. The solvent used was 0.1 mol L− 1 NaNO3 3.1. Extraction, purification and structural characterization of the
including 0.3 g L− 1 NaN3. A laser light scattering Mini Dawn Treos de­ polysaccharide extracted from Araçá pulp
tector from Wyatt was associated with a Wyatt differential refractom­
eter. The selected flow rate was 0.5 mL min− 1 at 30 ◦ C, the dn/dc was The Araçá Crude Polysaccharide (ACP) was obtained after hot water
equal to 0.155 and sample concentration injected was 4 mg mL− 1. The extraction and precipitation with ethanol (Fig. 1). A relatively low
samples were dissolved in water under stirring during 24 h at 25 ◦ C and ethanol concentration (60%) was used to avoid the precipitation of
filtrated on a 0.2 μm porous membrane before injection. neutral polysaccharides and eventually oligosaccharides and proteins
able to exist in aqueous extract of pulp or peel (Berriaud, Milas, &
2.8. Nuclear magnetic resonance (NMR) spectroscopy Rinaudo, 1998; Rinaudo, 2005). Then, the precipitate may be assumed
as a charged polymer, mainly pectic materials. The extraction yield of
Mono-dimensional (1H-) and bi-dimensional (HSQC) NMR spectra ACP reached 4.7% (w/w) of the dry fruit (AP). The ACP fraction was
were acquired at 70 ◦ C with a Bruker Avance III 400 MHz or Bruker then purified by centrifugation in order to remove insoluble particles,
Avance III HD 600 MHz spectrometers, equipped with a BBI 5 mm probe followed by addition of monovalent salt (Fig. 1). This step allows to
(400 MHz) or with a 5 mm CPP-TCI probe (600 MHz) (Bruker, Billerica, exchange ions, mainly calcium counterions, existing often in pectin
Massachusetts, USA). extracts for sodium and to screen the long-range electrostatic repulsions
The samples were solubilized in D2O and the chemical shifts of the (Rinaudo, 2005). Then, after ethanol precipitation, the Araçá Purified
polysaccharide were expressed as δ (ppm), using the resonances of –CH3 Pectin (APP) fraction was obtained, with the yield of 75% (w/w) in
groups of acetone (1H at δ 2.22; 13C at δ 30.20) as internal references. relation to the starting ACP fraction (i.e. 3.5% w/w of the initial dried
The data were collected and processed using the Software TOPSPIN, material).
version 3.1 (Bruker Biospin, Rheinstetten, Germany). Concerning the composition of the sample, the content of residual
Complementary 1H NMR measurements were performed at 353 K on proteins was tested but the APP fraction was free of proteins. According
a Bruker AVANCE I instrument equipped with a 5 mm probe operating at to the literature, the Araçá fruit has a low protein content (0.75–1.03 g/
400.2 MHz. An amount of 3 mg of sample was dissolved in 0.5 mL of 100 g) (Pereira et al., 2018). In addition, considering the conditions used
D2O, and 32 scans were recorded. Chemical shifts are expressed using for isolation of APP, absence of proteins is usually demonstrated in
H2O at 4.25 ppm at 80 ◦ C as reference. different ionic polysaccharides (Berriaud et al., 1998; Rinaudo, 2005).
The values of degree of methyl-esterification (DM) were determined Considering that the ability of pectins to form hydrogels can be

4
S. da Costa Amaral et al. Food Hydrocolloids 121 (2021) 106845

affected by intrinsic parameters, such as the degree of methyl- was obtained by dosage of the total carbohydrates (Dubois et al., 1956)
esterification (DM), the APP fraction was submitted to chemical modi­ and the reducing carbohydrates (Miller, 1959). Through Fig. 2 it is
fication to obtain four pectins with different DM. After 15, 30, 60, and possible to observe that 1 h of acid hydrolysis gives the maximum value
105 min of de-esterification reaction in alkaline conditions, it was ob­ of carbohydrates, which we adopt as the best condition for the deter­
tained samples named APP-15, APP-30, APP-60, and APP-105, mination of total neutral monosaccharides for the APP sample.
respectively. From the hydrolysis kinetics, it was observed that the APP sample
By conductimetry, it was possible to determine the content of free showed a total reducing carbohydrates of 0.64 (±0.01) mg mL− 1 after 1
uronic acid (on –COO-Na+ form) on each sample. Firstly, the titration h of hydrolysis corresponding to 50.4% as the maximum yield hydrolysis
was performed on APP sample. It is found 162.4 × 10− 5 -COO- per g, of the initial dried sample. After 8 h of hydrolysis, 0.22 (±0.05) mg mL− 1
which means 0.162 mol of GalA in 100 g of dried sample (Table 1). Each of monosaccharide are produced i.e. the fraction of sample equals
sample was titrated to get the number of ionic units in the dried weight, 17.3%. The decrease of carbohydrate content to 0.22 (±0.05) mg mL− 1
which was expressed in g/100 g sample (Table 1). Taking into account at 8 h of hydrolysis confirmed the occurrence of neutral monosaccharide
the total number of GalA units obtained on fully de-esterified sample, degradation in longer hydrolysis times with 2 mol L− 1 TFA at 100 ◦ C.
the number of esterified GalA is deduced and following the degree of The difference between the degrees of acid hydrolysis with time may be
methyl-esterification (DM). The nearly constant value of total GalA− Na+ due to the difficulty of finding a compromise between destruction of the
after 60 min and 105 min confirms that total de-esterification was less stable monosaccharide (Ara) and absence of hydrolysis of the more
achieved after 60 min of reaction. resistant glycosidic bonds in polyuronates, such as pectins (Mankarios
For APP-60 it was found 363.1 × 10− 5 -COO- per g, which means et al., 1979).
0.363 mol GalA in 100 g of dried sample (Table 1). Taking into account The neutral carbohydrates analysis on these hydrolyses allows to
the molar mass of the repeat unit under sodium form = 198 g mol− 1, the determine Rha, Ara and Gal contents with inositol as reference giving a
fraction of GalA in the polysaccharide equals 71.9 g/100 g sample. Then, total neutral monosaccharide (TNM) content of 14.3 (±0.6) g/100 g and
28.1 g/100 g should be represented by neutral saccharides. 11.2 (±0.8) g/100 g of dried APP sample for 1 and 8 h of hydrolysis,
The content of uronic acid was also determined by Blumenkrantz and respectively (Table 2). Considering the maximum degree of hydrolysis,
Asboe-Hansen (1973) method and the value obtained (76.9 ± 2.4 g/100 it is concluded that the content in neutral carbohydrates may correspond
g) is in relatively good agreement with the value of 71.9 g/100 g ob­ to 28.4 g/100 g dried sample (14.3/50.4 × 100) completing the total
tained by titration (Table 1). Furthermore, the identification of the galacturonic acid determined as 71.9 g/100 g given by titration. In
uronic acid type of the APP sample was determined as GalA by addition, from Table 2, it may be concluded that rhamnose involved in
HPAEC-PAD. RG-I zone is progressively released due to the neighbor galacturonic acid
The degree of methyl-esterification (DM) of the APP sample obtained preventing hydrolysis when arabinose is labile.
from conductometric titration equals DM = 55.9% characterizing this According to the literature, similar monosaccharide composition was
polymer as a high methoxyl (HM) type (Munarin et al., 2012). A similar observed in water extracted pectins from other fruits belonging to the
high DM (60%) was observed for the GW pectin extracted from the Myrtaceae family such as gabiroba (Campomanesia xanthocarpa) (Bar­
Gabiroba (Campomanesia xanthocarpa Berg) (Barbieri et al., 2019) and bieri et al., 2019) and in a mixture of two species of guavira.
for the Murta (Ugni molinae Turcz) pectin (57%) (Taboada et al., 2010) (Campomanesia pubescens and C. adamantium) (Schneider, Iacomini,
both species belonging to the Myrtaceae family. & Cordeiro, 2019), which also presented Ara as the major neutral
In order to determine the value of total neutral monosaccharides monosaccharide. However, Araçá pectin showed a higher amount of
(TNM) of the APP sample it was first selected the best hydrolysis con­ galacturonic acid (76.9% by Blumenkrantz and Asboe-Hansen (1973)
ditions, taking into account the wide variation on the stability of method – Table 1) compared to those pectins containing 33.5% and
glycosidic bonds to acid hydrolysis and that different monosaccharides 44.6% of GalA, respectively.
are destroyed by acid at different rates (Mankarios, Jones, Jarvis, This high content of GalA (76.9%) presented by Araçá pectins is
Threfall, & Friend, 1979). interesting considering that this parameter is desired for commercial
Therefore, considering the balance between the total hydrolysis time purposes, since this value is closer to that observed for commercial citrus
and the degradation of free monosaccharides, a kinetics of hydrolysis pectins reported by Kravtchenko et al. (1992) and by Ma et al. (2020),
which presented 76.4–77.1% and 68.2%, respectively. Furthermore, we
highlight that Araçá pectin was obtained by a less drastic extraction
Table 1
Uronic acid and DM of pectins obtained from the Araçá pulp.
method with using the boiling water extraction, which is a cheaper and

GalA− Na+ Uronic acid DM


(x10− 3)a (%)a
Sample GalA− Na GalA- Total GalAc
+ b
Metb GalAb

APP 162.4 32.1 39.8 71.9 76.9 ± 55.9


2.4
APP-15 208.0 41.2 30.7 71.9 70.2 ± 43.5
1.3
APP-30 324.1 64.2 7.7 71.9 66.3 ± 11.9
5.1
APP-60 363.1 71.9 0 71.9 66.9 ± 0
1.3
APP- 345.8 68.4 0 68.4 Nd 0
105

Nd = not determined.
a
Quantified by conductometric titration. GalA− Na+ expressed in -COO- per
100 g. DM: degree of methyl-esterification expressed in molar ratio x100.
b
GalA-Met, free Gal-A and total GalA expressed in g/100 g of initial dried Fig. 2. Hydrolysis kinetics of the APP sample. Concentration of total carbo­
sample. hydrates (phenol-sulfuric) and reducing carbohydrates (DNS) at different hy­
c
(g/100 g) Determined by Blumenkrantz and Asboe-Hansen (1973) method. drolysis times (from 0.5 h – up to 8 h) on 1.27 mg (dried weight) in 1 mL.
Measurements were done in triplicate and the values represent the averages ± Measurements were done in triplicate and the values represent the averages
SD (standard deviations). ± SD.

5
S. da Costa Amaral et al. Food Hydrocolloids 121 (2021) 106845

Table 2 of the monosaccharide composition obtained for APP equals 27.6


Monosaccharide composition of pectins obtained from the Araçá fruit. considering the rhamnose content obtained by NMR. This value was
Monosaccharide composition (g/100 g) similar to the ratio of the S3–N commercial citrus pectin (22.3) (Fra­
casso, Perussello, Carpine, Petkowicz, & Haminiuk, 2018) and indicates
Sample GalAe Rha Ara Gal TNMf
predominance of long HG regions (Zhang et al., 2018). The large dif­
a,b
APP 71.9 0.3 ± 0.1 13.6 ± 0.6 0.4 ± 0.1 14.3 ± 0.6 ference between rhamnose obtained by acid hydrolysis (0.3–0.6 g/100
APP a,c 71.9 0.6 ± 0.1 9.6 ± 0.6 1.0 ± 0.2 11.2 ± 0.8
APP d – 2.6 29 Nd Nd
g) and by NMR (2.6 g/100 g) may be attributed mainly to rhamnose
involved in the RG-I blocks resistant to hydrolysis. A significant decrease
Nd = Not determined. (p < 0.001) in Ara from 13.6 after 1 h of hydrolysis to 9.6 after 8 h, can
a
Monosaccharide composition, determined by GLC using inositol as standard.
be observed. In fact, arabinose in the pectin side chains is released more
The values represent the average ± SD of duplicate data.
b rapidly and is more instable after their release during acid hydrolysis
Monosaccharide composition and TNM after 1 h of hydrolisys.
c
Monosaccharide composition and TNM after 8 h of hydrolisys.
(Mankarios et al., 1979).
d
Monosaccharide composition by 1H NMR. Therefore, it is demonstrated that this hydrolysis is never complete
e
g/100 g of total uronic acid determined by titration. GalA was identified by due to resistant GalA blocks and that, in addition, the monosaccharide
HPAEC-PAD. content depends on time of hydrolysis (including instability of arabi­
f
TNM: Total neutral monosaccharides after hydrolysis, determined by GLC nose). In the initial sample, neutral saccharidic chains is found around
using inositol as standard. The values represent the average ± SD of duplicate 30 wt % as obtained from 1H NMR in which the side chains mainly
data. consist of Ara and, in smaller quantities, Rha and Gal.

environmentally friendly method compared to the acidic extraction used 3.2. NMR structure characterization
to obtain commercial citrus pectins. In addition, the purification step by
1
precipitation with 60% ethanol in presence of NaCl gives pure pectin H/13C heterocorrelated HSQC NMR spectrum was employed to
under sodium form. provide detailed structural information about the APP sample isolated
However, when comparing the values of neutral monosaccharides, from Araçá pulp. In Fig. 3 the HSQC NMR spectrum revealed 1H/13C
the neutral side chains of APP pectin is mainly composed of Ara (13.6% correlations from HG assigned to the anomeric carbon of →4)-α-D-
± 0.6 after 1 h hydrolysis), unlike commercial citrus pectins, which 6MeGalAp-(1→ at δ 4.96/99.6 (unit A1) and of →4)-α-D-GalAp-(1→ at δ
generally have Gal (6.8–12.4%) as the main neutral monosaccharide 5.13/99.0 (unit B1) (Dias, Barbieri, Fetzer, Corazza, & Silveira, 2020).
and a lower content of Ara (3.1–3.3%) (Kravtchenko et al., 1992; Ma The unesterified and methyl-esterified units of α-D-GalAp ring correla­
et al., 2020). This difference may play a role on solubility of Araçá tions were showed in Table 3. The signal at δ 3.81/52.8 corresponding to
pectins, considering that arabinose side-chains have strong methyl group (O–CH3) of α-D-6Me-GalAp (1 → 4) units, whereas the
water-binding capacity (Zeng et al., 2020). acetyl group (O–Ac) was observed at δ 2.09/20.09 ppm (Colodel,
An important intrinsic parameter for the gelling ability is the high Vriesmann, Teófilo, & Petkowicz, 2018; Patova et al., 2019).
proportion of HG blocks (Chan et al., 2017). The GalA/Rha molar ratio Characteristic correlations attributed to the presence of RG-I blocks

Fig. 3. 1H/13C HSQC spectrum of the APP sample from the Araçá pulp. The sample was dissolved in D2O and data were collected at a probe temperature of 70 ◦ C and
measured relative to acetone at 1H (δ 2.22) and 13C (δ 30.2). The chemical shifts (δ, ppm) corresponding to each letter are described in Table 3.

6
S. da Costa Amaral et al. Food Hydrocolloids 121 (2021) 106845

Table 3
1
H/13C HSQC chemical shifts (δ, ppm) of the APP sample from the Araçá pulp.
Glycosil units Chemical shifts, δ (ppm)

1 2 3 4 5 6 -OCH3

5a 5b
13
A →4)-α-D-6MeGalAp-(1→ C 99.6 68.1 68.1 78.2 70.4 – nd 52.8
1
H 4.96 3.73 3.98 4.45 5.05 – 3.81
13
B →4)-α-D-GalAp-(1→ C 99.0 68.1 68.1 78.2 71.5 – nd –
1
H 5.13 3.73 3.98 4.40 4.64 – –
13
C →2)-α-L-Rhap-(1→ C nd nd nd nd nd – 16.5 –
1
H – 1.17 –
13
D →2,4)-α-L-Rhap-(1→ C nd nd nd nd nd – 16.5 –
1
H – 1.25 –
13
E t-β-L-Araf-(1→ C 101.6 76.6 74.4 81.9 63.0 –
1
H 5.10 4.13 4.04 3.90 3.80 3.72 –
13
F →3)-α-L-Araf-(1→ C 107.0 79.4 83.6 82.2 61.3 – – –
1
H 5.17 4.36 3.99 4.19 3.81 3.73 –
13
G →5)-α-L-Araf-(1→ C 107.6 81.2 76.6 80.2 66.4 – –
1
H 5.08 4.13 4.04 4.28 3.89 3.81 –
13
H →4)-β-D-Galp-(1→ C 104.3 71.3 72.5 77.6 74.6 – 61.0 –
1
H 4.61 3.54 3.75 4.13 3.70 – 3.81 –

nd: not determined.

was also observed at δ 1.17/16.5 (H-6/C-6) to →2)-α-L-Rhap-(1→ (units and APP-105, respectively. Then, the values obtained by NMR are in
C6) and at δ 1.25/16.5 ppm to substituted →2,4)-α-L-Rhap-(1→ (unit D6) relatively good agreement with the values obtained by titration
(Fig. 3, Table 3), in agreement with the presence of Rha determined by (Table 1), excluding APP-15 (which showed a DM value of 18 by NMR
GLC (Table 2). and a DM value of 43.5 by titration). It is also shown that rhamnose
In addition, the 1H/13C correlations at δ 5.10/101.6 (H-1/C-1), δ content remains constant around 2.6 g/100 g in the samples. Arabinose
4.13/76.6 (H-2/C-2), δ 4.04/74.4 (H-3/C-3), δ 3.90/81.9 (H-4/C-4) decreases slightly from 29 g (APP) to 21 g/100 g (APP-105) in agree­
were assigned to terminal β-L-Araf-(1→ (unit E). Despite few reports of ment with its low stability also observed previously in strong acid hy­
the presence of the t-β-L-Araf-(1→ units, this signal was recently drolysis (Table 2). Clearly, the monosaccharide contents (arabinose,
observed for pectins obtained from Myrtaceae family fruits as Guavira rhamnose) are larger than the values found after acid hydrolysis (1 h or
pomace (Schneider et al., 2019) and Gabiroba pulp (Dias, Barbieri, 8 h) especially for rhamnose involved in the resistant RG-I block. Those
López-Fetzer, Corazza, & Silveira, 2020). yields in neutral monosaccharides obtained by NMR look slightly
The other anomeric signals (H-1/C-1) at δ 5.17/107.0 and δ 5.08/ overestimated to complete the GalA content.
107.6 ppm were attributed to →3)-α-L-Araf-(1→ (unit F) and →5)-α-L- The titration and 1H NMR data allow to conclude, from the molar
Araf-(1→ (unit G), respectively. The correlations corresponding to →4)- ratio GalA/Rha, that Araçá pectin is formed by long GalA blocks (HG
β-D-Galp-(1→ (unit H), were also shown at δ 4.61/104.3, 3.54/71.3, blocks) interrupted by rhamnogalacturonan blocks (RG-I).
3.75/72.5, 4.13/77.6, 3.70/74.6, 3.81/61.0 assigned as H-1/C-1, H-2/
C-2, H-3/C-3, H-4/C-4, H-5/C-5 and H-6/C-6 of →4)-β-D-Galp-(1→ (unit 3.3. Molecular weight characterization
H), respectively (Fig. 3, Table 3), suggesting the presence of arabinans,
galactans or type I arabinogalactans (AG-I), which can be linked as a side SEC results are given in Figs. 5 and 6. A clear modification in the
chain in RG-I (Dias et al., 2020; Schneider et al., 2019). Thus, according molecular weight distribution occurs as soon as the de-esterification
to the monosaccharide composition together with the NMR analyzes it is reaction starts.
possible to suggest that the APP sample has a structure formed mostly by In the initial sample (Fig. 5A), the chromatogram shows that the
HG blocks with a lower proportion of RG-I blocks with a side chain unmodified sample APP demonstrates 3 different series of molecules. A
consisting of arabinans, galactans or AG-I. shoulder in the light scattering signal at lower elution time at ~25 min,
The bidimensional analysis was completed with 1H NMR studied may be attributed to a small amount of aggregates of weight-average
after different times of de-esterification reaction as shown in Fig. 4. A molar mass in the range of 107 (lower than 1 wt %). Then, around 20
large modification occurs as soon as hydrolysis starts allowing to follow wt % of molecules with Mw ~1.5 × 106 g mol− 1 are eluted at 28 min
the demethylation rate. The spectra allows to quantify Rha, acetyl and corresponding to the maximum of the light scattering signal. At ~33
methyl substituents, and Ara relatively easily from the integrals around min, the maximum in the RI detector signal indicates that around 80 wt
1.3 ppm, 2.1 ppm, 3.8 ppm and around 5.11 ppm, respectively. In % of molecules have a Mw ~160,000 g mol− 1. The same type of distri­
addition, around 5–5.1 ppm are H-1 and 4.98 ppm is the H-5 of Me-GalA bution was obtained previously on different commercial pectins
and around 4.67 ppm is the H-5 of free GalA units (as proposed by (Malovikova, Rinaudo, & Milas, 1993; Zhang et al., 2018).
Grasdalen et al., 1988). Interestingly, the ratio between the signal of H-5 As soon as de-esterification reaction starts, the chromatogram is
of free GalA and total of H-5 also approach the DM determination seriously modified with only two fractions as shown in Fig. 5B. Around
avoiding some overlap of the 3.8 ppm signal. 20 wt % of molecules are eluted first with an average Mw ~800,000 g
The evolution of the spectra allows to draw few conclusions con­ mol− 1 followed by 80 wt % with Mw ~60,000 g mol− 1. The molecular
cerning the molecular structure of the different samples as a function of weight distributions of the de-esterified samples (APP-15, APP-30, APP-
the de-esterification reaction time. Firstly, it is shown that the acetyl as 60, and APP-105) are summarized in Fig. 6, where the refractive index
well as the methyl content obtained by 1H NMR decreases progressively detector signal indicates that all the samples have nearly the same mo­
compared to H-1 or H-5 signals. Concerning the degree of acetylation, it lecular weight distribution excluding the initial sample. In Table 4, the
is obtained a ratio Acetyl content/GalA starting as 0.018 (i.e.~2% of weight-average molecular weights (Mw) of the different samples are
GalA units being acetylated) and decreasing to 0.004. The analysis al­ listed showing the molecular weight stability when de-esterification
lows also to determine the DM percentage, giving the values of 58%, reaction time increases.
18%, 10%, 0%, and 0% for the samples APP, APP-15, APP-30, APP-60 These data indicate that the molecular weight of de-esterified pectins

7
S. da Costa Amaral et al. Food Hydrocolloids 121 (2021) 106845

Fig. 4. 1H NMR spectrum of the APP samples isolated


from the Araçá fruit obtained at 80 ◦ C in D2O
(chemical shifts are expressed in δ, ppm). (A) Initial
sample APP, (B) sample obtained after 15 min (APP-
15) of de-esterification reaction; (C) sample obtained
after 30 min (APP-30) of de-esterification reaction;
(D) sample obtained after 60 min (APP-60) of de-
esterification reaction. In the spectrum, region A:
integration of peaks from H-1 of →4)-α-D-GalAp-(1→
units, H-1 and H-5 of →4)-α-D-6MeGalAp-(1→ units.
Region B: integration of peak from H-5 of →4)-α-D-
GalAp-(1→ units.

samples (APP-15, APP-30, APP-60, and APP-105) remained unchanged Thereby, during the alkaline hydrolysis going to DM = 0 after 60 min
for the different alkaline treatments performed. The only change of de-esterification reaction, the molecular weight distributions remain
compared with APP is related to the dissolution of pectins in alkaline unchanged as soon as pectins are in contact with alkaline medium.
conditions which may dissociate association of molecules (forming ag­
gregates) or/and exchanging few remaining Ca2+ counterions in excess
of sodium even if the purification was performed in excess of NaCl.

8
S. da Costa Amaral et al. Food Hydrocolloids 121 (2021) 106845

Table 5
Observed critical concentration for gelation of the Araçá pectin samples in
excess of acid and calcium.
Samples Concentration (mg mL− 1) Addition of

HCl CaCl2

APP 2–10 N N
APP-15 1 D D
2 and 3 D H
APP-30 1 D D
2 and 3 D H
APP-60 1 and 2 D D
3 D H

N: no gel; D: dispersed microgel; H: homogeneous gel.

1996).
Different concentrations of the APP, APP-15, APP-30 and APP-60
samples were solubilized in water. Then, HCl or CaCl2 was added to
observe the gel formation (Table 5).
No gel was obtained with APP, but the de-esterified samples APP-15,
APP-30 and APP-60 were able to start the gelation process at a con­
centration of 1 mg mL− 1 after the addition of HCl or calcium solutions,
Fig. 5. SEC-MALLS-RI elution profile of APP samples: (A) unmodified APP; (B) forming a dispersed microgel.
APP-60. Light scattering (LS-90◦ - green), and refractive index (RI - black) de­ The formation of homogeneous gel was observed for all the de-
tectors. (For interpretation of the references to color in this figure legend, the esterified samples at a concentration of 3 mg mL− 1 after the addition
reader is referred to the Web version of this article.) of CaCl2. This ability to form gels in the presence of calcium is in
agreement with the chemical structure of Araçá pectin which is rich in
GalA blocks.
Therefore, to better understand the ability of the Araçá pectin to give
good gels in presence of calcium, the APP-15 sample was selected for
rheological studies.

3.5. Characterization of gelling properties

In order to study the ability for gelation in presence of calcium ions in


aqueous solution, a gel was prepared with the obtained de-esterified
sample APP-15 dialyzed against 1 mol L− 1 CaCl2 solution. The rheo­
logical behavior of the formed gels was investigated in oscillatory shear
and oscillatory compressed experiments.
A typical oscillatory shear experiment is shown on Fig. 7 where the
Fig. 6. SEC-RI elution profile of APP samples after different times of de- dependence of the elastic (G′ ) and viscous (G′′ ) moduli are plotted as a
esterification reaction in alkaline conditions: unmodified APP (Red); APP-15 function of angular frequency at constant applied oscillatory strain (γ =
(Green); APP-30 (Pink); APP-60 (Brown); APP-105 (Blue). (For interpretation 0.1%). On the all range of investigated angular frequency, the elastic
of the references to color in this figure legend, the reader is referred to the Web modulus (G′ ) is almost 10 times higher than the viscous modulus (G′′ ).
version of this article.) This last one is constant and approximately equal to 5 × 103 Pa at 1 rad/
s, whereas the elastic modulus G′ is equal 2 × 104 at ω = 1 rad/s and
continuously increases with a power law ω0.17. A similar ratio between
Table 4
Mw (g mol− 1) of the unmodified and de-esterified pectins samples obtained by
SEC experiments.
Samples APP APP-15 APP-30 APP-60 APP-105

Mw 480,000 194,000 173,000 176,000 180,000

3.4. Critical concentration for gelation

In Table 5, the observed gelation as a function of pectin concentra­


tions are given in acidic and calcium excesses. For the tested conditions,
the unmodified APP sample was not able to form a gel even at high
polymer concentrations (10 mg mL− 1). This result was expected
considering that this sample presented a high DM (DM = 55.9%). HM
pectins does not contain sufficient blocks of free carboxyl groups to form
gel or precipitate with calcium ions (Sriamornsak, 2003). In this case,
the gelation of HM pectins are mainly imposed by hydrogen bonding and
Fig. 7. Viscoelasticity moduli of an oscillatory shear experiment as a function
hydrophobic interactions in the presence of co-solutes, such as sucrose, of angular frequency of APP-15 at 10 g L− 1 in CaCl2 at 28 ◦ C for an applied
which reduces the water activity, at pH lower than 3.5 that causes oscillatory strain γ = 0.1%. Open points correspond to the viscous modulus G”
protonation of carboxylate residues (Picot-Allain et al., 2020; Rinaudo, (Pa) and full dots correspond to the elastic modulus G’ (Pa).

9
S. da Costa Amaral et al. Food Hydrocolloids 121 (2021) 106845

G’ and G” was also observed by Löfgren, Walkenström, and Hermansson


(2002) on pectin gels. Vincent and Williams (2009) using micro­
rheological measurements investigated and discussed the result of
calcium-mediated junction structures which exhibit the signature of
semi-flexible networks or cross-linked flexible networks as a function of
a parameter proportional to the calcium concentration and inversely
proportional to the degree of methyl-esterification.
A typical oscillatory compression experiment is given on Fig. 8. On
this figure, the loss modulus is fairly constant and equal to E” (Pa) ≈2 ×
104 Pa whereas the storage modulus E′ (Pa) evolves slowly following a
power law ω 0.17 with a value E’≈5.5 × 104 Pa at an oscillatory
compression frequency of 1 rad/s.
Finally, Fig. 9 shows an axial stress evolution as a function of the
compression strain of the gel at 28 ◦ C obtained with rough parallel plates
geometry. The axial stress shows two linear behaviors. Below 2% strain,
a slope EI = 8 × 104 Pa identified as the Young modulus of the gel is
observed. Above 2% strain, the axial stress evolves linearly with a slope
of EII = 2.05 × 104 Pa. Interestingly, when a new oscillatory shear strain
is done after an axial stress compression, viscoelastic moduli still behave
as on Fig. 7.
It is found that for oscillatory compression and oscillatory shear Fig. 9. Axial stress versus strain obtained by a compression test between a
experiments viscous terms are constants and respectively equal to E” = rough plate-plate geometry. Open points were measured on APP-15 in 1 mol
2 × 104 Pa and G” = 5 × 103 Pa (cf. Figs. 7 and 8). Their ratio defines a L− 1 CaCl2 at 28 ◦ C. Dotted lines correspond to best fits obtained onto linear part
of stress versus strain. (I) corresponds to a Young modulus EI = 8 × 104 Pa and
Trouton number Tr = E”/G” equal to 4. As defined here, this Trouton
(II) to a Young modulus EII = 2.05 × 104 Pa.
number can be compared to the Trouton number encountered in the
literature and defined as the ratio of the elongational to the shear vis­
cosities. For example, using an extensional rheometer experiment Ng, Hecht, 1980; Li, Hu, & Li, 1993).
Mun, Boger, and James (1996) found a Trouton ratio equal to 3 for Those main characteristics are the signature of a physical gel in
Newtonian fluids and 3.9 for a non Newtonian polymeric fluid. Tir­ which pectin gel is stabilized by cooperative linkage with calcium
taatmadja and Sridhar (1993) and Amin and Wang (2020) have shown counter-ions.
experimentally and numerically the called Trouton ratio drastically In summary, although the APP (DM ~ 56%) does not form gel in
depends onto polymer molecular weight and extensional strain. Finally, presence of acid or calcium excess (as characteristic of HM pectins), as
our experiment is in good agreement with expected Trouton ratio Tr = 4 soon as DM decreases, it is shown that, over a critical polymer concen­
for a non slippage compression test (Macosko, 1994, p. 286). Moreover, tration around 2 mg mL− 1, the de-esterified pectins APP-15 are able to
Araçá pectin gels behaves as a transient percolated network similarly as form gel in the presence of acid and calcium. These results suggest that
the associative telechelic network described by Amin and Wang (2020). this pectin may be used in the formulation of hydrogels based on calcium
For those reasons but also because our gels keep the same behavior ions interactions for the development of pharmaceutical, cosmetic and
during oscillations test before and after a compression test, we claim that applied products in the biomedical area.
our gels are physical gels.
Regarding the storage moduli E′ and G′ , they both increase slightly 4. Conclusion
with a power law ω 0.17 with E’ ≈ 5.5 × 104 Pa and G’ = 2 × 104 Pa at
respectively ω = 1 rad/s and ω = 1 rad/s. Using the following relation E’ In this work, for the first time, pectins from Araçá (Psidium cat­
= 2(1+ν) G′ which links the Young modulus E′ and the shear modulus tleianum Sabine), a Brazilian fruit pulp, were obtained by water
G’, we obtain a Poisson coefficient of ν = 0.375 which is in good extraction and precipitation with ethanol under sodium form in pres­
agreement with Poisson coefficient for polymeric materials (Geissler & ence of NaCl (APP). From the experimental results discussed in this
work, it is pointed the necessity to associate different techniques to get
valuable chemical structure. Based on titration of GalA units after
completed-methylation, APP consists of around 72% of uronic acid. The
initial degree of methylation is around 56% and then, those pectins may
be classified as a HM pectin with a low degree of acetylation.
The 1H NMR and 1H/13C HSQC NMR analyses confirmed the pres­
ence of homogalacturonans and type I ramnogalacturonans in APP, with
side chains containing arabinogalactans and/or arabinans and gal­
actans. In addition, from NMR studies after partial de-esterification in
alkaline conditions, it is confirmed that the main chain composition is
stable (constant GalA and rhamnose contents), with just the decrease of
methylation and acetylation degrees with time of treatment going to
zero after 60 min of treatment.
On APP de-esterified samples with different DM, firstly, the molec­
ular weight distributions indicate the presence of two ranges of chain
length which remain unchanged independently of the time of alkaline
treatment. Secondly, after partial de-esterification of APP, a homoge­
neous strong physical gel is obtained by dialysis against CaCl2 on APP-
Fig. 8. Viscoelasticity moduli of an oscillatory compression experiment of APP- 15, demonstrating that the Araçá fruit is an interesting raw material to
15 at 10 g L− 1 in CaCl2 at 28 ◦ C. Storage modulus E′ correspond to full dots and
isolate pectins for gelling applications.
loss modulus E′′ correspond to open dots. The oscillatory compression fre­
At end, this work contributes providing a low cost and simple
quency sweep ω (rad/s) was made at constant strain of ε = 0.05%.

10
S. da Costa Amaral et al. Food Hydrocolloids 121 (2021) 106845

method of purification to obtain pure pectins with specifications closer Carbohydrate Polymers, 214, 250–258. https://doi.org/10.1016/j.
carbpol.2019.03.045
to those required commercially.
Berriaud, N., Milas, M., & Rinaudo, M. (1998). Characterization and properties of
hyaluronic acid (hyaluronan). In D. Severian (Ed.), Polysaccharides in medicine and
CRediT authorship contribution statement biotechnology (pp. 313–334). New York: Marcel Dekker.
Biazotto, K. R., Mesquita, L. M. S., Neves, B. V., Braga, A. R. C., Tangerina, M.,
Vilegas, W., et al. (2019). Brazilian biodiversity fruits: Discovering bioactive
Sarah da Costa Amaral: Conceptualization, Data curation, Formal compounds from underexplored sources. Journal of Agricultural and Food Chemistry,
analysis, Funding acquisition, Investigation, Methodology, Project 67(7), 1860–1876. https://doi.org/10.1021/acs.jafc.8b05815
Bittencourt, G. M., Firmiano, D. M., Fachini, R. P., Lacaz-Ruiz, R., Fernandes, A. M., &
administration, Resources, Software, Supervision, Validation, Visuali­
Oliveira, A. L. (2019). Application of green technology for the acquisition of extracts
zation, Writing – original draft, Writing – review & editing. Denis Roux: of araçá (Psidium grandifolium mart. Ex DC.) using supercritical CO2 and pressurized
Conceptualization, Data curation, Formal analysis, Funding acquisition, ethanol: Characterization and analysis of activity. Journal of Food Science, 84(6),
Investigation, Methodology, Project administration, Resources, Soft­ 1297–1307. https://doi.org/10.1111/1750-3841.14584
Blumenkrantz, N., & Asboe-Hansen, G. (1973). New method for quantitative
ware, Supervision, Validation, Visualization, Writing – original draft, determination of uronic acids. Analytical Biochemistry, 54(2), 484–489. https://doi.
Writing – review & editing. François Caton: Conceptualization, Data org/10.1016/0003-2697(73)90377-1
curation, Formal analysis, Funding acquisition, Investigation, Method­ Bradford, M. M. (1976). A rapid and sensitive method for the quantitation of microgram
quantities of protein utilizing the principle of protein-dye binding. Analytical
ology, Project administration, Resources, Software, Supervision, Vali­ Biochemistry, 72(1–2), 248–254. https://doi.org/10.1016/0003-2697(76)90527-3
dation, Visualization, Writing – original draft, Writing – review & Cao, L., Lu, W., Mata, A., Nishinari, K., & Fang, Y. (2020). Egg-box model-based gelation
editing. Marguerite Rinaudo: Conceptualization, Data curation, of alginate and pectin: A review. Carbohydrate Polymers, 242, 116389. https://doi.
org/10.1016/j.carbpol.2020.116389
Formal analysis, Funding acquisition, Investigation, Methodology, Cárdenas, A., Goycoolea, F. M., & Rinaudo, M. (2008). On the gelling behaviour of
Project administration, Resources, Software, Supervision, Validation, ‘nopal’ (Opuntia ficus indica) low methoxyl pectin. Carbohydrate Polymers, 73(2),
Visualization, Writing – original draft, Writing – review & editing. 212–222. https://doi.org/10.1016/j.carbpol.2007.11.017
Cardoso, J. S., Oliveira, P. S., Bona, N. P., Vasconcellos, F. A., Baldissarelli, J.,
Shayla Fernanda Barbieri: Conceptualization, Data curation, Formal Vizzotto, M., et al. (2018). Antioxidant, antihyperglycemic, and antidyslipidemic
analysis, Funding acquisition, Investigation, Methodology, Project effects of Brazilian-native fruit extracts in an animal model of insulin resistance.
administration, Resources, Software, Supervision, Validation, Visuali­ Redox Report, 23(1), 41–46. https://doi.org/10.1080/13510002.2017.1375709
Chan, S. Y., Choo, W. S., Young, D. J., & Loh, X. J. (2017). Pectin as a rheology modifier:
zation, Writing – original draft, Writing – review & editing. Joana Léa
Origin, structure, commercial production and rheology. Carbohydrate Polymers, 161,
Meira Silveira: Conceptualization, Data curation, Formal analysis, 118–139. https://doi.org/10.1016/j.carbpol.2016.12.033
Funding acquisition, Investigation, Methodology, Project administra­ Ciriminna, R., Chavarría-Hernández, N., Hernández, A. I. R., & Pagliaro, M. (2015).
tion, Resources, Software, Supervision, Validation, Visualization, Pectin: A new perspective from the biorefinery standpoint. Biofuels, Bioproducts &
Biorefining, 9, 368–377. https://doi.org/10.1002/bbb.1551
Writing – original draft, Writing – review & editing. Colodel, C., Vriesmann, L. C., Teófilo, R. F., & Petkowicz, C. L. O. (2018). Extraction of
pectin from ponkan (Citrus reticulata blanco cv. Ponkan) peel: Optimization and
structural characterization. International Journal of Biological Macromolecules, 117,
Declaration of competing interest 385–391. https://doi.org/10.1016/j.ijbiomac.2018.05.048
Denardin, C. C., Hirsch, G. E., da Rocha, R. F., Vizzotto, M., Henriques, A. T.,
Moreira, J. C. F., et al. (2015). Antioxidant capacity and bioactive compounds of four
There is no conflict of interest.
Brazilian native fruits. Journal of Food and Drug Analysis, 23(3), 387–398. https://
doi.org/10.1016/j.jfda.2015.01.006
Acknowledgements Dias, I. P., Barbieri, S. F., López-Fetzer, D. E., Corazza, M. L., & Silveira, J. L. M. (2020).
Effects of pressurized hot water extraction on the yield and chemical
characterization of pectins from Campomanesia xanthocarpa Berg fruits. International
The authors gratefully acknowledge the following Brazilian agencies Journal of Biological Macromolecules, 146, 431–443. https://doi.org/10.1016/j.
for financial support: the Coordination for the Improvement of Higher ijbiomac.2019.12.261
Dranca, F., & Oroian, M. (2018). Extraction, purification and characterization of pectin
Education Personnel – CAPES, the National Council for Scientific and
from alternative sources with potential technological applications. Food Research
Technological Development – CNPq, and the Federal University of International, 113, 327–350. https://doi.org/10.1016/j.foodres.2018.06.065
Paraná - Brazil. J.L.M.S. is a research member of the CNPq Foundation Dubois, M., Gilles, K. A., Hamilton, J. K., Rebers, P. A., & Smith, F. (1956). Colorimetric
Method for determination of sugars and related substances. Analytical Chemistry, 28
and UFPR (nº 476950/2013-9; 308296/2015-0; 309225/2018-3); S.C.A
(3), 350–356. https://pubs.acs.org/doi/10.1021/ac60111a017.
is the beneficiary of a post-graduation scholarship (nº 141692/2018-9) Emporio Laszlo. (2021). https://www.emporiolaszlo.com.br/oleo-essencial-de-araca-ro
provided by CNPQ and a Sandwich scholarship at the University Gre­ sa.html Accessed in 2021/03/02.
noble Alpes provided by CAPES-PRINT - RIBBBA-UFPR (nº Fracasso, A. F., Perussello, C. A., Carpiné, D., Petkowicz, C. L. D. O., & Haminiuk, C. W. I.
(2018). Chemical modification of citrus pectin: Structural, physical and rheologial
88887.467969/2019-00); and S.F.B. is the beneficiary of a post-doctoral implications. International Journal of Biological Macromolecules, 109, 784–792.
scholarship from Coordination of Superior Level Staff Improvement – https://doi.org/10.1016/j.ijbiomac.2017.11.060
CAPES (nº 88887.335103/2019-00). The authors would like to thank E. Franzon, R. C., de Campos, L. Z. O., Proença, C. E. B., & Sousa-Silva, J. C. (2009). Araçás
do gênero Psidium: Principais espécies, ocorrência, descrição e usos – documentos 266.
Bayma from PCANS technical platform at CERMAV (CNRS) for SEC Embrapa Cerrados.
experiments, I. Jeacomine from the NMR Centers of RMN-ICMG Gawkowska, D., Cybulska, J., & Zdunek, A. (2018). Structure-related gelling of pectins
(FR2607) - Grenoble-France, A. Santana Filho and G. Sassaki from and linking with other natural compounds: A review. Polymers, 10(7), 762. https://
doi.org/10.3390/polym10070762
UFPR-Curitiba-Brazil for recording the NMR spectra, K. Mischiatti and Geissler, E., & Hecht, A. M. (1980). The Poisson ratio in polymer gels. Macromolecules, 13
R. Bagatin for HPAEC-PAD and GLC analysis, respectively, the Brazilian (5), 1276–1280. https://doi.org/10.1021/ma60077a047
Agricultural Research Corporation/Embrapa Forests, and M. Mazza for Grasdalen, H., Bakøy, O. E., & Larsen, B. (1988). Determination of the degree of
esterification and the distribution of methylated and free carboxyl groups in pectins
providing the Araçá fruits.
by 1H NMR spectroscopy. Carbohydrate Research, 184, 183–191. https://doi.org/
10.1016/0008-6215(88)80016-8
References Haminiuk, C. W. I., Sierakowski, M. R., Vidal, J. R. M. B., & Masson, M. L. (2006).
Influence of temperature on the rheological behavior of whole araçá pulp (Psidium
cattleianum Sabine). LWT-Food Science and Technology, 39(4), 427–431. https://doi.
Amaral, S. C., Barbieri, S. F., Ruthes, A. C., Bark, J. M., Winnischofer, S. M. B., &
org/10.1016/j.lwt.2005.02.011
Silveira, J. L. M. (2019). Cytotoxic effect of crude and purified pectins from
Khotimchenko, M. (2020). Pectin polymers for colon-targeted antitumor drug delivery.
Campomanesia xanthocarpa Berg on human glioblastoma cells. Carbohydrate
International Journal of Biological Macromolecules, 158, 1110–1124. https://doi.org/
Polymers, 224, 115140. https://doi.org/10.1016/j.carbpol.2019.115140
10.1016/j.ijbiomac.2020.05.002
Amin, D., & Wang, Z. (2020). Nonlinear rheology and dynamics of supramolecular
Kravtchenko, T. P., Voragen, A. G. J., & Pilnik, W. (1992). Analytical comparison of three
polymer networks formed by associative telechelic chains under shear and
industrial pectin preparations. Carbohydrate Polymers, 18, 17–25. https://doi.org/
extensional flows. Journal of Rheology, 64(3), 581–600. https://doi.org/10.1122/
10.1016/0144-8617(92)90183-Q
1.5120897
Li, Y., Hu, Z., & Li, C. (1993). New method for measuring Poisson’s ratio in polymer gels.
Barbieri, S. F., Amaral, S. C., Ruthes, A. C., Petkowicz, C. L. O., Kerkhoven, N. C.,
Journal of Applied Polymer Science, 50(6), 1107–1111. https://doi.org/10.1002/
Silva, E. R. A., et al. (2019). Pectins from the pulp of gabiroba (Campomanesia
app.1993.070500619
xanthocarpa Berg): Structural characterization and rheological behavior.

11
S. da Costa Amaral et al. Food Hydrocolloids 121 (2021) 106845

Lima, A. S., Maia, D. V., Haubert, L., Oliveira, T. L., Fiorentini, A. M., Rombaldi, C. V., Rahelivao, M. P., Andriamanantoanina, H., Heyraud, H., & Rinaudo, M. (2013). Structure
et al. (2020). Action mechanism of araçá (Psidium cattleianum Sabine) hydroalcoholic and properties of three alginates from Madagascar seacoast algae. Food
extract against Staphylococcus aureus. Lebensmittel-Wissenschaft und -Technologie- Hydrocolloids, 32, 143–146. https://doi.org/10.1016/j.foodhyd.2012.12.005
Food Science and Technology, 119, 108884. https://doi.org/10.1016/j. Ralet, M.-C., Dronnet, V., Buchholt, H. C., & Thibault, J.-F. (2001). Enzymatically and
lwt.2019.108884 chemically de-esterified lime pectins: Characterisation, polyelectrolyte behaviour
Löfgren, C., Walkenström, P., & Hermansson, A. M. (2002). Microstructure and and calcium binding properties. Carbohydrate Research, 336(2), 117–125. https://
rheological behavior of pure and mixed pectin gels. Biomacromolecules, 3(6), doi.org/10.1016/S0008-6215(01)00248-8
1144–1153. https://doi.org/10.1021/bm020044v Rinaudo, M. (1996). Physicochemical properties of pectins in solution and gel states. In
Ma, X., Chen, W., Yan, T., Wang, D., Hou, F., Miao, S., et al. (2020). Comparison of citrus J. Visser, & A. G. J. Voragen (Eds.), Pectins and pectinases (pp. 237–252). New York,
pectin and apple pectin in conjugation with soy protein isolate (SPI) under USA.
controlled dry-heating conditions. Food Chemistry, 309, 125501. https://doi.org/ Rinaudo, M. (2005). Advances in characterization of polysaccharides in aqueous solution
10.1016/j.foodchem.2019.125501 and gel state. In S. Dimitriu, & M. Dekker (Eds.), Polysaccharides: Structural diversity
Macosko, C. W. (Ed.). (1994). Rheology: Principles, measurements and applications. New and functional versatility (pp. 237–252). New York, USA.
York, NY: Wiley-VCH. Chap. 7 Extensional rheometry. Schneider, V. S., Iacomini, M., & Cordeiro, L. M. C. (2019). β-L-Araf-containing arabinan
Malovikova, A., Rinaudo, M., & Milas, M. (1993). On the characterization of and glucuronoxylan from guavira fruit pomace. Carbohydrate Research, 481, 16–22.
polygalacturonate salts in dilute solution. Carbohydrate Polymers, 22(2), 87–92. https://doi.org/10.1016/j.carres.2019.06.005
https://doi.org/10.1016/0144-8617(93)90070-K Sriamornsak, P. (2003). Chemistry of pectin and its pharmaceutical uses: A review.
Mankarios, A. T., Jones, C. F. G., Jarvis, M. C., Threfall, D. R., & Friend, J. (1979). University International Journal, 3(1–2), 206–228.
Hydrolysis of plant polysaccharides and GLC analysis of their constituent neutral Stephen, A. M., Phillips, G. O., & Williams, P. A. (2006). Food polysaccharides and their
sugars. Phytochemistry, 18(3), 419–422. https://doi.org/10.1016/S0031-9422(00) applications. United States of America: CRC Press.
81879-8 Taboada, E., Fisher, P., Jara, R., Zúñiga, E., Gidekel, M., Cabrera, J. C., et al. (2010).
Medina, A. L., Haas, L. I. R., Chaves, F. C., Salvador, M., Zambiazi, R. C., Da Silva, W. P., Isolation and characterisation of pectic substances from murta (Ugni molinae Turcz)
et al. (2011). Araçá (Psidium cattleianum Sabine) fruit extracts with antioxidant and fruits. Food Chemistry, 123(3), 669–678. https://doi.org/10.1016/j.
antimicrobial activities and antiproliferative effect on human cancer cells. Food foodchem.2010.05.030
Chemistry, 128(3), 916–922. https://doi.org/10.1016/j.foodchem.2011.03.119 Thibault, J. F., & Rinaudo, M. (1985). Interactions of mono- and divalent counterions
Miller, G. L. (1959). Use of dinitrosalicylic acid reagent for determination of reducing with alkali- and enzyme-deesterified pectins in salt-free solutions. Biopolymers, 24,
sugar. Analytical Chemistry, 31, 426–428. https://doi.org/10.1021/ac60147a030 2131–2143. https://doi.org/10.1002/bip.360241109
Munarin, F., Tanzi, M. C., & Petrini, P. (2012). Advances in biomedical applications of Tirtaatmadja, V., & Sridhar, T. (1993). A filament stretching device for measurement of
pectin gels. International Journal of Biological Macromolecules, 51(4), 681–689. extensional viscosity. Journal of Rheology, 37(6), 1081–1102. https://doi.org/
https://doi.org/10.1016/j.ijbiomac.2012.07.002 10.1122/1.550372
Nagel, A., Sirisakulwat, S., Carle, R., & Neidhart, S. (2014). An acetate hydroxidegradient Ventura, I., Jammal, J., & Bianco-Peled, H. (2013). Insights into the nanostructure of
for the quantitation of the neutral sugar and uronic acid profile ofpectins by HPAEC- low-methoxyl pectin–calcium gels. Carbohydrate Polymers, 97(2), 650–658. https://
PAD without postcolumn pH adjustment. Journal of Agricultural and Food Chemistry, doi.org/10.1016/j.carbpol.2013.05.055
62(9), 2037–2048. https://doi.org/10.1021/jf404626d Vincent, R. R., & Williams, M. A. (2009). Microrheological investigations give insights
Naqash, F., Masoodi, F. A., Rather, S. A., Wani, S. M., & Gani, A. (2017). Emerging into the microstructure and functionality of pectin gels. Carbohydrate Research, 344
concepts in the nutraceutical and functional properties of pectin-A review. (14), 1863–1871. https://doi.org/10.1016/j.carres.2008.11.021
Carbohydrate Polymers, 168, 227–239. https://doi.org/10.1016/j. Vinholes, J., Lemos, G., Barbieri, R. L., Franzon, R. C., & Vizzotto, M. (2017). In vitro
carbpol.2017.03.058 assessment of the antihyperglycemic and antioxidant properties of araçá, butiá and
Ng, S. L., Mun, R. P., Boger, D. V., & James, D. F. (1996). Extensional viscosity pitanga. Food Bioscience, 19, 92–100. https://doi.org/10.1016/j.fbio.2017.06.005
measurements of dilute solutions of various polymers. Journal of Non-newtonian Fluid Voragen, A. G. J., Coenen, G.-J., Verhoef, R. P., & Schols, H. A. (2009). Pectin, a versatile
Mechanics, 65(2–3), 291–298. https://doi.org/10.1016/0377-0257(96)01463-2 polysaccharide present in plant cell walls. Structure chemistry, 20, 263–275. https://
Ngouémazong, D. E., Nkemamin, N. F., Cardinaels, R., Jolie, R. P., Fraeye, I., doi.org/10.1007/s11224-009-9442-z
Loey, A. M. V., et al. (2012). Rheological properties of Ca2+-gels of partially Vriesmann, L. C., Petkowicz, C. L. O., Carneiro, P. I. B., Costa, M. E., & Beleski-
methylesterified polygalacturonic acid: Effect of “mixed” patterns of Carneiro, E. (2009). Acidic polysaccharides from Psidium cattleianum (araçá).
methylesterification. Carbohydrate Polymers, 88(1), 37–45. https://doi.org/10.1016/ Brazilian Archives of Biology and Technology, 52, 259–264. https://doi.org/10.1590/
j.carbpol.2011.11.051 S1516-89132009000200001
Noreen, A., Nazlic, Z.-I.-U., Akrama, J., Rasulb, I., Manshaa, N. Y., Iqbald, R., et al. Wolfrom, M. L., & Thompson, A. (1963a). Reduction with sodium borohydride. In
(2017). Pectins functionalized biomaterials, a new viable approach for biomedical R. L. Whistler, M. L. Wolfrom, & J. N. BeMiller (Eds.), Methods in carbohydrate
applications: A review. International Journal of Biological Macromolecules, 101, chemistry (pp. 65–68). New York and London: Academic Press Inc.
254–272. https://doi.org/10.1016/j.ijbiomac.2017.03.029 Wolfrom, M. L., & Thompson, A. (1963b). Acetylation. In R. L. Whistler, M. L. Wolfrom,
Patel, S. (2012). Exotic tropical plant Psidium cattleianum: A review on prospects and & J. N. BeMiller (Eds.), Methods in carbohydrate chemistry (pp. 211–215). New York
threats. Reviews in Environmental Science and Biotechnology, 11, 243–248. https://doi. and London: Academic Press Inc.
org/10.1007/s11157-012-9269-8 Yapo, B. M. (2011). Pectic substances: From simple pectic polysaccharides to complex
Patova, O. A., Smirnov, V. V., Golovchenko, V. V., Vityazev, F. V., Shashkov, A. S., & pectins - a new hypothetical model. Carbohydrate Polymers, 86(2), 373–385. https://
Popov, S. V. (2019). Structural, rheological and antioxidant properties of pectins doi.org/10.1016/j.carbpol.2011.05.065
from Equisetum arvense L. and Equisetum sylvaticum L. Carbohydrate Polymers, 209, Zhang, H., Chen, J., Li, J., Yan, L., Li, S., Ye, et al. (2018). Extraction and characterization
239–249. https://doi.org/10.1016/j.carbpol.2018.12.098 of RG-I enriched pectic polysaccharides from Mandarin citrus peel. Food
Pereira, E. S., Vinholes, J., Franzon, R. C., Dalmazo, G., Vizzotto, M., & Nora, L. (2018). Hydrocolloids, 79, 579–586. https://doi.org/10.1016/j.foodhyd.2017.12.002
Psidium cattleianum fruits: A review on its composition and bioactivity. Food Zheng, J., Chen, J., Zhang, H., Wu, D., Ye, X., Linardt, R. J., et al. (2020). Gelling me.
Chemistry, 258, 95–103. https://doi.org/10.1016/j.foodchem.2018.03.024 chanism of RG-I enriched citrus pectin: Role of arabinose side-chains in cation-and
Picot-Allain, M. C. N., Ramasawmy, B., & Emmambux, M. N. (2020). Extraction, acid-induced gelation. Food Hydrocolloids, 101, 105536. https://doi.org/10.1016/j.
characterisation, and application of pectin from tropical and sub-tropical fruits: A foodhyd.2019.105536
review. Food Reviews International, 1–31. https://doi.org/10.1080/
87559129.2020.1733008

12

You might also like