Open access peer-reviewed chapter

Pectin - Extraction, Purification, Characterization and Applications

Written By

Virginia Rodríguez Robledo* and Lucía Isabel Castro Vázquez

Submitted: 15 February 2019 Reviewed: 03 March 2019 Published: 03 December 2019

DOI: 10.5772/intechopen.85588

From the Edited Volume

Pectins - Extraction, Purification, Characterization and Applications

Edited by Martin Masuelli

Chapter metrics overview

2,689 Chapter Downloads

View Full Metrics

Abstract

Fruits, vegetables, and other plant-based foods are particularly important as they are source of dietary carbohydrates, and therefore much of the energy in the adult diet. Plant food also contains a wide range of dietary components rich in bioactive phytochemicals and is essential to the human body that may provide desirable health benefits beyond basic nutrition. Pectin is one of the nonstarch polysaccharides (NSPs), which constitutes the major fraction of the plant cell wall in association and/or substituted with other polysaccharides, and they cover a great variety of biological functions and chemical structures. Generally, pectin is isolated from by-products of agro-foods using extraction technologies with the emergence of novel and effective techniques that inclined toward a cleaner process. Pectin is widely used both in food sector (as gelling, thickening, and stabilizer agent) and in pharmaceutical industries (bioactive components) including biomedical application (drug delivery, tissue engineering, and wound healing) as innovative applications.

Keywords

  • new sources
  • pectin isolations
  • innovative application
  • food sector
  • tissue engineering
  • drug delivery

1. Introduction

Pectins represent a group of structurally heteropolysaccharides, composed mainly by covalently α-1,4-linked D-galacturonic acid (GalA) units, found in primary cell walls and middle lamella of higher plants [1, 2]. Pectin contributes to the firmness and structure of plant tissue, being involved in the intercellular adhesion and mechanical resistance of the cell wall. They also have an important role in the development of plant cells providing turgidity and resistance [3] Figure 1. These polysaccharides have been used in the food and beverage industries for many years. The principal applications of pectin are as a gelling agent, stabilizer, emulsifier, and thickening agent [4, 5, 6].

Figure 1.

Structure of primary plant cell wall (Copyright Figure 1, [56]).

In the food sector, this traditional usage is being complemented by the emerging utilization of pectin as a fat replacer and health-promoting functional ingredient [4, 7, 8]. Pectin also provides an important source of dietary fiber that has been identified as emerging prebiotic with improved therapeutic properties for gut microbiota modulation [9, 10, 11].

Pectin has also been used for medicine and pharmaceutical purposes as a carrier for controlled drugs or bioactive release [12], for example, in drug delivery [13, 14].

Advertisement

2. Pectin chemical structure

Nowadays, there are a consensual classification of pectin based on three main structural domains: homogalacturonan (HGA), alternating with two types of highly branched rhamnogalacturonan regions designated as RG-I and RG-II [2]. Other structural classes of pectic polysaccharides also comprise xylogalacturonan, apiogalacturonan, arabinan, galactan, and arabinogalactan I [15, 16]. The pectin structure largely governs its physicochemical properties and its applications for several purposes. In this context, it is important to note that the carboxylic groups or hydroxyls may be methyl-esterified and/or O-acetyl-esterified [16]. The O-acetyl-esterification occurs predominantly at the O-3 position and occasionally at the O-2 position. However, not only these chemical variables are important but also the composition of neutral sugars, the degree of branching, and the degree of polymerization (molar weight), which can also modify the structure of this complex polysaccharide [17, 18]. The pectin gelation process is strongly influenced by the pectin esterification degree (DM). So, depending on the DM of pectin (defined as percentage of carboxyl groups esterified) [19], two different procedures of pectin gelation can be distinguished. High-methoxylated pectins (DM > 50%) are pectin formed in the presence of cosolutes as sucrose, at a concentration > 55%, and under pH < 3.5 [20]. Low-methoxylated pectins (DM < 50%) are gels created with the presence of divalent cations, mainly Ca2+. Gelation process is due to the formation of junction zones between HGA regions of different pectin chains through calcium bridges between dissociated carboxyl groups [21].

High and low methoxyl pectins have diverse physicochemical properties and therefore have varied applications. High methoxyl pectin can be used as a gelling agent, emulsifier, stabilizer, and thickener in the food industry for the production of jams and jellies. Low methoxyl pectin can be used like a fat replacer, ice cream, yoghurt, bakery glazing, emulsified meat products, and low calorie products [22]. The diversity of pectin structures influences the physicochemical properties and also in its different technological applications, biological activities, biofunctionality, and therapeutic properties.

2.1 Homogalacturonan (HGA)

HGA is a linear homopolymer of α-1,4-linked GalA and is known as the “smooth region.” It is an abundant and widespread domain of pectin, accounting for approximately 60–65% of total pectin amount [23]. The carboxyl groups present at C-6 in the GalA units can be partially methyl-esterified. Around 70–80% of GalA units, methyl-esterified could also be O-acetyl-esterified at O-2 or O-3 positions, depending on its provenance [2]. The amount of GalA units present in a HGA chain is estimated to be around 100–200 units [24]. The smooth region can sometimes be joined by one or two α-1,2-linked L-rhamnopyranose units and most of the pectins have this structure. In addition, GalA units may be substituted at the C-2 or C-3 positions with residues of xylose or apiose, producing domains known as xylogalacturonan or apiogalacturonan, respectively [25].

Regarding other structural classes of pectin polysaccharides, xylogalacturonans (XGAs) are HGA substituted with β-linked-D-xylose-(1-3) at O-3 single unit side chains [26]. The degree of xylosidation can vary between 25 and 75% for instance in watermelon and apple, respectively [27]. Part of the GalA residues in XGA is methyl-esterified independently of the xylose substitutions [28].

2.2 Rhamnogalacturonan I (RG-I)

RG-I comprises a backbone of repeating disaccharide units of [-α-L-Rhap-1,4-α-D-GalpA-1,2-]n [29, 30]. The RG-I backbone may enclose up to 300 rhamnosyl and 300 galactosyluronic acid residues. The composition of neutral sugar side chains can be a glycosyl residue up to 50, resulting in a large and highly variable family of polysaccharides with a range of glycosidic linkages [23]. The highly branched nature of RG-I has led to the name “hairy region.” The GalA residues of RG-I could be O-acetyl-esterified but they are not methyl-esterified [31]. In most cases, rhamnose residues are preferably substituted at the C-4 position with neutral sugar side chains, like arabinose and galactose forming arabinan, galactan, and arabinogalactans [2, 32].

Regarding to arabinogalactans, RG-I with galactose and arabinose side chains possess two structural different forms designed as arabinogalactans I (AGI) and arabinogalactans II (AGII). The AGI consist in a linear chain of 1,4-linked β-D-galactose, containing up to 25% α-L-arabinose residues 1,5-linked in short side chains, connected predominantly to O-4 of the rhamnosyl residues [1]. This type of arabinogalactan has been widely spread in citrus, apple, potatoes, soybean, lupin, cabbage, onion, tomato, and kiwi. It is know that HG and RGI are covalently linked and cannot be separated without a chain-cleavage using, for example, enzymes such as endopolygalacturonase or chromatographic fractionation methods [33]. AGII similarly contains chains of β-D-galactopyranosyl units, but glycosidic linkages occur at C-1, C-3, and C-6 in galactose molecules [34]. This type of arabinogalactan is highly ramified, and it may also include an L-arabinopyranosyl residue at the end of the chain of β-1,6-D-galactopyranosyl units.

2.3 Rhamnogalacturonan-II (RG-II)

RG-II is a homopolymer composed of a backbone that consists of 7–11 galacturonic acid residues branched with 4 side chains, preferably at C-2 and C-3, which may include apiose, 2-O-methyl-L-fucose, 2-O-methyl-D-xylose, 3-C-carboxy-5-deoxy-L-xylose (aceric acid), 3-deoxy-D-manno-octulosonic acid, and 3-deoxy-D-lyxoheptulosaric acid [15, 35]. The side chains of RGII consist of 12 different types of sugars with over 20 different linkages [2]. The RGII, the most structurally complex pectin domain, is highly conserved across many plant sources. In addition, RG-II has the ability to form borate esters dimers [36].

Advertisement

3. Food sources

Nowadays, fruit and some food by-products can be considered as raw materials to produce value-added products enriched with pectin. Pectin is not only a gelling agent but also a thickening, stabilizing, and emulsifying agent [37], and even it has been used as a fat replacer and health-promoting functional ingredient [4, 8]. Thus, pectin must be taken into account within the set of new opportunities for the development of innovative products.

3.1 Conventional sources

Historically, commercial pectin is mostly extracted from citrus peel (85.5%), followed by apple pomace (14.0%) and, to a smaller extent, of the sugar industry (0.5%) [7, 38, 39, 40, 41, 42]. In the cases of apple, citrus, and sugar beet, where the backbone of HG is formed by at least 72–100 D-galacturonic acid residues, the degree of methylation is one of the key parameters related to gelling aptitude. The gelation process is affected by factors including the number and distribution of free carboxyl groups and the molecular weight, but also by other factors as calcium concentration, pH, ionic strength, and temperature [1, 43].

Regarding to citrus fruit, the amount of pectin has been estimated to account for as much as 13.4–29.1% of the dry weight [44]. The yield of pectin recovery from lime, orange, lemon and, lime peels using enzymatic extraction showed no significant effect compared with the extraction yields obtained using acid isolation technique [45].

The second most used source of pectin is apple pulp and apple pomace containing among 4.2–19% of pectin rich in neutral sugars as arabinose, galactose, rhamnose, glucose, and xylose [42]. Apple pectin displays a more elastic-viscous gel; however, citrus pectin produces a more elastic-brittle gel [46, 47].

Sugar beet pulp is also considered as an outstanding low cost source for commercial pectin coming from the sugar industry activity [32]. The combined effect of reduced pH, increased temperature, extraction time, and solvent-solid relation of the extraction process have been related with higher recovery rates of sugar beet pectin ranged between 6.5 and 24.8% [48]. Sugar beet pectin are characterized by showing a diminished gelling properties that have been attributed to its higher acetyl group content, lower molecular weight, higher amount of proteins bound covalently in the lateral chains and greater neutral sugar level [49, 50, 51]. Nevertheless, sugar beet pectin processes remarkable emulsifying properties, greater to pectin extracted from other conventional sources, as consequence of the ferulic acid residues on the arabinan side chains of RG-I, which allow to create covalently cross-linked hydrogels increasing its industrial potential [52, 53].

3.2 Nonconventional sources

Currently, other vegetable wastes constitute a new source of pectin with interesting food applications as valorizing new products and as functional ingredient with health-promoting benefits [54, 55, 56]. Nowadays, there are several unconventional feasible sources of pectin coming from food, vegetable residues, and plant species, which have different pectin contents and physicochemical properties. In this sense, sunflower head residues possess a very interesting gelling properties based on its high molecular weight and high GalA content. This kind of pectin contains around 3.5–5.0% water soluble high-methoxyl pectin and between 12 and 14% of insoluble low-methoxyl pectin [57].

Pectin from olive pomace, a semi-solid by-product from the olive oil industry has low molecular weight and high content of neutral sugars (as beet pectin) and a high percentage of acetylation, which facilitate the formation of gel by ionic interactions with calcium [58, 59]. Among 5–8% of pectin from olive pomace have suitable emulsifying properties for commercial uses and also possess antioxidant and functional properties for its high dietary fiber levels. Carrot pomace and carrot peel, resulting from juice production, are vegetable wastes that possess approximately 22–25% of the total dietary fiber (29.6%) [60]. Pectin from carrot peels displays a very high level of linearity, a lower DM, and a better solubility than pectin extracted from carrot pomace [54] what makes it especially interesting for gelling mechanisms.

Pumpkin pulp and peel showed 7.46% of pectin under extraction at pH 4 with EDTA as chelating agent [61]. Moreover, the pumpkin cake is suggested to have a positive effect on gut bacteria [62]. In the same way, pectin from banana peels comprises around 1/4 of total fruit weight and contains a low amount of water-soluble pectin for gelling, thickening, or stabilizing purposes [63].

Watermelon rinds, which account for approximately one-third of total fruit mass and are usually discarded, were proposed as another possible source of pectin. The pectin content of wet watermelon rinds has displayed 19–21% (w/w) [64] revealing its potential for use in the food industry. Both fresh and lyophilized watermelon pectin have showed high degree of methyl esterification and low molar mass [65]. These properties reveal its suitable application as emulsifier, stabilizing agent, and thickening agents.

Tomato residues, coming from the canning industry, lead large quantities of tomato pomace and tomato peels that contain 7.55 and 22.6% of pectin, respectively, expressed on a dry weight [66]. The structure of pectin molecules isolated from unripe tomato consist in complex biopolymer consists of HGs held together by RG-I regions [67, 68].

Another unconventional pectin sources are summarized in Table 1.

Extraction technique Plant source Operating conditions Pectin content (%) Effect on yield and quality Ref.
Solvent extraction (SE) Cocoa husks Solvent: water, citric acid or hydrochloric acid; pH: 2.5 or 4.0; temperature: 50 or 95°C; time: 1.5 or 3.0 h 3.38–12.60 The highest pectin yield (7.62%) was obtained using citric acid (pH 2.5, 95°C, 3.0 h), while the highest uronic acid content in pectin (65.20%) resulted by using water (95°C, 3.0 h). Extraction with citric acid produced pectin with a wider DM range. [86]
Sour orange peel Solvent: water; liquid/solid ratio: 20:1, 30:1, or 40:1 (v/w); temperature: 75, 85, or 95°C; time: 30, 60, or 90 min 17.95 The yield of pectin extracted at the optimal condition (95°C, 90 min, and liquid/solid ratio of 25:1) was 18.35%. GalA content and DE of the extracted pectin ranged from 57 to 83%, and 17–30.5%, respectively. [87]
Durian rinds Solvent: water; solid/liquid ratio: 1:5–1:15 (g/mL); pH: 2–3; temperature: 75–95°C; time: 20–60 min 9.10 Under optimal conditions (solid/liquid ratio of 1:10 g/mL, pH of 2.8, 43 min, 86°C) [88]
Lime peel Solvent: water: nitric acid; pH: 1.5, 2.3, or 3.1; temperature: 60, 70, or 80°C 15.7 Higher pectin solution concentrations were obtained at the lower pH values. Increased temperature and especially acidity caused a faster decrease of DE, effect that was particularly significant during extractions at pH = 1.5. [89]
Apple pomace Enzymes: endo-xylanase and endo-cellulaseb; solid/liquid ratio: 1 g/15 mL; enzyme dose: 50 U/g; pH: 5.0; temperature: 40°C; time: 10 h 6.85 Treatment with endo-xylanase resulted in the highest pectin yield (19.8%) and very high DM (73.4%). Pectin extracted by endo-cellulase treatment was characterized by the high GalA content (70.5%). Simultaneous use of both enzymatic preparations resulted in a 10.2% extraction yield and a pectin rich in galacturonic acid (74.7%). [90]
Rapeseed cake Commercial enzymes: celluclast and alcalase; enzyme/rapeseed cake ratio: 1:50–1:65 (v/w); celluclast/alcalase ratio: 0:5, 1:4, 2:3, 3:2, 4:1, or 5:0 (v/v); time: 90, 180, 270, 360, or 450 min 6.85 Pectin yield (6.85%) without significant loss of GalA was a 1:50 enzyme/RSC ratio with a celluclast/alcalase ratio of 1:4 for a 270 min hydrolysis time. Enzymes indicated different functions: alcalase led to the destruction of protein-carbohydrate complexes, while celluclast slightly cleaved some linkages of carbohydrate. [91]
Lime peel Enzymes: Laminex C2K, Multifect B, GC220, and GC880; pH: 3.5–6.5; temperature: 40–70°C 23 Laminex C2K preparation proved to the most effective, as in optimum conditions (4 h treatment at pH 3.5, 50°C) gave a high yield (23%) and a pectin with good composition and properties (gelling, stabilization). [92]
Ultrasound-assisted extraction (UAE) Passion fruit peel Extraction solvent: 1.0 mol/L HNO3; pH 2.0, peel/solvent ratio of 1:30 (g/mL); temperature: 45–85°C; power intensity: 132.8–664.0 W/cm2; time and frequency (constant): 10 min, 20 kHz 10.0–30.30 The highest pectin yield (12.67%) was obtained at 85°C and a power intensity of 664 W/cm2. Despite the fact that pectin isolation reached the highest level, the isolate did not display the best composition, as the GalA content and DE showed minimum values. [63]
Mango peel Extraction solvent: citric acid, pH 2.5, peel-solvent ratio of 1:40 (g/mL), time and frequency (constant): 15 min, 20 kHz; temperature: 20 or 85°C 9.20–31.80 Extraction yield of pectins varied greatly with the increase in temperature, from 2.09% (at 20°C) to 17.15% (at 85°C). A significant influence of temperature was also observed for GalA content (increase from 29.35 to 53.35%) and molecular weight (increase from 378.4 to 2320 kDa). [93, 94]
Jack fruit peel Extraction solvent: distilled water); liquid-solid ratio: 10:1–20:1 (mL/g) pH: 1–2; sonication time: 15–30 min; extraction temperature: 50–70°C 8.94–14.5 Optimal conditions for the extraction were: liquid-solid ratio of 15:1 mL/g, pH of 1.6, sonication time of 24 min, and temperature of 60°C. [95]
Grapefruit peel Emitter surface: 13 mm or 25 mm; power density: 0.20–0.53 W/mL; duty cycle: 33–80%; temperature: 30–80°C; solid-liquid ratio: 1/30–1/70 (g/mL); sonication time: 10–60 min 21.5 Heating significantly improved the extractability and extraction rate of pectin, leading to higher yield (26.74%) in shorter extraction time (52 min). The optimized parameters were: ultrasound power density 0.40 W/mL, duty cycle 50%, temperature 60°C, S/L 1/50 g/mL. [8]
Subcritical water extraction (SWE) Apple pomace and citrus peel Solid to liquid ratio of 1:30 g/mL, extraction time of 5 min (constant); extraction temperature: 130, 150, or 170°C for apple pomace; 100, 120, or 140°C for citrus peel 14.0–21.9 The highest yield (21.95%) of citrus pectin (68.88% GalA content) was obtained at 120°C, and the highest yield of apple pectin (16.68%) was gained at 150°C. [96]
Sugar beet pulp Temperature: 110, 120, or 130°C; extraction time: 20, 30, or 40 min; liquid/solid ratio: 30, 40, or 50 (w/w); extraction pressure: 8, 10, or 12 MPa 6.5–24.8 Increase in all parameters enhanced the extraction up to a certain level past which a decrease of pectin recovery was recorded. The optimum extraction conditions to obtain a maximum yield of 24.63% (with 59.12% GalA and 21.66% arabinose content) were as follows: L/S ratio of 44.03, 120.72°C extraction temperature, extraction time of 30.49 min and extraction pressure of 10.70 MPa. [48, 97]
Microwave-assisted extraction (MAE) Waste papaya peel Microwave power: 320, 480, or 640 W pH: 1, 2, or 3; time: 20, 100, or 180 s; solid-liquid ratio: 1:5, 1:15, or 1:25 g/mL 11.11–49.83 All the process variables had significant effect on pectin yield. The optimal conditions for reaching a maximum pectin yield (25.41%) were: microwave power of 512 W, pH of 1.8, time of 140 s, and solid-liquid ratio of 1:15 g/mL. [98, 99]
Waste mango peel Microwave power: 160, 320, or 480 W; pH: 2, 3, or 4; time: 60, 120, or 180 s; solid-liquid ratio: 1:10, 1:20, or 1:30 g/mL 9.20–31.80 For all process parameters, a similar influence was observed: the increase in their level was positively correlated with the pectin yield up to a certain point, beyond which their effect on pectin extraction was negative. The maximum pectin yield (28.86%) could be obtained at a microwave power of 413 W, pH of 2.7, time of 134 s, and solid-liquid ratio of 1:18 g/mL [88]
Tangerine peels Microwave power: 600, 700, or 800 W; temperature: 40, 50, or 60°C; time: 30, 40 or 50 s 19.9 The optimal extraction parameters were: microwave power 704 W, 52.2°C, extraction temperature, and extraction time of 41.8 min [47]
Opuntia ficus-indica cladodes Microwave power: 200, 400, or 600 W pH: 1.5, 2.25, or 3; time: 1, 2, or 3 min; solid-liquid ratio: 2:20, 2:35, or 2:50 g/mL 12.56 The optimum conditions to obtain a maximum pectin recovery of 12.56% were 2.16 min, pH 2.26, 517 W microwave power and 2 g/30.66 mL of solid-liquid ratio. FTIR analysis indicated a GalA content of 34.4% and no alterations in the chemical structure of pectin following microwave treatment. [100]

Table 1.

Main sources of commercial pectin.

Extraction techniques and their effects on pectin yield. Adapted from [56].

Advertisement

4. Pectin isolation process

Since pectin is a native polysaccharide that is found in the cell wall and middle lamellae of many land-growing plants, especially fruits and vegetables, an extraction process for its isolation is mandatory. The innovative applications (food and nonfood) that we will discuss in the following pages begin definitely with pectin isolation from the plant, fruit, or vegetal material.

This entire pectin production process is well documented in the literature and includes a raw material pretreatment stage, an extraction operation, and a postextraction stage (Figure 2). Generally, pectin is isolated by using chemical extraction process from by-products of agro-foods also called “Conventional Method.” However, these methods require the disposal of, for example, acidic waste water that can cause serious environmental problems [69]. Currently, in connection with the emerging concept of “Green Chemistry,” several novel or nonconventional extraction methods and more effective techniques have been developed for pectin extraction (Table 1).

Figure 2.

The industrial pectin production process (Copyright Figure 1, [73]).

4.1 Chemical extraction methods

Chemical extraction using acids are the most commonly employed methods in the production of pectin, and it has an impact on content and composition both of pectin monomeric as xylose, arabinose, and galacturonic acid and physicochemical properties (emulsifying activity and viscosity, among others) [70].

For example, the acid extraction from pumpkin pulp and peel using pH 4 with EDTA showed around 7.5% of pectin as chelating agent [61]. Higher yields of pectin were obtained (up to 200 mg g−1) by combining the acid extraction with a microwave extraction, and those also displayed a large amount of phenolic compounds, proteins, and neutral sugars [18].

Other example for extraction yield of carrot pectin can range from 5.0 to 15.2% depending on parameters such as pH, temperature, heating time, sample provenance, and liquid/solid ratio [71].

The composition of the pectin from fabas and beans varied according to the extraction conditions: the neutral sugar galactose, arabinose, and rhamnose increased under slighter extraction conditions while glucose, mannose, and xylose sugars predominated under severe extraction. The maximum yield of extraction (15.75%) was recorded at pH 1.5 and at a temperature of 85°C for an 80-min extraction period and solid to liquid (1:25) ratio, while the highest degree of esterification (54.62%) occurred at pH 2.5 and at temperature of 90°C for a 60-min extraction period [72].

4.2 Novel technique extraction

Currently, novel and more effective techniques are inclined toward a cleaner (green or natural) extraction process. Among nonconventional extraction methods for pectin isolation are found (i) ultrasound-assisted extraction (UAE) (ii) subcritical water extraction (SWE), and (iii) microwave-assisted extraction (MAE). Table 1 summarizes the main pectin extraction process including solvent extraction (SE).

Other relevant novel technique for pectin isolation is enzyme-assisted extraction (EAE), which among other benefits, the unavoidable presence of trace chemical solvents in products from solvent-based extraction processes [73]. The enzymes are characterized to catalyze reactions such as hydrolysis, with a high level of selectivity, which either reduces the amount of solvent/chemical needed or increase yield for the same amount of solvent [74], in a way that is not feasible with acid-based hydrolysis. A distinction is possible between two approaches of EAE of pectin: (i) using enzymes that degrade pectin and help isolate pectin fragments, for example, galacturonic acid and (ii) using enzymes capable of deconstructing plant cell wall and isolating pectin [75].

Concretely combined enzymatic and ultrasonic treatment has been evaluated for efficient extraction of pectin in comparison with other extraction methods. The sequential treatment proposed by Yang et al. with enzyme (Celluclast 1.5 L) followed by ultrasound of sisal waste attained a much higher pectin yield of 31.1% than the ultrasound followed by enzyme, 14.6%. The pectin yield attained with the combined enzymatic/ultrasonic extraction was also higher than enzymatic extraction (9.4%), ultrasonic extraction (11.9%), and acidic extraction (5.8%) [76].

Advertisement

5. Innovative applications

Main pectin applications have been focused to food sector, while nonfood applications are currently presented as innovative alternative. Nonfood application includes the use in medical and pharmaceutical industries, where the health promoting benefits and bioactivities of pectin has shown potential for biomedical applications including drug delivery, tissue engineering, and wound healing [77].

5.1 Food sector

In the food sector, pectin traditional usage as a gelling agent, thickening agent, and stabilizer is being complemented by the emerging utilization of pectin as a fat replacer and also as health-promoting functional ingredient [4, 7, 8, 56].

Apple pomace and citrus peel remain the main sources for the production of commercial pectins, although other sources are being considered (Table 1) to the rising demand and growing interest for the development of innovative products [20, 55, 56].

5.2 Nonfood pectin applications

Nonfood applications include the use of pectin in pharmaceutical industry, where pectin bioactivity has shown an outstanding potential for biomedical applications as bioactive components [12] and also includes drug and gene delivery [13, 14], tissue engineering [2], and wound healing patches [56].

Within biomedical application, drug delivery systems improve the conventional drug administration as they provide in situ controlled and sustained release of active biomolecules. For most drug delivery systems, natural polymers, where pectin is included, are employed as inert, biocompatible carriers. These compounds have interesting properties such as the mucoadhesiveness, the ease of dissolution in basic environments associated to its resistance to proteases and amylases, making pectin suitable to release drugs in the colon [77]. And the ability to form gels in acid environments, instead, enhances the contact time of drugs for gastric or ocular treatments [78, 79]. Furthermore, pectin has been found to recognize galectin molecules, which are involved in different stages of cancer pathologies, being particularly appealing to target tumor cells for chemiotherapic treatments [80].

Furthermore, pectin has been described as an emerging prebiotic with the ability to modulate the bacterial composition of the colon microbiota [81] being able to exert beneficial effects on health [70]. The prebiotic activity is one of the most outstanding health benefits. Pectin, indigestible food-ingredients, has considered as an emerging prebiotic that possess prebiotic effect into the colon microbiota [9–11]. The beneficial effects of the pectin-derived oligosaccharides are essentially explained by the positive effects caused into bacterial population [82]. The consumption of pectins generates increasing levels of acetate, propionate, and butyrate, derived from the intestinal fermentation of pectin. These short-chain fatty acids possess an outstanding role in the prevention and treatment of metabolic syndrome, diarrhea, and intestinal disorders mainly Crohn’s disease and ulcerative colitis [83]. An example is the pumpkin cake, which is suggested to have a positive effect on gut bacteria [62].

Natural polymers have been employed as scaffolds to stimulate specific cell functions and to direct cell-cell interactions [77]. Few studies report the use of pectin for tissue regeneration; however, pectin hydrogels were found to have a great potential for bone tissue engineering applications, as they promote the nucleation of a mineral phase if immersed in adequate physiological solutions [84], with the formation of biomimetic constructs better mimicking the natural architecture of the bone [77].

Finally, other innovative applications in biomedical area are focused to wound healing patches. Hydrogel films on wounds or ulcers were aimed to prevent bacterial infection, support the autolytic debridement, and maintain a moist healing environment [77]. The natural properties of pectin impart several advantages to the wound dressings, such as hydrophilicity, which permits the removal of exudates, the retention of an acid environment, which may act as barrier against bacteria, and the ability for binding active molecules as drugs or growth factors to heal the wounds. As a wound heals, the cells around it are stimulated by growth factors to proliferate and grow into the wound [85].

A wide variety of hydrocolloid pectin-based wound dressings have been patented and are nowadays commercially available (Table 2).

Type of dressing Composition
Hydrocolloid sheets, uses: cavity or flat shallow wounds with low to medium exudates; absorbent; conformable; suitable for “problematic” areas: heel, elbow, sacrum
CombiDERM® (ConvaTec Ltd.) Cellulose, pectin, Salsorb90 (acrylic polymer)
DuoDERM® (ConvaTec Ltd.) Carboxymethylcellulose, pectin, propylene glycol
Granuflex® (ConvaTec Ltd.) Polyurethane foam sheet coated with pectin, gelatin and carboxymethylcellulose
Hydrocoll® (Hartmann) Pectin, gelatin and carboxymethylcellulose
Hydrocolloid paste, uses: useful debriding agent, conformable, may be left in place for several days
GranuGel® paste (ConvaTec Ltd.) Pectin, carboxymethylcellulose, propylene glycol
CitruGel® (Advances medical) Pectin, carboxymethylcellulose

Table 2.

Pectin-containing hydrocolloid wound dressing (Copyright Table 4, [77]).

References

  1. 1. Chan SY, Choo WS, Young DJ, Loh XJ. Pectin as a rheology modifier: Origin, structure, commercial production and rheology. Carbohydrate Polymers. 2017a;161:118-139
  2. 2. Mohnen D. Pectin structure and biosynthesis. Current Opinion in Plant Biology. 2008;11(3):266-277
  3. 3. Parre E, Geitmann A. Pectin and the role of the physical properties of the cell wall in pollen tube growth of solanum chacoense. Planta. 2005;220(4):582-592
  4. 4. Min B, Bae IY, Lee HG, Yoo SH, Lee S. Utilization of pectin-enriched materials from apple pomace as a fat replacer in a model food system. Bioresource Technology. 2010;101(14):5414-5418
  5. 5. Nakamura A, Furuta H, Maeda H, Takao T, Nagamatsu Y. Structural studies by stepwise enzymatic degradation of the main backbone of soybean soluble polysaccharides consisting of galacturonan and rhamnogalacturonan. Bioscience, Biotechnology, and Biochemistry. 2002;66(6):1301-1313
  6. 6. Saha D, Bhattacharya S. Hydrocolloids as thickening and gelling agents in food: A critical review. Journal of Food Science and Technology. 2010;47(6):587-597
  7. 7. Ciriminna R, Chavarria-Hernandez N, Hernandez AIR, Pagliaro M. Pectin: A new perspective from the biorefinery standpoint. Biofuels, Bioproducts and Biorefining. 2015;9(4):368-377
  8. 8. Peng Q , Xu Q , Yin H, Huang L, Du Y. Characterization of an immunologically active pectin from the fruits of lycium ruthenicum. International Journal of Biological Macromolecules. 2014;64:69-75
  9. 9. Babbar N, Dejonghe W, Gatti M, Sforza S, Elst K. Pectic oligosaccharides from agricultural by-products: Production, characterization and health benefits. Critical Reviews in Biotechnology. 2016;36(4):594-606
  10. 10. Gomez B, Gullon B, Yanez R, Schols H, Alonso JL. Prebiotic potential of pectins and pectic oligosaccharides derived from lemon peel wastes and sugar beet pulp: A comparative evaluation. Journal of Functional Foods. 2016;20:108-121
  11. 11. Li PJ, Xia JL, Nie ZY, Shan Y. Pectic oligosaccharides hydrolyzed from orange peel by fungal multi enzyme complexes and their prebiotic and antibacterial potentials. LWT-Food Science and Technology. 2016;69:203-210
  12. 12. Gunter EA, Popeyko OV. Calcium pectinate gel beads obtained from callus cultures pectins as promising systems for colon-targeted drug delivery. Carbohydrate Polymers. 2016;147:490-499
  13. 13. Liu L, Fishman ML, Kost J, Hicks KB. Pectin-based systems for colon-specific drug delivery via oral route. Biomaterials. 2003;24(19):3333-3343
  14. 14. Sriamornsak P. Application of pectin in oral drug delivery. Expert Opinion on Drug Delivery. 2011;8(8):1009-1023
  15. 15. Yapo BM. Pectic substances: From simple pectic polysaccharides to complex pectins—A new hypothetical model. Carbohydrate Polymers. 2011;86(2):373-385
  16. 16. Kaya M, Sousa AG, Crepeau MJ, Sorensen SO, Ralet MC. Characterization of citrus pectin samples extracted under different conditions: Influence of acid type and pH of extraction. Annals of Botany. 2014;114(6):1319-1326
  17. 17. Sila DN, Van Buggenhout S, Duvetter T, Fraeye I, De Roeck A, Van Loey A, et al. Pectins in processed fruit and vegetables: Part II–Structure-function relationships. Comprehensive Reviews in Food Science and Food Safety. 2009;8(2):86-104
  18. 18. Yoo SH, Lee BH, Lee H, Lee S, Bae IY, Lee HG, et al. Structural characteristics of pumpkin pectin extracted by microwave heating. Journal of Food Science. 2012;77(11):C1169-C1173
  19. 19. Gnanasambandam R, Proctor A. Determination of pectin degree of esterification by diffuse reflectance Fourier transform infrared pectroscopy. Food Chemistry. 2000;68(3):327-332
  20. 20. Muller-Maatsch J, Bencivenni M, Caligiani A, Tedeschi T, Bruggeman G, Bosch M, et al. Pectin content and composition from different food waste streams. Food Chemistry. 2016b;201:37-45
  21. 21. Fraeye I, Duvetter T, Doungla E, Van Loey A, Hendrickx M. Fine-tuning the properties of pectin calcium gels by control of pectin fine structure, gel composition and environmental conditions. Trends in Food Science & Technology. 2010;21(5):219-228
  22. 22. Tromp RH, de Kruif CG, Van Eijk M, Rolin C. On the mechanism of stabilisation of acidified milk drinks by pectin. Food Hydrocolloids. 2004;18(4):565-572
  23. 23. Voragen AGJ, Coenen GJ, Verhoef RP, Schols HA. Pectin, a versatile polysaccharide present in plant cell walls. Structural Chemistry. 2009;20(2):263-275
  24. 24. Bonnin E, Dolo E, Le Goff A, Thibault JF. Characterisation of pectin subunits released by an optimised combination of enzymes. Carbohydrate Research. 2002;337(18):1687-1696
  25. 25. Ovodov YS. Current views on pectin substances. Russian Journal of Bioorganic Chemistry. 2009;35(3):269-284
  26. 26. Zandleven J, Beldman G, Bosveld M, Schols HA, Voragen AGJ. Enzymatic degradation studies of xylogalacturonans from apple and potato, using xylogalacturonan hydrolase. Carbohydrate Polymers. 2006;65(4):495-503
  27. 27. Le Goff A, Renard C, Bonnin E, Thibault JF. Extraction, purification and chemical characterisation of xylogalacturonans from pea hulls. Carbohydrate Polymers. 2001;45(4):325-334
  28. 28. Yu LY, Mort AJ. Partial characterization of xylogalacturonans from cell walls of ripe watermelon fruit: Inhibition of endopolygalacturonase activity by xylosylation. Progress in Biotechnology. Vol. 14. 1996. pp. 79-88
  29. 29. Renard C, Crepeau MJ, Thibault JF. Structure of the repeating units in the rhamnogalacturonic backbone of apple, beet and citrus pectins. Carbohydrate Research. 1995;275(1):155-165
  30. 30. Maxwell EG, Belshaw NJ, Waldron KW, Morris VJ. Pectin—An emerging new bioactive food polysaccharide. Trends in Food Science and Technology. 2012;24(2):64-73
  31. 31. Caffall KH, Mohnen D. The structure, function, and biosynthesis of plant cell wall pectic polysaccharides. Carbohydrate Research. 2009;344(14):1879-1900
  32. 32. Yapo BM, Wathelet B, Paquot M. Comparison of alcohol precipitation and membrane filtration effects on sugar beet pulp pectin chemical features and surface properties. Food Hydrocolloids. 2007;21(2):245-255
  33. 33. Coenen GJ, Bakx EJ, Verhoef RP, Schols HA, Voragen AGJ. Identification of the connecting linkage between homo- or xylogalacturonan and rhamnogalacturonan type I. Carbohydrate Polymers. 2007;70(2):224-235
  34. 34. Leivas CL, Iacomini M, Cordeiro LM. Pectic type II arabinogalactans from starfruit (Averrhoa carambola L.). Food Chemistry. 2016;199:252-257
  35. 35. Palin R, Geitmann A. The role of pectin in plant morphogenesis. Bio Systems. 2012;109(3):397-402
  36. 36. Ishii T, Matsunaga T, Pellerin P, O’Neill MA, Darvill A, Albersheim P. The plant cell wall polysaccharide rhamnogalacturonan II self-assembles into a covalently cross-linked dimer. The Journal of Biological Chemistry. 1999;274(19):13098-13104
  37. 37. Lara-Espinoza C, Carvajal-Millan E, Balandran-Quintana R, Lopez-Franco Y, Rascon-Chu A. Pectin and pectin-based composite materials: Beyond food texture. Molecules. 2018;23(942):1-35
  38. 38. Yapo BM, Robert C, Etienne I, Wathelet B, Paquot M. Effect of extraction conditions on the yield, purity and surface properties of sugar beet pulp pectin extracts. Food Chemistry. 2007;100(4):1356-1364
  39. 39. Masmoudi M, Besbes S, Ben Thabet I, Blecker C, Attia H. Pectin extraction from lemon by-product with acidified date juice: Rheological properties and microstructure of pure and mixed pectin gels. Food Science and Technology International. 2010;16(2):105-114
  40. 40. Masmoudi M, Besbes S, Chaabouni M, Robert C, Paquot M, Blecker C, et al. Optimization of pectin extraction from lemon by-product with acidified date juice using response surface methodology. Carbohydrate Polymers. 2008;74(2):185-192
  41. 41. Shalini R, Gupta DK. Utilization of pomace from apple processing industries: A review. Journal of Food Science and Technology. 2010;47(4):365-371
  42. 42. Rascon-Chu A, Martinez-Lopez AL, Carvajal-Millan E, de Leon-Renova NEP, Marquez-Escalante J, Romo-Chacon A. Pectin from low quality ‘Golden Delicious’ apples: Composition and gelling capability. Food Chemistry. 2009;116(1):101-103
  43. 43. Gigli J, Garnier C, Piazza L. Rheological behaviour of low-methoxyl pectin gels over an extended frequency window. Food Hydrocolloids. 2009;23(5):1406-1412
  44. 44. Ververis C, Georghiou K, Danielidis D, Hatzinikolaou DG, Santas P, Santas R, et al. Cellulose, hemicelluloses, lignin and ash content of some organic materials and their suitability for use as paper pulp supplements. Bioresource Technology. 2007;98(2):296-301
  45. 45. Methacanon P, Krongsin J, Gamonpilas C. Pomelo (citrus maxima) pectin: Effects of extraction parameters and its properties. Food Hydrocolloids. 2014;35:383-391
  46. 46. Liu H, Zhu DS, Xu XM, Guo SD, Jin ZY. Rheological, textural and microstructure analysis of the high-methoxy pectin/gelatin mixed systems. Journal of Texture Studies. 2007;38(5):577-600
  47. 47. Chen HM, Fu X, Luo ZG. Effect of molecular structure on emulsifying properties of sugar beet pulp pectin. Food Hydrocolloids. 2016;54:99-106
  48. 48. Li DQ , Du GM, Jing WW, Li JF, Yan JY, Liu ZY. Combined effects of independent variables on yield and protein content of pectin extracted from sugar beet pulp by citric acid. Carbohydrate Polymers. 2015;129:108-114
  49. 49. Ralet MC, Crepeau MJ, Buchholt HC, Thibault JF. Polyelectrolyte behaviour and calcium binding properties of sugar beet pectins differing in their degrees of methylation and acetylation. Biochemical Engineering Journal. 2003;16(2):191-201
  50. 50. Guo XB, Guo XM, Yu SJ, Kong FS. Influences of the different chemical components of sugar beet pectin on the emulsifying performance of conjugates formed between sugar beet pectin and whey protein isolate. Food Hydrocolloids. 2018;82:1-10
  51. 51. Funami T, Zhang GY, Hiroe M, Noda S, Nakauma M, Asai I, et al. Effects of the proteinaceous moiety on the emulsifying properties of sugar beet pectin. Food Hydrocolloids. 2007;21(8):1319-1329
  52. 52. Colquhoun IJ, Ralet MC, Thibault JF, Faulds CB, Williamson G. Structure identification of feruloylated oligosaccharides from sugar-beet pulp by NMR-spectroscopy. Carbohydrate Research. 1994;263(2):243-256
  53. 53. Zaidel DNA, Meyer AS. Biocatalytic cross-linking of pectic polysaccharides for designed food functionality: Structures, mechanisms, and reactions. Biocatalysis and Agricultural Biotechnology. 2012;1(3):207-219
  54. 54. Christiaens S, Uwibambe D, Uyttebroek M, Van Droogenbroeck B, Van Loey AM, Hendrickx ME. Pectin characterisation in vegetable waste streams: A starting point for waste valorisation in the food industry. LWT-Food Science and Technology. 2015;61(2):275-282
  55. 55. Muller-Maatsch J, Bencivenni M, Caligiani A, Tedeschi T, Bruggeman G, Bosch M, et al. Pectin content and composition from different food waste streams. Food Chemistry. 2016a;201:37-45
  56. 56. Dranca F, Oroian M. Extraction, purification and characterization of pectin from alternative sources with potential technological applications. Food Research International. 2018;113:327-350
  57. 57. Iglesias MT, Lozano JE. Extraction and characterization of sunflower pectin. Journal of Food Engineering. 2004;62(3):215-223
  58. 58. Jimenez A, Rodriguez R, Fernandez-Caro I, Guillen R, Fernandez-Bolanos J, Heredia A. Olive fruit cell wall: Degradation of pectic polysaccharides during ripening. Journal of Agricultural and Food Chemistry. 2001;49(1):409-415
  59. 59. Rubio-Senent F, Rodriguez-Gutierrez G, Lama-Munoz A, Garcia A, Fernandez-Bolanos J. Novel pectin present in new olive mill wastewater with similar emulsifying and better biological properties than citrus pectin. Food Hydrocolloids. 2015;50:237-246
  60. 60. Stabnikova O, Wang JY & Ivanov V. Value-added biotechnological products from organic wastes. In: Wang LK, Ivanov V, Tay JH, Hung YT editors. Environmental Biotechnology. Totowa-NJ, USA: Humana Press; 2010;10:343-394
  61. 61. Kost’alova Z, Hromadkova Z, Ebringerova A. Structural diversity of pectins isolated from the styrian oil-pumpkin (cucurbita pepo var. styriaca) fruit (vol 93, pg 163, 2013). Carbohydrate Polymers. 2014;99:831-831
  62. 62. Jun HI, Lee CH, Song GS, Kim YS. Characterization of the pectic polysaccharides from pumpkin peel. LWT-Food Science and Technology. 2006;39(5):554-561
  63. 63. Oliveira TIS, Rosa MF, Cavalcante FL, Pereira PHF, Moates GK, Wellner N, et al. Optimization of pectin extraction from banana peels with citric acid by using response surface methodology. Food Chemistry. 2016;198:113-118
  64. 64. Banerjee J, Singh R, Vijayaraghavan R, MacFarlane D, Patti AF, Arora A. Bioactives from fruit processing wastes: Green approaches to valuable chemicals. Food Chemistry. 2017;225:10-22
  65. 65. Petkowicz CLO, Vriesmann LC, Williams PA. Pectins from food waste: Extraction, characterization and properties of watermelon rind pectin. Food Hydrocolloids. 2017;65:57-67
  66. 66. Del Valle M, Camara M, Torija ME. Chemical characterization of tomato pomace. Journal of the Science of Food and Agriculture. 2006;86(8):1232-1236
  67. 67. Round AN, Rigby NM, MacDougall AJ, Morris VJ. A new view of pectin structure revealed by acid hydrolysis and atomic force microscopy. Carbohydrate Research. 2010;345(4):487-497
  68. 68. Guillon F, Moise A, Quemener B, Bouchet B, Devaux MF, Alvarado C, et al. Remodeling of pectin and hemicelluloses in tomato pericarp during fruit growth. Plant Science. 2017;257:48-62
  69. 69. Yapo BM. Lemon juice improves the extractability and quality characteristics of pectin from yellow passion fruit by-product as compared with commercial citric acid extractant. Bioresource Technology. 2009;100(12):3147-3151
  70. 70. Pacheco MT, Villamiel M, Moreno R, Moreno FJ. Structural and rheological properties of pectins extracted from industrial sugar beet by-products. Molecules. 2019;24:392-408
  71. 71. Jafari F, Khodaiyan F, Kiani H, Hosseini SS. Pectin from carrot pomace: Optimization of extraction and physicochemical properties. Carbohydrate Polymers. 2017;157:1315-1322
  72. 72. Korish M. Faba bean hulls as a potential source of pectin. Journal of Food Science and Technology. 2015;52(9):6061-6066
  73. 73. Adetunji LR, Adekunle A, Orsat V, Raghavan V. Advances in the pectin production process using novel extraction techniques: A review. Food Hydrocolloids. 2017;62:239-250
  74. 74. Puri M, Sharma D, Barrow CJ. Enzyme-assisted extraction of bioactives from plants. Trends in Biotechnology. 2012;30(1):37-44
  75. 75. Panouillé M, Thibaul JF, Bonnin E. Cellulase and protease preparations can extract pectins from various plant byproducts. Journal of Agricultural and Food Chemistry. 2006;54(23):8926-8935
  76. 76. Yang Y, Wang Z, Hu D, Xiao K, Wu JY. Efficient extraction of pectin from sisal waste by combined enzymatic and ultrasonic process. Food Hydrocolloids. 2018;79:189-196
  77. 77. Munarin F, Tanzi MC, Petrini P. Advances in biomedical applications of pectin gels. International Journal of Biological Macromolecules. 2012;51:681-689
  78. 78. Sriamornsak P, Wattanakorn N, Takeuchi H. Study on the mucoadhesion mechanism of pectin by atomic force microscopyand mucin-particle method. Carbohydrate Polymers. 2010;79:54-59
  79. 79. Ludwig A. The use of mucoadhesive polymers in ocular drug delivery. Advanced Drug Delivery Reviews. 2005;57:1595-1639
  80. 80. Glinsky VV, Raz A. Modified citrus pectin anti-metastatic properties: One bullet, multiple targets. Carbohydrate Research. 2009;344:1788-1791
  81. 81. Ferreira-Lazarte A, Kachrimanidou V, Villamiel M, Rastall RA, Moreno FJ. In vitro fermentation properties of pectins and enzymatic-modified pectins obtained from different renewable bioresources. Carbohydrate Polymers. 2018;199:482-491
  82. 82. Gullón B, Gullón P, Tavaria F, Pintado M, Gomes AM, Alonso JL, et al. Structural features and assessment of prebiotic activity of refined arabinoxylooligosaccharides from wheat bran. Journal of Functional Foods. 2014;6:438-449
  83. 83. Islamova ZI, Ogai DK, Abramenko OI, Lim AL, Abduazimov BB, Malikova MK, et al. Comparative assessment of the prebiotic activity of some pectin polysaccharides. Pharmaceutical Chemistry Journal. 2017;51(4):288-291
  84. 84. Munarin F, Giuliano L, Bozzini S, Tanzi MC, Petrini P. Materials Science and Engineering C. 2010;30:491-496
  85. 85. Harris R, Nasi W. Gels for Use in Wound Management. World Patent no. WO/2008/015475; 2008
  86. 86. Chan SY, Choo WS. Effect of extraction conditions on the yield and chemical properties of pectin from cocoa husks. Food Chemistry. 2013;141(4):3752-3758
  87. 87. Hosseini SS, Khodaiyan F, Yarmand MS. Aqueous extraction of pectin from sour orange peel and its preliminary physicochemical properties. International Journal of Biological Macromolecules. 2016;82:920-926
  88. 88. Maran JP, Priya B. Multivariate statistical analysis and optimization of ultrasound-assisted extraction of natural pigments from waste red beet stalks. Journal of Food Science and Technology. 2016;53(1):792-799
  89. 89. Andersen NM, Cognet T, Santacoloma PA, Larsen J, Armagan I, Larsen FH, et al. Dynamic modelling of pectin extraction describing yield and functional characteristics. Journal of Food Engineering. 2017;192:61-71
  90. 90. Wikiera A, Mika M, Starzynska-Janiszewska A, Stodolak B. Endo-xylanase and endo-cellulase-assisted extraction of pectin from apple pomace. Carbohydrate Polymers. 2016;142:199-205
  91. 91. Jeong HS, Kim HY, Ahn SH, Oh SC, Yang I, Choi IG. Optimization of enzymatic hydrolysis conditions for extraction of pectin from rapeseed cake (Brassica napus L.) using commercial enzymes. Food Chemistry. 2014;157:332-338
  92. 92. Dominiak M, Sondergaard KM, Wichmann J, Vidal-Melgosa S, Willats WGT, Meyer AS, et al. Application of enzymes for efficient extraction, modification, and development of functional properties of lime pectin. Food Hydrocolloids. 2014;40:273-282
  93. 93. Wang M, Huang B, Fan C, Zhao K, Hu H, Xu X, et al. Characterization and functional properties of mango peel pectin extracted by ultrasound assisted citric acid. International Journal of Biological Macromolecules. 2016;91:794-803
  94. 94. Koubala BB, Kansci G, Mbome LI, Crepeau MJ, Thibault JF, Ralet MC. Effect of extraction conditions on some physicochemical characteristics of pectins from “Amelioree” and “mango” mango peels. Food Hydrocolloids. 2008;22(7):1345-1351
  95. 95. Moorthy IG, Maran JP, Ilakya S, Anitha SL, Sabarima SP, Priya B. Ultrasound assisted extraction of pectin from waste Artocarpus heterophyllus fruit peel. Ultrasonics Sonochemistry. 2017;34:525-530
  96. 96. Wang X, Chen QR, Lu X. Pectin extracted from apple pomace and citrus peel by subcritical water. Food Hydrocolloids. 2014;38:129-137
  97. 97. Chen HM, Fu X, Luo ZG. Properties and extraction of pectin-enriched materials from sugar beet pulp by ultrasonic-assisted treatment combined with subcritical water. Food Chemistry. 2015;168:302-310
  98. 98. Maran J, Prakash K. Process variables influence on microwave assisted extraction of pectin from waste Carcia papaya L. peel. International Journal of Biological Macromolecules. 2015;73:202-206
  99. 99. Koubala B, Christiaens S, Kansci G, Van Loey AM, Hendrickx ME. Isolation and structural characterisation of papaya peel pectin. Food Research International. 2014;55:215-221
  100. 100. Lefsiha K, Giacomazzab D, Dahmoune F, Mangione MR, Bulone D, San Biagio PL, et al. Pectin from opuntia ficus indica: Optimization of microwave-assistedextraction and preliminary characterization. Food Chemistry. 2017;221:91-99

Written By

Virginia Rodríguez Robledo* and Lucía Isabel Castro Vázquez

Submitted: 15 February 2019 Reviewed: 03 March 2019 Published: 03 December 2019