Open access peer-reviewed chapter

New Sources of Pectin: Extraction, Processing, and Industrial Applications

Written By

Stefani Cortés-Camargo, Angélica Román-Guerrero, Erik Alpizar-Reyes and César Pérez-Alonso

Submitted: 23 November 2022 Reviewed: 17 December 2022 Published: 13 January 2023

DOI: 10.5772/intechopen.109579

From the Edited Volume

Utilization of Pectin in the Food and Drug Industries

Edited by Maruf Ahmed

Chapter metrics overview

512 Chapter Downloads

View Full Metrics

Abstract

One of the most important polysaccharides in the vegetal kingdom is pectin. This class of natural polysaccharide is found primarily in citrus fruits and apple pomace. Pectin has been used in different sectors of the industry, among which the food, pharmaceutical, cosmetic, and paper industries stand out. Today, there is a growing demand for this type of hydrocolloid, where both the scientific and industrial fields have focused on using new sources of pectin and developing novel extraction methods. This chapter describes the chemical structure of pectin and its main chemical characteristics. Then, the conventional sources from which pectin is obtained are exposed as well as its main industrial applications. Subsequently, the physicochemical and functional properties of pectins obtained from unconventional sources are described and analyzed as well as the main technologies used for their extraction. Finally, the most recent advances in the role played by pectin in the industrial sector are described.

Keywords

  • pectin
  • extraction
  • functional properties
  • husks
  • hulls
  • Cactaceae
  • new applications

1. Introduction

Pectin is considered one of the main polysaccharides found in plant sources; it participates in the constitution of cell walls of higher plants, impacting the physical and nutritional contribution of products of plant origin. Pectin is a globally recognized polysaccharide with great relevance in the global biopolymer market due to its inherent functional properties and vast applications in the food, pharmaceutical, and biomedical industries [1]. It is a macromolecule capable of forming flexible polymer chains that lead to forming hydrogel-type structures [2]. Its functional properties are associated with the extraction conditions and influenced by the source used. The primary sources of commercial pectin are citric fruits and apples; however, non-conventional sources have been investigated, such as agro-industrial sub-products and residues, pulps, husks, hulls, peels, Cactaceae, and vegetables, among others [3]. Furthermore, pectin has been functionalized through chemical or enzyme reactions that lead to changes and improvements in its physicochemical properties, such as molecular weight, degree of esterification (DE), and surface charge, which in turn contributes to the development of new functional or improved properties, along with new approaches and applications [4].

Advertisement

2. Pectin structure

Pectin is a negatively charged branched heteropolysaccharide, composed of up to 17 different monosaccharides with more than 20 types of linkages [5, 6]. This polysaccharide was first reported in 1825 by Braconnot and defined as a biopolymer rich in galacturonic acid (GalA; up to 65%) [7]. Although the precise structure of pectin has not yet been fully elucidated due to its complexity, three major polysaccharide domains are recognized; as shown in Figure 1, the most abundant is based on a linear homopolymer of α-(1–4)-linked-D-galacturonic acid (GalpA, GalA) residues that can be methyl esterified at the C-6 position and to a lesser extent O-acetylated in C-2 and C-3; this domain is defined as homogalacturonan (HG) [5, 7]. In the rhamnogalacturonan I (RG-I) domain, the rhamnose (Rhap, Rha) residues disrupt the HG structure to form a preferably ramified structure of pectin (20–35%) due to the presence of the repeating disaccharide [→4)-α-D-GalpA-(1→2)- α-L-Rhap-(1-]. Here, the GalA residues are not methyl esterified, and attachment of neutral sugar side chains [α-L-arabinose (Araf, Ara) and β-D-galactose (Galp, Gal)] to the C-4 positions of Rha residues can be suitable, leading to linear side chains (LC-A) when α(1→5)-L-Araf or linear type I (β(1→4)-L- Galp) or branched side chains (LC-B) when α(1→2,3)-L-Araf or branched type II β(1→3,6)-D-Galp and arabinogalactans. The branching design of the structure in RG-I depends on the pectin source, the extraction conditions, and the presence of other sugars such as xylose (D-Xyl), fucose (L-Fuc), and glucuronic acid (D-GlucA), among others [9]. The RG-II domain (1–8%) is constituted of around nine α(1→4)-linked GalpA units partially methyl esterified with four heteropolymer side chains attached, mainly composed of 11 monosaccharide residues, including apiose (D-Api), 2-O-methyl-L-fucose, 2-O-methyl-D-xylose. 3-C-carxy-5deoxy-L-xylose, 3-deoxy-D-manno-octulosonic acid (Kdo), and 3-deoxy-D-lyxoheptulosaric acid (D-Dha), which are linked with up to 22 glycoside bonds [10, 11].

Figure 1.

The schematic representation of the pectin structure contains the HG, RG-I, and RG-II domains. L-AcA: L-Aceric acid. Adapted from [8].

Some investigations about the basic structure of pectin establish that although the pectin source may influence the structure diversity by partially modifying the chain conformation of the macromolecule, the RG-II region seemed to be well preserved among the different sources [12]. Moreover, pectins contain functional groups besides carbohydrate type, such as phenolic acids, methanol, acetic acid, and some amide groups. Methanol and acetic acid are relevant in the esterification of galacturonic acid residues for developing the inherent structure functionalities of pectin. The degree of methylation (DM) is a helpful tool for describing the structure of pectin and potential applications; high methoxy pectins (HM) contain more than 50% of carboxyl groups in methylated form, while those with lower content are defined as low methoxy pectins (LM). Most common native pectins are characterized by being methyl esterified. Likewise, acetylation in pectins rarely occurs in native pectins. The degree of acetylation (DA) in pectins is defined as the percentage of galacturonosyl residues that can be acetylated per unit of monosaccharide. DA can be larger than 100% and is usually found in the branched RG regions. In pectins from citrus and apple, the acetyl groups in the HG region are present in low content, rather than in pectins from sugar beet and potato, where higher amounts have been found [13, 14]. Amidation of pectins does not occur naturally; instead, it is induced chemically or enzymatically to improve the functional properties such as solubility in water, gelling, and rheological properties through modifying some non-esterified carboxyl groups into amide groups by using various amino compounds [15, 16, 17].

Advertisement

3. Conventional sources of pectin and their applications

Commercial pectin is generally obtained from citrus peels (25% dry matter) and apple pomace (15–18% dry matter), their processing subproducts, and sugar beet pulp (25% dry matter) [18]. The most significant part of commercial pectins includes 85.5% from citrus peels, 14% from apple pomace, and ~ 0.5% from sugar beet pulp [19]. Industrial processes for the extraction of pectin are based on the thermal hydrolysis of the citric peels (mainly from orange, lemon, and lime), apple pomace, and sugar beet pulp by using hot mineral acids like HCl, H2SO4, or HNO3 (~pH 1.5) at ~85°C [20], where the control of the extraction conditions is of great relevance for minimizing the de-esterification and depolymerization of the polysaccharide and improving the functional properties of pectins as gelling, fiber enrichment, stabilizer, texture, and rheology control agent [21]. Notably, these pectin extraction processes generate large amounts of acidic industrial wastes and high energy consumption [22]. Hence, recent investigations have explored the use of more green technologies to overcome these environmental issues and enhance yield extraction [23]. Table 1 shows some physicochemical properties of pectins obtained from conventional sources using different extraction procedures.

SourceExtraction conditions and yieldFunctional PropertiesReferences
Citrus peelLimeHHP extraction
Enzyme treatment
pH: 4.5; 50°C, 4 h.
Yield: 26.1%
DE: 75.7%; GalA: 82.8%
MW:308.4 kDa; ηint 5.0 dL/g
Enhanced solubilization
Shorter gel setting time
[23]
Enzyme treatment
pH 3.5; 50 °C, 4 h.
Yield: 23%
DE: > 82%; GalA:81–84%
MW: 69 kDa; Gelling properties
[24]
OrangeUSE: 150 W, 20 kHz, 10 min, 50 °C
Citric acid pH 1.5
Yields: 28.1%
GalA: 72%; DE: 37.5%
Surface tension: 42.1–46.6 mN/m (0.1–0.5% wt.)
WHC: 3.10 gwater/gpectin
OHC:1.32 goil/gpectin
[25]
PomeloConventional extraction
Nitric acid pH 2.0; 90 °C; 1.5 h
Yield: 23.19%
DE: 57.9%; MW: 353 kDa
Viscoelastic solution (<1% wt.)
Weak gelation (<1% wt.)
Newtonian behavior (<0.4% wt.)
Pseudoplastic behavior (>0.4% wt.)
[26]
MandarinHHP: 500 MPa, 10 min
Citric acid pH 1.4
Yield: 21.95%
GalA: 75.4–84.4%; DE: 67.7–70.4%
MW:1201–2626 kDa
Pseudoplastic behavior (3% wt.)
[27, 28]
GrapefruitUSE:800 W;20 kHz; 58% amplitude
67.8 °C; 30 min
Yield: 27.3%
DE: 65.5%; GalA: 50%
ηint: 3.26 dL/g; MW: 109.5 kDa
[29, 30]
Conventional extraction
H2SO4 pH; 80 °C 1 h.
Yield:33.6%
DE: 71.7%; ηint: 18.36 dL/g
MW 2.3x105 kDa
[31]
Apple pomaceGranny Smith; Royal; and Golden varieties
Pomace used for commercial pectin
Chemical or enzyme treatment
Yield:4.2–19.8%*
GalA:18.0–67.9%*; DE:52.5–76.4% *
DM:58–88%; MW:68–790 kDa *
[32, 33, 34]
Sugar beet pulpConventional extraction
HCl, pH 1.2, 90 °C, 3 h.
DE: 38.6–40.8%
Pseudoplastic behavior (2%wt.)
Emulsifying activity
[35, 36]

Table 1.

Physicochemical properties of some pectin extracted from conventional sources.

Depending on the variety of pomace.


HHP: High hydrostatic pressure; MW: Molecular weight; ηint: Intrinsic viscosity; USE: Ultrasound-assisted extraction; WHC: Water holding capacity; OHC: Oil holding capacity.

Pectin is widely used in the food industry as an excellent thickener agent for producing jellies and jams, a pH stabilizer in dairy products and low-calorie products, and an emulsifier in pharmaceutics for the design of drugs to treat gastrointestinal disorders, blood cholesterol reduction, and cancer treatment as well as good former of edible films and coatings, foams, and paper substitutes [17, 24]. Because of the functional properties of pectins, both LM and HM, many applications in food, industrial, and pharmaceutic sectors can be considered (Figure 2).

Figure 2.

Principal applications of commercial pectin in food, packaging, and pharmaceutical industries [17].

Most commercial pectins are facilitated to dissolution when a dextrose content is present. An additional pectin classification is based on its gelling capacity, which is relevant in product processing and preservation. Pectins are classified as rapid-set pectin, when gels are formed, preferably at high temperatures, generally used for jams because it reduces the possibility that the fruit rises to the surface before the pectin gel is set, and slow-set pectin, which is preferred in jellies because it allows handling the product before the gel setting without damaging the texture and firmness of the product [25].

Despite the presence of extensive contents of pectin polysaccharide in fruit subproducts, like citrus peels, apple pomace, or sugar beet pulp, it is not the most critical parameter to define a lucrative extraction and the best functional properties for this functional agent [17]; the exploration of novel sources of pectins is raising the attention of scientists and technologists.

Advertisement

4. Vegetable sources of pectin

New sources of pectin that are receiving significant interest in the scientific field are those obtained from different kinds of vegetables, such as pumpkin, eggplant, chayote, and Opuntia ficus indica cladodes.

Several studies have been conducted on the extraction of pumpkin pectin using different extraction methods, such as the chemical acid treatment (0.1 M HCl) or enzymatic extraction, where the last has given much higher yields than the acid extraction [26]. Pumpkin pectin fraction A was obtained from raw pumpkin with an enzyme preparation of cellulase and α-amylase. Pumpkin pectin fraction B was obtained by treating the solution of fraction A with pronase to reduce the protein content. The pumpkin pectin fractions A and B yielded 10.03 and 8.08 g/100 g, respectively.

The DE values of about 47% for pumpkin pectin fractions A and B were not significantly different, while the GalA contents represent 75.02 and 78.22 g/100 g, respectively. This finding indicated that both fractions are mainly composed of HG [27]. Small amounts (about 10 g/100 g) of six different neutral sugars were found in both pectin fractions, including rhamnose, arabinose, galactose, glucose, xylose, and mannose.

FT-IR and 1D NMR analyses revealed that the pumpkin pectin backbone is mainly composed of 1,4-D galacturonic acid, in which a considerable portion of galacturonic acid residues is present as methyl esters, and L-rhamnose is involved in the linear region of the backbone through α-1,2 linkages. The emulsifying capacity and stability of pumpkin pectin fraction A were 63.7 and 58.3%, respectively. At the same time, both properties were not detected in pumpkin pectin fraction B. Pectin fraction A exhibited emulsifying properties in the water–oil mixture, evidencing the presence of hydrophobic protein components in the pectin structure. In contrast, protein removal in fraction B resulted in a loss of emulsifying properties [26]. Therefore, pumpkin pectin could be used as an emulsifying agent in the preparation of oil-in-water emulsions for the beverage industry as long as residual hydrophobic protein components are not removed.

Eggplant fruit (Solanum melongena L.), a popular vegetable with an elongated oval shape and dark purple peels, grows worldwide, especially in tropical and subtropic regions. Under optimal extraction conditions by the ultrasound-assisted extraction method (UAE) (ultrasound power of 50 W, irradiation time of 30 min, and pH of 1.5), the pectin extracted from the peels of this vegetable (EPP) indicated that the EPP had a high GalA content (66.08 g/100 g) [28]. Considering the Food and Agriculture Organization (FAO) and European Union recommendations, the GalA content of pectin used as a food additive or pharmaceutical purpose should not be lower than 65 g/100 g pectin. This pectin had a high DE (61.22%) and was categorized as HM pectin (DE > 50%). EPP had a protein content of 2.53 g/100 g, which can be attributed to the difference in raw materials and extraction techniques. However, FAO suggests that the protein content of pectin should not be higher than 15.6 g/100 g [24]. In addition, EPP showed good values in functional features such as water-holding capacity (WHC) and oil-holding capacity (OHC). Under the optimal extraction conditions, EPP exhibited a WHC of 6.22 ± 0.21 g water per g EPP, while the OHC was 2.12 ± 0.15 g oil per g EPP. The emulsifying activity (EA) and emulsifying stability (ES) of EPP were evaluated, EA was about 56.16%, and the highest emulsion stability was 96.36 ± 0.80 at 4°C. EPP also exhibited antioxidant activity, determined by the DPPH radical scavenging method, reaching a highest antioxidant activity at a concentration of 50 mg/mL (94%), which was still lower than the antioxidant activity performed by the ascorbic acid, with an IC50 value of 1.39 mg/mL; this activity is due to the higher total phenolic content (TPC = 96.81 ± 2.18 mg GAEa/g pectin) associated to the EPP. The GalA content of the extracted pectin can be also effective in the antioxidant activity due to active portions in its structure [29].

Chayote is one of the most cultivated vegetables in the world. The major producing countries are Mexico, Brazil, and China [30]. The UAE method has been used to extract chayote pectin (PEUO) [31]. Using a liquid/solid ratio of 50 mL/g, a temperature of 70°C, and an ultrasonic time of 40 min as optimal extraction conditions. The yield was around 6.19%. Under these extraction conditions, PEUO exhibited a low DE (17.6%), indicating that the chayote pectin could be considered as LM pectin. This property could be attributed to the harsh extraction conditions that would promote the de-esterification of polygalacturonic chains. The GalA content in PEUO accounted for 57.25%. To our knowledge, the ripeness, blanching, ultrasound, and other effects may influence the GalA content in the extracted pectin [31], besides the contribution to improve the depolymerization of polysaccharides, releasing the water-soluble pectin from the plant tissue [32]. The molecular weight in pectins significantly affects the emulsification, rheology, and their colloid stability. In this sense, the weight-average molecular weight and number-average molecular weight of PEUO were 2.47 × 106 g/mol and 1.29 × 106 g/mol, respectively, and the polydispersity index was 1.91. Polydispersity index higher than 1 suggests that PEUO extracted by UAE represents a heterogeneous natural polysaccharide with a broad range of polymer size distribution [31]. The monosaccharide composition of PEUO indicated the presence of five monosaccharides, where glucose (Glu) represents the most abundant monosaccharide (90.6%), followed by Gal (8%), D-Xyl (0.6%), Ara (0.6%), and Rha (0.2%). Besides, the content of Gal was significantly higher than that of Ara, indicating that the RG-I region may have been highly branched with galactan or arabinogalactan. Rheological properties of PEUO aqueous dispersions (<5%wt.) exhibited a non-Newtonian behavior [31]. Other functional properties like WHC and OHC for PEUO showed suitable values for both WHC (3.14 ± 0.42 g water/g PEUO) and OHC (3.73 ± 0.30 g oil/g PEUO). High WHC in PEUO makes it suitable as a food industry thickener. EA and ES were determined at 4°C and 25°C. The ES for PEUO emulsions were 88.36 ± 5.63% and 81.28 ± 4.82% after 1 day, and these values changed after 30 days to 85.33 ± 4.16% and 77.59 ± 5.19%, respectively. The lower temperature (4°C) was presumably more suitable for storing the PEUO emulsion. These results provide further evidence that chayote pectin may have great potential to be applied as an emulsifier and stabilizer in the food industry [31, 33]. Regarding the antioxidant activity of PEUO, it was higher when compared to pectin extracted from apples. Due to its techno-functional properties, PEUO may be used as a gelling agent and preservative in jam production or as a viscosity enhancer in beverages.

Another source of pectin that has received much attention is the Opuntia ficus indica (OFI) cladodes. This pectin has been extracted by acid water, ultrasound, and enzyme treatments [34, 35]. The pectin obtained by ultrasound under optimal conditions (sonication time of 70 min, temperature of 70°C, pH of 1.5, and water:solid ratio of 30 mL/g) reached an extraction yield of 18.14% ± 1.41%, with a GalA content of 68.87%. This pectin had a DE of 41.42%, classifying it as an LM pectin [36]. This DE value was higher than that achieved when the OFI pectin was extracted by the chemical process, which was 30.67% [37]. WHC in OFI pectin was 4.84 g water/g OFI pectin, ultrasound-induced cavitations in the pectin structure improving the water penetration and its absorption [38]. WHC for OFI pectin extracted by the chemical process was higher (5.64 g water/g OFI pectin) [34]. OHC for OFI pectin extracted with ultrasound was 1.01 g oil/g OFI pectin, slightly lower than pectin extracted by the chemical method (1.24 g oil/g OFI pectin) [34]. EA and ES were determined at two pectin concentrations (2 and 4% w/v). EA values were 19.23% and 26.92%, respectively, showing that the emulsion stability depends on the pectin concentration. OFI pectin at 4% maintained stability of more than 57% of the emulsion after 30 min of incubation at 80°C, unlike the 2% pectin solution, which could not retain more than 40% of the emulsion. This stability of the emulsions could be attributed to the rise of viscosities in the pectin solutions caused by the formation of a layer of pectin around each oil droplet, delaying the coalescence phenomenon [39, 40]. This stability was affected by the high pectin extraction temperature (> 45°C) [41]. ES in OFI pectin extracted with acid water at 2% displayed higher values (90.45%) [34]. Differences in the ES are due to differences in the extraction methods, which affect the average molecular weight and the GalA content in the structure of pectin and therefore influencing the long-term stability in the emulsions [42]. When enzyme treatments were used for OFI pectin extraction, the optimal conditions were cellulase/xylanase at an LS ratio of 22 mL/g, cellulase/xylanase ratio of 2 U/U, and enzymes/matter ratio of 4 U/g, reaching an extraction yield of 17.91% [35], being more effective than the chemical treatment, which resulted in an extraction yield of 6.13 ± 0.60% [34]. Enzyme-assisted extraction of pectin depends on the choice of enzymatic activities based on the strength of pectin connection with cellulose and xylan and their abundance in the cell wall of the plant source [43].

For OFI pectin, the total sugar content was 89.94%, the main monosaccharide was GalA (66.66 ± 2.46%), with a DE of 35.04%, which was higher than that reported by Lira-Ortiz et al. [44] for pectin from prickly pear fruits (Opuntia albicarpa; DE 30.7%). OFI pectin had a WHC of 5.42 ± 0.16 g water /g OFI pectin, slightly lower than that for pectin extracted by the chemical process (5.64 g water/g OFI pectin) [34]. Various intrinsic factors, like the chemical structure of the biomaterial, and extrinsic factors, such as the pH, temperature, and ionic strength, can affect the WHC [45]. The OHC value of pectins was 1.23 ± 0.42 g oil/g OFI pectin. It was like the OHC of the OFI pectin extracted by the acid water method [34]. Thus, the oil retention power depends essentially on the hydrophilic nature and the overall charge density of the constituents [45]. EA values for OFI pectin emulsions at 2 and 4% were 26.9% and 30.77%, respectively. These values were lower than the ones found by Bayar et al. [34] for a 2% concentration of pectin extracted by the chemical process from the OFI cladodes (35%), proving that the extraction process influences the functional properties of pectin macromolecules [46]. The ES rates were 14.31% and 87.48% for 2% and 4% of pectin hydrocolloid in the emulsions, this long-term stability when emulsions were submitted to temperature treatment at 80°C is due to the high viscosity of pectin solution and by the formation of layers around the fat globules by the pectin [39].

Advertisement

5. Unconventional sources of pectin: hulls or husks and seeds

It is well known that the primary sources of pectin extraction are those obtained from citrus fruits or apples, due to their high yield and physicochemical properties that make them useful for various applications in the food and pharmaceutical industries. However, in recent years, new extraction sources have been sought that may represent alternatives to overexploited sources and that also have the advantage of allowing the use of organic by-products, such as the case of hulls or husks and seeds, from which pectins with specific physicochemical properties of high utility for multiple applications can be obtained.

Table 2 shows current research work regarding unconventional sources for obtaining pectins, classified as hulls or husks that come from dry fruits (almonds, pistachios, walnuts, and cocoa), pods, and legume seeds (soy, peas, faba beans, and riang), cereal leaves (Zea mays) and seeds of different fruits (Nicandra physaloides Linn., Gaertn, papaya, jackfruit, creeping fig and sesame). In addition, its extraction methods and its most outstanding properties are also described.

Pectin sourceExtraction conditions and YieldFunctional propertiesReference
Hulls or husks
Almond hullFirst part: acidification with citric acid of almond hull pectin whose optimal conditions were pH = 1.4, liquid-solid ratio (LSR) 20.13, 90°C for 58.65 min followed by filtration.
Second part: a mixture of pectin supernatant with 96% ethanol at a ratio of 1:1 v/v, then the precipitate obtained was dried in an oven at 50°C.
Yield: 26.32% wt.
Extraction of LM pectin
DE: 26.4%
Forming forms gels using Ca2+ at a pH 3–7
Do not need sugar to form gel
High polydispersity due to the small chains formed during the extraction process.
[28]
Pistachio hullConventional and ultrasound-assisted,
Acidification with a citric acid solution
Yield: 32.3%.
Extraction of LM pectin DE = 19.29%
Maximum emulsifying capacity with 6% wt. pectin, EA index: 172.85 ± 0.59 m2/g ES index: 158.28 ± 3.41 min, High creaming stability
Shear-thinning behavior.
[47]
A green husk of walnuts (Juglans regia L.)Walnuts husks from different regions of cultivation.
Walnut husks powder heated in an acid medium.
Ethanol precipitation.
The pectin was decolorized using acetone.
The soil and climate conditions where the walnut husks were obtained caused variations in the properties of the pectins obtained. DE: higher 65%.
Lamellar and leaf-shaped structures, depending on the region of cultivation
[48]
Cocoa pod husk (Theobroma cacao)Microwave-assisted extraction using an acidified medium with oxalic acid.
pH 1.16
15 min.
Liquid/solid ratio:25
Yield: 9.64%.
The decrease in pH during extraction produced a decrease in the esterification degree which reduced the gelling ability of pectin. This could be observed by FTIR spectroscopy.[49]
Legumes (from seeds or pods)
Soy hullThermal treatment by microwave irradiation
Acidification 0.6% wt. citric acid or sodium citrate (SC).
Pectin precipitation with ethanol.
The pectin extracted with SC had better stability of emulsions, smaller droplet sizes and greater emulsifying capacity.
Applied into mayonnaise, achieving uniformly distributed drops and high stability.
[50]
Soy hullPectin extraction was carried out from milled soy hulls, from which galactomannan was removed.
Acid medium, HCl followed by HNO3.
Low uronic acid content.
Low yield.
Xylogalacturonan and rhamnogalacturonans as major components.
Cannot form gels by adding Ca2+.
[51]
Pea hullTo optimize the extraction, a central composite design was carried out where the effect of pH, temperature, and time on the yield and purity of the pectins was evaluated using two different acid media: citric acid and HNO3.
pH 2.0
Yield: 3.5–9.8% with citric acid
Yield: 1.4–8.0% with HNO3.
Purity: >65%, related to the high uric acid content.
LM pectin
Mainly composed of xylogalacturonan.
[52]
Faba bean hullMicrowave assisted extraction (640 W).
pH = 1.5 with 1 M HCl, 9 min.
Ethanol precipitation.
Yield: 14.86%
HM pectin
DE: 54.08%.
[53]
Riang (Parkia timoriana (DC.) Merr.) pod huskAcid water
pH = 2 with HNO3, heated to 90°C for 90 min.
Final pH adjusted to 4.5.
Yield: 15.0%
HM pectin
DE: ~66%.
Pseudoplastic behavior at a concentration > 2%w/v,
Newtonian behavior at concentrations <2% w/v.
High antioxidant activity and high content of phenolic compounds (mainly tannins).
[54]
Cereal
Zea mays huskExtraction with high-power ultrasound (US) application.
Pretreatment in a primary medium, followed by enzymatic hydrolysis with cellulase (pH = 5.2), precipitation of pectin with ethanol, and subsequent lyophilization to obtain the MP fraction.
Second treatment consisted of high-power ultrasound at 20 kHz, plus the steps of the first treatment to obtain the MP-US fraction.
Formation of thermo-irreversible and soft gels at pH = 6 in the presence of Ca2+
Interaction with the iron (II) ion to form thermo-reversible weak gels.
LM pectin and high-water solubility due to the ultrasound treatment.
[55]
Fruit seeds
Papaya seedsAcid medium (citric acid)
80°C, pH 1.5, 60 min
Yield: 8.66%
LM: 9.22%.[56]
Jackfruit seeds sheatsPectin extraction was from the slimy sheats of the jackfruit seed (JS).
Water acidified with oxalic acid
90°C, 1 h.
Yield: 35.52%
Total phenolic content: 65.7 mg GAE/g
Antioxidant activity:
DPPH method: 25.29 ± 4.03%
FRAP: 10.4 μM
[57]
Creeping fig fruit seedsChemical extraction in acid conditions.LM pectin
DE: ~20%.
Form gels at low pH values with the addition of glucono-δ-lactone
Viscous solutions are obtained at pH = 4.5 with Na+ and K+ ions for favoring the formation of the gel.
Gel strength depends on the type of salt added and its concentration.
[58]
Nicandra physaloides (Linn.) Gaertn seedsEnzyme inactivation of the NPG seeds were inactivated with heating.
Aqueous extraction at 60°C. Different fractions of pectin were obtained.
Yield: 9.17–10.56%
LM pectin
DE = 46.93%.
Spontaneous gel formation at 1.5%.
Gel formation at <1.5% in the presence of NaCl and KCl.
[59]
Sesame seed hullDefatted seed
Acid medium (HCl). Fractionation with ethanol (30%, 50%, and 90%).
Maximum yield: 75.6% at 30% ethanol
High antioxidant activity
Able to stabilize emulsions.
[60]

Table 2.

Unconventional sources of pectins: Hulls or husks and seeds.

The most widely used pectin extraction method for hulls or husk and seeds is the conventional one, which consists of acidifying the sample, for which different types of organic (citric and oxalic acid) and inorganic (HCl and HNO3) acids are used; the type of acid used influences the extraction conditions and the properties of the pectin obtained [30]. Subsequently, a heat treatment is carried out using a conventional hot plate, or, for more efficient extraction, it can be assisted by microwaves [49] or by high-power ultrasound [61]. Finally, separation is carried out using the ethanol solvent, and the pectin obtained is dried.

Among the properties shared by pectins obtained from hulls or husks and seeds is that they are primarily LM pectins, with very varied DEs, and they also can form gels in the presence of ions such as calcium, sodium, or potassium [28]. However, the esterification degree influences the properties of the gels formed [49]. Besides, this type of pectin regularly achieves the formation of stable emulsions [24, 62], a shear-thinning rheological behavior and can even, in some cases, present antioxidant activity [14, 54, 63].

One of the main disadvantages of obtaining pectin from hulls or husks and seeds is its low yield (regularly less than 15%), since the extraction is carried out by conventional methods, where conditions such as the type of acid influence the yield obtained. However, research is currently being carried out on new methods that allow a more efficient extraction and higher yield to encourage the use of unconventional sources of pectin.

Advertisement

6. Applications of new sources of pectin

Nowadays, green chemistry leads to environmentally friendly bioproduct extraction approaches. Because bioproducts are biocompatible, they have a wide range of applications [33]. The synthesis and production of bioproducts use substantially less energy and solvent, and they can now be scaled up with a small initial expenditure [14, 40]. Biomaterial formulations, sometimes inspired by biomimicking nature’s behavior, are specifically tailored for applications involving human consumption products and innovative biobased materials [64]. In the field of biomaterials, hydrogels have gained popularity owing to their specific properties, such as biodegradability, biocompatibility, a soft-wet feel, and resemblance to organic tissue. Hydrogels with tridimensional crosslinked polymeric structures made from natural polymers have been extensively studied because of the increasing need for biomaterials with novel features for human consumption-related applications [65]. Pectin, a biopolymer found in the cell walls of fruits and vegetables, is extensively employed in the food, pharmaceutical, and textile sectors due to its ability to produce a thick gel-like solution [65]. Pectin is a gelling ingredient in the production of jams, jellies, and marmalades. Over the past decade, intense new research has yielded a new understanding of its molecular structure and physiological function, opening the gate to novel manufacturing techniques and entirely new applications, such as new advanced biomaterials, for example, calcium phosphate pectin for bone restoration and biobased construction, and building materials, for example, pectin aerogels for thermal insulation [64, 66].

According to the scientific literature, we can classify applications of pectins in two ways: first, according to their physicochemical properties, and last, according to their field of application. The specific application of each of the novel pectin sources is intimately linked to their particular physicochemical characteristics; please see Figure 3. For example, LM pectin is believed to be a helpful stabilizer for dairy products. This is due to low methoxyl pectin gels in the presence of divalent cations, in this specific instance, calcium ions. The capacity of HM pectin to gel at moderately lower pH values (pH 2–3.5) in the addition of soluble substances, such as sucrose, makes it suitable for use in the preparation of jams and jellies [67]. An LM pectin is an attractive option for use as a gelling agent in manufacturing low-calorie jams due to its ability to form a gel without added sugar. Unlike gums, which impart a slimy mouth feel, the use of pectin to increase the viscosity in soft drinks and beverages gives a clean mouth feel; this may be due to the low viscosity of low-concentration pectin solutions at the shear rate of the mouth [67].

Figure 3.

Different classifications for pectin.

From the viewpoint related to their field of application, pectins may be categorized into pectins for food products and pectins for drug and therapeutic applications. Please see Table 3. Current research trends in food packaging promote the development of biodegradable, renewable, and environmentally friendly materials. Pectin-based edible coatings are among the most recent advancements in the world of food packaging. Including additional biopolymers, such as cellulose and natural compounds with antioxidant and antibacterial properties, has enhanced and strengthened these coatings.

ApplicationSourceExtraction techniqueReference
Pectins for food productsStabilizer for dairy productsGrapefruit peelAcid hydrolysis[68]
Food films, gelling agents, and plasticizerLime peelCitric acid-microwave extraction[69]
Food packagingLemon waste peelMicrowave extraction[70]
Films and emulsionsCitrus[71]
Antioxidants in food formulationsJackfruit peelUltrasonic-microwave extraction[72]
Pectins for drug and therapeutic applicationsDrug delivery systemsCitrus[73]
Drug delivery systemsFig skinUltrasonic-microwave extraction[74]
Tissue engineeringLemon peelsAcid hydrolysis[75]
Bioprinting of 3D scaffoldsCitrus peels[76]
Wound healingAkebia trifoliata fruit peelAcid hydrolysis[77]
Accelerated wound healingCyclea Barbata MiersCold acid hydrolysis[78]
Skin wound healingPapaya fruit[79]

Table 3.

Applications of novel pectins.

Additionally, researchers have discovered the biological functions of pectin, consequently increasing its application in the pharmaceutical industry, including drug delivery systems, skin and bone tissue engineering, and wound dressings [65, 76]. Pectin is most widely used in the formulation of drugs for oral administration, such as tablets, gels, hydrogels, beads, aerogels, and coated and compression-coated doses. The ability of pectin to withstand acidic conditions and higher temperatures allows for the development of drug delivery systems able to load and release drugs at a specific location. Pectin has primarily been considered a colon-specific drug delivery vehicle that reduces systemic toxicity while increasing bioactivity and medication stability.

Pectin also has significant potential for use in tissue engineering. Pectins may promote mineral nucleation in this application if immersed in the appropriate physiological conditions, resulting in biomimetic structures that more closely resemble the natural architecture of bone. Furthermore, pectins are responsible for wound healing treatments’ gelling protection and anti-inflammatory effects [76]. By crosslinking pectins, calcium ions aid in its gelation. Solubilized pectin forms an acidic environment that acts as a bacterial or viral barrier, and pectin hydrogels allow for the loading and release of drugs such as antibiotics, analgesics, and tissue repair agents. Other physiological effects of pectin have been described, such as prebiotic, antimicrobial, antiglycation, and antioxidant. Pectin has also been used to nano-encapsulate bioactive substances, thereby increasing their shelf life and stability.

The exploration of new sources of pectin, involving the introduction of cleaner and new sustainable extraction techniques, demands more research to guarantee that an industrial application is sustainable and competitive in the current market.

Advertisement

7. Conclusions

Pectin is one of the primary polysaccharides present in plants; it contributes to the physical and nutritional value of plant-based goods. It’s a macromolecule that can create flexible polymer chains. Source and extraction circumstances affect its functioning characteristics. Citric fruits and apples are the principal sources of commercial pectin, although non-conventional sources have been examined, including agro-industrial sub-products and wastes, pulps, husks, hulls, peels, Cactaceae, and vegetables. Pectin has been functionalized by chemical or enzyme processes that affect its physical characteristics, such as molecular weight, degree of esterification (DE), and surface charge, leading to new functional or enhanced qualities as well as new techniques and applications. Pumpkin, eggplant, chayote, and Opuntia ficus indica cladodes are new sources of pectin. Due to their high production and physicochemical qualities, citrus fruits and apples are the principal sources of pectin extraction. In recent years, new extraction sources have been sought that may represent alternatives to overexploited sources and that allow the use of organic by-products, such as hulls or husks and seeds, from which pectin with specific physicochemical properties can be obtained for multiple applications. Intense new research has yielded a new understanding of its molecular structure and physiological function, opening the door to novel manufacturing techniques and entirely new applications, such as calcium phosphate pectin for bone restoration and pectin aerogels for thermal insulation.

Advertisement

Acknowledgments

The authors wish to acknowledge the partial financial support of this research to the Universidad Autónoma del Estado de México through project 6661/2022SF.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. 2023 Pectin Market Outlook Report - Market Size, Market Split, Market Shares Data, Insights, Trends, Opportunities, Companies: Growth Forecasts by Product Type, Application, and Region from 2022 to 2030. https://www.researchandmarkets.com/reports/5686321/2023-pectin-market-outlook-report-market-size#product--description [Accessed: November 21, 2022)
  2. 2. Freitas CMP, Coimbra JSR, Souza VGL, et al. Structure and applications of pectin in food, biomedical, and pharmaceutical industry: A review. Coatings. 2021;11:922. DOI: 10.3390/COATINGS11080922
  3. 3. Herrera-Rodríguez SE, Pacheco N, Ayora-Talavera T, et al. Advances in the green extraction methods and pharmaceutical applications of bioactive pectins from unconventional sources: A review. Studies in Natural Products Chemistry. 2022;73:221-264. DOI: 10.1016/B978-0-323-91097-2.00015-7
  4. 4. Muñoz-Almagro N, Montilla A, Moreno FJ, et al. Modification of citrus and apple pectin by power ultrasound: Effects of acid and enzymatic treatment. Ultrasonics Sonochemistry. 2017;38:807-819. DOI: 10.1016/J.ULTSONCH.2016.11.039
  5. 5. Muñoz-Almagro N, Montilla A, Villamiel M. Role of pectin in the current trends towards low-glycaemic food consumption. Food Research International (Ottawa, Ont). 2021;140. Epub ahead of print 1 February. DOI: 10.1016/J.FOODRES.2020.109851
  6. 6. Zhang H, Chen J, Li J, et al. Extraction and characterization of RG-I enriched pectic polysaccharides from mandarin citrus peel. Food Hydrocolloids. 2018;79:579-586. DOI: 10.1016/j.foodhyd.2017.12.002
  7. 7. Willats WGT, Knox JP, Mikkelsen JD. Pectin: New insights into an old polymer are starting to gel. Trends in Food Science & Technology. 2006;17:97-104. DOI: 10.1016/J.TIFS.2005.10.008
  8. 8. Leclere L, Van CP, Michiels C. Anti-cancer activities of pH- or heat-modified pectin. Frontiers in Pharmacology. Oct 2013;4:128. DOI: 10.3389/FPHAR.2013.00128/BIBTEX
  9. 9. Atmodjo MA, Hao Z, Mohnen D. Evolving views of pectin biosynthesis. Annual Review of Plant Biology. 2013;64:747-779. DOI: 10.1146/annurev-arplant-042811-105534
  10. 10. Ndeh D, Rogowski A, Cartmell A, et al. Complex pectin metabolism by gut bacteria reveals novel catalytic functions. Nature. 2017;544(7648):65-70. DOI: 10.1038/nature21725
  11. 11. O'Neill MA, Ishii T, Albersheim P, et al. Rhamnogalacturonan II: Structure and function of a borate cross-linked cell wall pectic polysaccharide. Annual Review of Plant Biology. 2004;55:109-139. DOI: 10.1146/ANNUREV.ARPLANT.55.031903.141750
  12. 12. Dranca F, Oroian M. Extraction, purification and characterization of pectin from alternative sources with potential technological applications. Food Research International. 2018;113:327-350. DOI: 10.1016/j.foodres.2018.06.065
  13. 13. Ngenefeme FTJ, Eko NJ, Mbom YD, et al. A one pot green synthesis and characterisation of iron oxide-pectin hybrid nanocomposite. Open Journal of Composite Materials. 2013;03:30-37. DOI: 10.4236/ojcm.2013.32005
  14. 14. Alpizar-Reyes E, Cortés-Camargo S, Román-Guerrero A, et al. Chapter 14: Tamarind gum as a wall material in the microencapsulation of drugs and natural products. In: Micro- and Nanoengineered Gum-Based Biomaterials for Drug Delivery and Biomedical Applications. 2022. pp. 347-382. DOI: 10.1016/ B978-0-323-90986-0.00016-9
  15. 15. Chen J, Niu X, Dai T, et al. Amino acid-amidated pectin: Preparation and characterization. Food Chemistry. 2020;309. Epub ahead of print 30 March. DOI: 10.1016/J.FOODCHEM.2019.125768
  16. 16. Zheng C, Zou Y, Huang Y, et al. Biosynthesis of amidated pectins with ultra-high viscosity and low gelation restriction through ultra-low temperature enzymatic method. Food Hydrocolloids. 2023;134. Epub ahead of print 1 January. DOI: 10.1016/J.FOODHYD.2022.108037
  17. 17. Chandel V, Biswas D, Roy S, et al. Current advancements in pectin: Extraction. Properties and Multifunctional Applications. Foods. 2022;11:1-30. DOI: 10.3390/foods11172683
  18. 18. Mesbahi G, Jamalian J, Farahnaky A. A comparative study on functional properties of beet and citrus pectins in food systems. Food Hydrocolloids. 2005;19:731-738. DOI: 10.1016/J.FOODHYD.2004.08.002
  19. 19. Ciriminna R, Chavarría-Hernández N, Inés Rodríguez Hernández A, et al. Pectin: A new perspective from the biorefinery standpoint. Biofuels, Bioproducts and Biorefining. 2015;9:368-377. DOI: 10.1002/BBB.1551
  20. 20. Chen J, Cheng H, Zhi Z, et al. Extraction temperature is a decisive factor for the properties of pectin. Food Hydrocolloids. 2021;112:106160. DOI: 10.1016/J.FOODHYD.2020.106160
  21. 21. Ciriminna R, Fidalgo A, Delisi R, et al. Controlling the degree of esterification of citrus pectin for demanding applications by selection of the source. ACS Omega. 2017;2:7991-7995. DOI: 10.1021/ACSOMEGA.7B01109/ASSET/IMAGES/AO-2017-01109E_M002.GIF
  22. 22. Panouillé M, Thibault JF, Bonnin E. Cellulase and protease preparations can extract Pectins from various plant byproducts. Journal of Agricultural and Food Chemistry. 2006;54:8926-8935. DOI: 10.1021/JF0617824
  23. 23. Munarin F, Tanzi MC, Petrini P. Advances in biomedical applications of pectin gels. International journal of biological macromolecules. 2012;51:681-689. DOI: 10.1016/J.IJBIOMAC.2012.07.002
  24. 24. Alpizar-Reyes E, Cruz-Olivares J, Cortés-Camargo S, et al. Structural, physicochemical, and emulsifying properties of pectin obtained by aqueous extraction from red pitaya (Hylocereus polyrhizus) peel. Revista Mexicana de Ingeniera Quimica. 2022;21:Alim2887–Alim2887. DOI: 10.24275/rmiq/Alim2887
  25. 25. Smith DA. Jams and preserves. Methods of manufacture. In: Caballero B, editor. Encyclopedia of Food Sciences and Nutrition. 2nd ed. USA: Elsevier; 2003. pp. 3409-3415. DOI: 10.1016/B0-12-227055-X/00660-X
  26. 26. Cui SW, Chang YH. Emulsifying and structural properties of pectin enzymatically extracted from pumpkin. LWT - Food Science and Technology. 2014;58:396-403. DOI: 10.1016/J.LWT.2014.04.012
  27. 27. Vincken JP, Schols HA, Oomen RJFJ, et al. If Homogalacturonan were a side chain of Rhamnogalacturonan I. implications for Cell Wall architecture. Plant Physiology. 2003;132:1781-1789. DOI: 10.1104/PP.103.022350
  28. 28. Hosseini SS, Khodaiyan F, Kazemi M, et al. Optimization and characterization of pectin extracted from sour orange peel by ultrasound assisted method. International Journal of Biological Macromolecules. 2019;125:621-629. DOI: 10.1016/J.IJBIOMAC.2018.12.096
  29. 29. Kazemi M, Khodaiyan F, Labbafi M, et al. Pistachio green hull pectin: Optimization of microwave-assisted extraction and evaluation of its physicochemical, structural and functional properties. Food Chemistry. 2019;271:663-672. DOI: 10.1016/j.foodchem.2018.07.212
  30. 30. Feliciano GD, Lima EAC, dos Pereira MJS, et al. Effect of a chayotte (Sechium edule) extract on the labeling of red blood cells and plasma proteins with technetium-99m: in vitro and in vivo studies. Cellular and Molecular Biology (Noisy-le-Grand, France). 2002;48:751-755
  31. 31. Ke J, Jiang G, Shen G, et al. Optimization, characterization and rheological behavior study of pectin extracted from chayote (Sechium edule) using ultrasound assisted method. International Journal of Biological Macromolecules. 2020;147:688-698. DOI: 10.1016/j.ijbiomac.2020.01.055
  32. 32. Su DL, Li PJ, Quek SY, et al. Efficient extraction and characterization of pectin from orange peel by a combined surfactant and microwave assisted process. Food Chemistry. 2019;286:1-7. DOI: 10.1016/J.FOODCHEM.2019.01.200
  33. 33. Velázquez-Gutiérrez SK, Alpizar-Reyes E, Cruz-Olivares J, et al. Ionic gelation encapsulation of sesame oil with sodium alginate-nopal mucilage blends: Encapsulation efficiency and oxidative stability. Revista Mexicana de Ingenieria Quimica. 2020;19:Alim349– Alim362. DOI: 10.24275/rmiq/Alim1642
  34. 34. Bayar N, Kriaa M, Kammoun R. Extraction and characterization of three polysaccharides extracted from Opuntia ficus indica cladodes. International Journal of Biological Macromolecules. 2016;92:441-450. DOI: 10.1016/J.IJBIOMAC.2016.07.042
  35. 35. Bayar N, Friji M, Kammoun R. Optimization of enzymatic extraction of pectin from Opuntia ficus indica cladodes after mucilage removal. Food Chemistry. 2018;241:127-134. DOI: 10.1016/j.foodchem.2017.08.051
  36. 36. Bayar N, Bouallegue T, Achour M, et al. Ultrasonic extraction of pectin from Opuntia ficus indica cladodes after mucilage removal: Optimization of experimental conditions and evaluation of chemical and functional properties. Food Chemistry. 2017;235:275-282. DOI: 10.1016/j.foodchem.2017.05.029
  37. 37. Forni E, Penci M, Polesello A. A preliminary characterization of some pectins from quince fruit (Cydonia oblonga mill.) and prickly pear (Opuntia ficus indica) peel. Carbohydrate Polymers. 1994;23:231-234. DOI: 10.1016/0144-8617(94)90183-X
  38. 38. Chemat F, Rombaut N, Sicaire AG, et al. Ultrasound assisted extraction of food and natural products. Mechanisms, techniques, combinations, protocols and applications. A review. Ultrasonics Sonochemistry. 2017;34:540-560. DOI: 10.1016/J.ULTSONCH.2016.06.035
  39. 39. Prajapati VD, Jani GK, Moradiya NG, et al. Pharmaceutical applications of various natural gums, mucilages and their modified forms. Carbohydrate Polymers. 2013;92:1685-1699. DOI: 10.1016/j.carbpol.2012.11.021
  40. 40. Alpizar-Reyes E, Román-Guerrero A, Cortés-Camargo S, et al. Chapter 2: Recent approaches in alginate-based carriers for delivery of therapeutics and biomedicine. In: Polysaccharide-based Biomaterials: Delivery of Therapeutics and Biomedical Applications. 2022. pp. 27-68. DOI: 10.1039/9781839166235-00027
  41. 41. Lv C, Wang Y, Wang LJ, et al. Optimization of production yield and functional properties of pectin extracted from sugar beet pulp. Carbohydrate Polymers. 2013;95:233-240. DOI: 10.1016/j.carbpol.2013.02.062
  42. 42. Funami T, Nakauma M, Ishihara S, et al. Structural modifications of sugar beet pectin and the relationship of structure to functionality. Food Hydrocolloids. 2011;25:221-229. DOI: 10.1016/J.FOODHYD.2009.11.017
  43. 43. Zykwinska A, Thibault JF, Ralet MC. Competitive binding of pectin and xyloglucan with primary cell wall cellulose. Carbohydrate Polymers. 2008;74:957-961. DOI: 10.1016/J.CARBPOL.2008.05.004
  44. 44. Lira-Ortiz AL, Reséndiz-Vega F, Ríos-Leal E, et al. Pectins from waste of prickly pear fruits (Opuntia albicarpa Scheinvar' Reyna'): Chemical and rheological properties. Food Hydrocolloids. 2014;37:93-99. DOI: 10.1016/j.foodhyd.2013.10.018
  45. 45. Elleuch M, Bedigian D, Roiseux O, et al. Dietary fibre and fibre-rich by-products of food processing: Characterisation, technological functionality and commercial applications: A review. Food Chemistry. 2011;124:411-421. DOI: 10.1016/J.FOODCHEM.2010.06.077
  46. 46. Leroux J, Langendorff V, Schick G, et al. Emulsion stabilizing properties of pectin. Food Hydrocolloids. 2003;17:455-462. DOI: 10.1016/S0268-005X(03)00027-4
  47. 47. Barış S, Elik A, Fahrettin G, et al. Pistachio hull as an alternative pectin source: Its extraction and use in oil in water emulsion system. Preparative Biochemistry & Biotechnology. 2022. DOI: 10.1080/10826068.2022.2098323 [Epub ahead of print]
  48. 48. La Torre C, Caputo P, Plastina P, et al. Green husk of walnuts (Juglans regia L.) from southern Italy as a valuable source for the recovery of glucans and Pectins. Fermentation. 2021;7:305. DOI: 10.3390/FERMENTATION7040305
  49. 49. Pangestu R, Amanah S, Juanssilfero AB, et al. Response surface methodology for microwave-assisted extraction of pectin from cocoa pod husk (Theobroma cacao) mediated by oxalic acid. Journal of Food Measurement and Characterization. 2020;14:2126-2133. DOI: 10.1007/s11694-020-00459-4
  50. 50. Liu D, Tang W, Huang XJ, et al. Structural characteristic of pectin-glucuronoxylan complex from Dolichos lablab L. hull. Carbohydrate Polymers. 2022;298:120023. DOI: 10.1016/J.CARBPOL.2022.120023
  51. 51. Reichembach LH, de Oliveira Petkowicz CL. New findings on acid-extractable pectins from soy hull. Carbohydrate Polymers. 2022;294:119831. DOI: 10.1016/J.CARBPOL.2022.119831
  52. 52. Gutöhrlein F, Drusch S, Schalow S. Extraction of low methoxylated pectin from pea hulls via RSM. Food Hydrocolloids. 2020;102:105609. DOI: 10.1016/J.FOODHYD.2019.105609
  53. 53. Hussain Dar A, Ul Q , Rizvi EH, et al. Microwave Assisted Extraction of Pectin from Dried Hull of Faba Bean. www.jpht.in [Accessed: November 21, 2022]
  54. 54. Buathongjan C, Israkarn K, Sangwan W, et al. Studies on chemical composition, rheological and antioxidant properties of pectin isolated from Riang (Parkia timoriana (DC.) Merr.) pod. International Journal of Biological Macromolecules. 2020;164:4575-4582. DOI: 10.1016/J.IJBIOMAC.2020.09.079
  55. 55. Higuera-Coelho RA, Lizarraga L, Ponce NMA, et al. Novel gelling pectins from Zea mays husks' agro-industrial residue and their interaction with calcium and iron (II). Bioactive Carbohydrates and Dietary Fibre. 2021;26:100273. DOI: 10.1016/J.BCDF.2021.100273
  56. 56. Madhuvanthi S, Selvapriya K, Nirmala RA, et al. Extraction and characterization of pectin derived from underutilized papaya seeds as a value-added product. Journal of Applied and Natural Science. 2022;14:127-132. DOI: 10.31018/JANS.V14I1.3269
  57. 57. Kumar M, Potkule J, Tomar M, et al. Jackfruit seed slimy sheath, a novel source of pectin: Studies on antioxidant activity, functional group, and structural morphology. Carbohydrate Polymer Technologies and Applications. 2021;2:100054. DOI: 10.1016/J.CARPTA.2021.100054
  58. 58. Chen R, Ratcliffe I, Williams PA, et al. The influence of pH and monovalent ions on the gelation of pectin from the fruit seeds of the creeping fig plant. Food Hydrocolloids. 2021;111:106219. DOI: 10.1016/J.FOODHYD.2020.106219
  59. 59. Liu H, Deng L, Dai T, et al. Emulsifying and emulsion stabilization mechanism of pectin from Nicandra physaloides (Linn.) Gaertn seeds: Comparison with apple and citrus pectin. Food Hydrocolloids. 2022;130:107674. DOI: 10.1016/J.FOODHYD.2022.107674
  60. 60. Liu HM, He MK, Yao YG, et al. Pectic polysaccharides extracted from sesame seed hull: Physicochemical and functional properties. International Journal of Biological Macromolecules. 2021;192:1075-1083. DOI: 10.1016/J.IJBIOMAC.2021.10.077
  61. 61. Selkälä T, Suopajärvi T, Sirviö JA, et al. Surface modification of cured inorganic foams with cationic cellulose nanocrystals and their use as reactive filter Media for Anionic dye Removal. ACS Applied Materials & Interfaces. 2020;12:27745-27757. DOI: 10.1021/acsami.0c05927
  62. 62. Duan X, Zhu Y, Shu C, et al. Extraction of pectin from Satsuma mandarin Peel: A comparison of high hydrostatic pressure and conventional extractions in different acids. Molecules. 2022;27:3747. DOI: 10.3390/MOLECULES27123747
  63. 63. Kumar M, Tomar M, Saurabh V, et al. Delineating the inherent functional descriptors and biofunctionalities of pectic polysaccharides. Carbohydrate Polymers. 2021;269:118319. DOI: 10.1016/J.CARBPOL.2021.118319
  64. 64. Alpizar-Reyes E, Concha JL, Martín-Martínez FJ, et al. Biobased spore microcapsules for asphalt self-healing. ACS Applied Materials & Interfaces. 2022;14:31296-31311. DOI: 10.1021/acsami.2c07301
  65. 65. Eivazzadeh-Keihan R, Noruzi EB, Aliabadi HAM, et al. Recent advances on biomedical applications of pectin-containing biomaterials. International Journal of Biological Macromolecules. 2022;217:1-18. DOI: 10.1016/J.IJBIOMAC.2022.07.016
  66. 66. Ciriminna R, Fidalgo A, Scurria A, et al. Pectin: New science and forthcoming applications of the most valued hydrocolloid. Food Hydrocolloids. 2022;127:107483. DOI: 10.1016/J.FOODHYD.2022.107483
  67. 67. Picot-Allain MCN, Ramasawmy B, Emmambux MN. Extraction, characterisation, and application of pectin from tropical and sub-tropical fruits: A review. Food Reviews International. 2020;38:282-312. DOI: 10.1080/87559129.2020.1733008
  68. 68. Khan AA, Butt MS, Randhawa MA, et al. Extraction and characterization of pectin from grapefruit (Duncan cultivar) and its utilization as gelling agent. International Food Research Journal. 2014;21:2195-2199
  69. 69. Rodsamran P, Sothornvit R. Microwave heating extraction of pectin from lime peel: Characterization and properties compared with the conventional heating method. Food Chemistry. 2019;278:364-372. DOI: 10.1016/J.FOODCHEM.2018.11.067
  70. 70. Dash KK, Ali NA, Das D, et al. Thorough evaluation of sweet potato starch and lemon-waste pectin based-edible films with nano-titania inclusions for food packaging applications. International Journal of Biological Macromolecules. 2019;139:449-458. DOI: 10.1016/J.IJBIOMAC.2019.07.193
  71. 71. Almasi H, Azizi S, Amjadi S. Development and characterization of pectin films activated by nanoemulsion and Pickering emulsion stabilized marjoram (Origanum majorana L.) essential oil. Food Hydrocolloids. 2020;99:105338. DOI: 10.1016/J.FOODHYD.2019.105338
  72. 72. Xu SY, Liu JP, Huang X, et al. Ultrasonic-microwave assisted extraction, characterization and biological activity of pectin from jackfruit peel. LWT. 2018;90:577-582. DOI: 10.1016/J.LWT.2018.01.007
  73. 73. Kaushik P, Priyadarshini E, Rawat K, et al. pH responsive doxorubucin loaded zein nanoparticle crosslinked pectin hydrogel as effective site-specific anticancer substrates. International Journal of Biological Macromolecules. 2020;152:1027-1037. DOI: 10.1016/J.IJBIOMAC.2019.10.190
  74. 74. Jannasari N, Fathi M, Moshtaghian SJ, et al. Microencapsulation of vitamin D using gelatin and cress seed mucilage: Production, characterization and in vivo study. International Journal of Biological Macromolecules. 2019;129:972-979. DOI: 10.1016/j.ijbiomac.2019.02.096
  75. 75. Demir D, Ceylan S, Göktürk D, et al. Extraction of pectin from albedo of lemon peels for preparation of tissue engineering scaffolds. Polymer Bulletin. 2021;78:2211-2226. DOI: 10.1007/S00289-020-03208-1/FIGURES/8
  76. 76. Lapomarda A, De Acutis A, Chiesa I, et al. Pectin-GPTMS-based biomaterial: Toward a sustainable bioprinting of 3D scaffolds for tissue engineering application. Biomacromolecules. 2020;21:319-327. DOI: 10.1021/ACS.BIOMAC.9B01332/ASSET/IMAGES/LARGE/BM9B01332_0010.JPEG
  77. 77. Cai T, Xiao P, Yu N, et al. A novel pectin from Akebia trifoliata var. australis fruit peel and its use as a wall-material to coat curcumin-loaded zein nanoparticle. International Journal of Biological Macromolecules. 2020;152:40-49. DOI: 10.1016/j.ijbiomac.2020.02.234
  78. 78. Sari M, Tamrin KJ, et al. A novel composite membrane pectin from Cyclea Barbata Miers blend with chitosan for accelerated wound healing. Polymer Testing. 2021;99:107207. DOI: 10.1016/J.POLYMERTESTING.2021.107207
  79. 79. Jáuregui KMG, Cabrera JCC, Ceniceros EPS, et al. A New formulated stable papin-pectin aerosol spray for skin wound healing. Biotechnology and Bioprocess Engineering. 2009;14:(4):450-456. DOI: 10.1007/S12257-008-0268-0

Written By

Stefani Cortés-Camargo, Angélica Román-Guerrero, Erik Alpizar-Reyes and César Pérez-Alonso

Submitted: 23 November 2022 Reviewed: 17 December 2022 Published: 13 January 2023