Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 4076 2023-03-20 19:43:32 |
2 format Meta information modification 4076 2023-03-22 03:16:57 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Martínez-Sabando, J.; Coin, F.; Melillo, J.H.; Goyanes, S.; Cerveny, S. Pectin-Based Material for Applications in Water Treatment. Encyclopedia. Available online: https://encyclopedia.pub/entry/42364 (accessed on 05 May 2024).
Martínez-Sabando J, Coin F, Melillo JH, Goyanes S, Cerveny S. Pectin-Based Material for Applications in Water Treatment. Encyclopedia. Available at: https://encyclopedia.pub/entry/42364. Accessed May 05, 2024.
Martínez-Sabando, Javier, Francesco Coin, Jorge H. Melillo, Silvia Goyanes, Silvina Cerveny. "Pectin-Based Material for Applications in Water Treatment" Encyclopedia, https://encyclopedia.pub/entry/42364 (accessed May 05, 2024).
Martínez-Sabando, J., Coin, F., Melillo, J.H., Goyanes, S., & Cerveny, S. (2023, March 20). Pectin-Based Material for Applications in Water Treatment. In Encyclopedia. https://encyclopedia.pub/entry/42364
Martínez-Sabando, Javier, et al. "Pectin-Based Material for Applications in Water Treatment." Encyclopedia. Web. 20 March, 2023.
Pectin-Based Material for Applications in Water Treatment
Edit

Climate change and water are inseparably connected. Extreme weather events cause water to become more scarce, polluted, and erratic than ever. Pectin is a biodegradable polymer, extractable from vegetables, and contains several hydroxyl and carboxyl groups that can easily interact with the contaminant ions. In addition, pectin-based materials can be prepared in different forms (films, hydrogels, or beads) and cross-linked with several agents to change their molecular structure. Consequently, the pectin-based adsorbents can be tuned to remove diverse pollutants such as divalent metals. 

pectin low methoxylation degree adsorption cross-linking

1. Pectin—Chemical Structure

Pectin is an anionic heteropolysaccharide that is present in vegetable cell walls. Pectin is particularly predominant in fruit peels, especially in citrus peel and apple pomace, but also in passion fruit rind, pomelo, and banana peel [1]. The pectin backbone chain comprises α-(1,4)-linked D-galacturonic acid (GalA) residues [2] (mainly homogalacturonan (HG), rhamnogalacturonan I (RG-I), rhamnogalacturonan (RG-II)) linked by α 1-4 glycosidic bond. The homogalacturonan (HG) is the most abundant and linear domain (around 60%).
The carbon 6 of the D-galacturonic acid residues can be methyl-esterified (COOCH3) or carboxylated (COO, in its deprotonated form). It also can be acetylated at the O-2 and/or O-3 position of the GalA residues, although this is less common [2]. Depending on the ratio of the methyl-esterification of these residues, pectin shows a different esterification degree (DE).
Carboxyl groups are hydrophilic and can coordinate with metal ions, whereas methyl-esterified are phobic. Therefore, by maximizing the number of COO (i.e., low degree of esterification), the adsorption efficiency against metal ions is much higher than pectin with a high amount of COOCH3 groups.
DE is a magnitude that is related to the gelling mechanism of pectin. It classifies pectin into two groups: high methoxyl pectins (HM), where more than 50% of the galacturonic acid residues are esterified, and low methoxyl (LM) pectins, where less than 50% of the galacturonic acid residues are methoxylated. Independent of the DE, pectin is a water-soluble biopolymer. Therefore, it should become insoluble to be used as an adsorbent in aqueous environments. This can be achieved by cross-linking the pectin with various cross-linking agents, as described in the next section.
Pectin sources (i.e., the plant used to extract pectin) strongly influence the galacturonic acid content and the degree of methoxylation obtained after extraction [3][4]. These two variables directly affect pectin’s ability to form a gel. The most common sources of pectin are apple pomace and citrus peels. Whereas the galacturonic acid content for apple pomace is between 20 and 44%, and the orange peel has been higher in comparison.
There are different extraction methods to isolate pectin from vegetables, which also affect the structure of the pectin, and it involves several steps (pretreatment, extraction, purification, concentration, precipitation, drying, and standardization) [5]. The preferable method is to mix the cell-wall materials in hot or cold acidified water [6][7] because the highest extraction yield is achieved. It is also possible to add chelating agents [8] such as cyclohexane-diamine-tetra acetic acid (CDTA), oxalate, or ethylene-diamine-tetra acetic acid (EDTA) to promote the release of pectin. On the other hand, if instead an acidic alkaline extraction is used, the length of the galacturonic acid, the methoxylation degree (DM), and acetylation (DA) decrease [9].
The possible uses of pectin in different industries are abundant and of different natures, mainly in the health and pharmacy sector, food applications, and packaging. As pectin is a natural component of vegetables and fruits, it is an exceptionally safe multifunctional food additive (E440) that is used, for example, as a texturizer or gelling agent. In addition, pectin is popular in several scientific fields because of its availability, safety, relatively low price, and functionality. Specific structural attributes developed by pectin’s functional groups or the attachment of chemicals on the molecule make it a good candidate for several purposes, such as food innovations [10], nutritional remediation [11], drug delivery [12], illness treatment [13], tissue engineering [14], and other approaches [15]. Notably, pectin is resistant to gastrointestinal hydrolyzing enzymes and acidic/alkaline media, which favors its application in colon delivery via an oral route under a specific condition.
In particular, pectin, with a low level of esterification, also has a high application value and broad application prospects as a functional food ingredient [16]. It also has been studied as a forthcoming biomaterial for tissue engineering and biomedical applications [17]. Pectin films have also been used in packaging, although there are some limitations because of their inadequate mechanical properties and high hydrophilicity of the pectin films [18][19]. This issue can be solved by blending with other polymers [20] or incorporating cross-linkers and/or filler material [21].

2. Insoluble Pectin—Cross-Linking Agents That Generate Gelation

As mentioned above, pectin is a water-soluble material. Therefore, it must be insolubilized to be used as an adsorbent to remove pollutants from water. Different ionic crosslinking agents were used for pectin The cations monovalent (Na+, K+), divalent (Ca2+, Cu2+, Sr2+ Ni2+ Zn2+ Cd2+ Pb2+ or Mg2+), or trivalent (Al3+, La3+, and Fe3+) are the most used. Pectin can also be cross-linked using non-ionic cross-linkers such as glutaraldehyde [22] or laccase [23]. However, ionic crosslinking is the most widely used and suitable method because it has the principal advantage of releasing contaminants (in some cases, recovering a priceless pollutant) and recovering the absorbent that can be reused.
Among all the possible ions, the divalent cations are the most commonly used elements for LM pectin gelation. After cross-linking (independent of the cross-linking agent), pectin undergoes gelation, producing a three-dimensional network in the form of xerogels (vacuum drying), hydrogels (wet gels), aerogels (supercritical CO2 drying), and cryo-gels (freeze-drying) depending on the drying conditions [24]. The primary characteristic of wet gels is their ability to bind considerable quantities of water, thus increasing their volume. In addition, the swelling favors the access of pollutants to pectin.
The gelation mechanism of pectin is mainly governed by the methyl-esterification degree. Therefore, the gel formation mechanism differs for high-methoxyl (HM) and low-methoxyl (LM) pectins [25]. Apart from the methyl-esterification degree, the gelling process is influenced by ionic strength, molecular weight, and pH [26]. In addition, after cross-linking, the final structure of pectin depends on the cross-linking cation used.
There are fewer reports on LM pectin cross-linking using monovalent cations such as sodium (Na+) and potassium (K+). Cross-linking with these monovalent cations is produced at a low pH (approximately from pH = 2 to pH = 4) [27][28][29][30]. The cross-linking with monovalent ions diminishes the repulsive charges between pectin chains and promotes chain-chain association via hydrogen bonding [27][30]. It was suggested that at pH = 4.5, the cross-linking would more effective using K+ than Na+ because of the charge screening, galacturonic acid de-esterification, hydrogen bond changes, and electronic attraction [27]. Moreover, by increasing the pH to higher values (alkaline conditions), it was also found [27] that Na+ ions generated much stronger gels than those induced by K+ in HM pectin. In accordance, Yoo and coworkers [31] also studied poly (methyl esterase) mediated de-esterified citrus pectin when cross-linked with Na+, K+, and Li+. In agreement with Chen et al. [27], they observed that stronger cross-linking was produced using Na+ followed by K+ at pH = 7, whereas the reverse was true for pH = 5. A similar study was performed by Strom et al. [30][31], where the influence of L+, Na+, and K+ on LM pectin rheology in an acid solution was evaluated. These investigations revealed that K+ formed the strongest gels, followed by Na+ and Li+. Furthermore, LM pectin was cross-linked by NaCl (0.05M) with and without ZnCl2 [29] and could form nanoparticles. Li+ does not form gels at any of the tested conditions. This behavior was ascribed to an increase in the ionic radius of the hydrated cation.
Trivalent cations have also been employed as cross-linking agents for pectin [32][33][34][35]. It was found [32][33] that independent of the methyl-esterification degree, Al3+ binds pectin chains had the condition of pH ≥ 4. In a rheological investigation [33], Ca2+, Cu2+, Al3+, and La3+ ions were used as cross-linking agents. All these cations can form pectin gels. It was stated that the weakest gel was formed with La3+, followed by Ca2+, similar to Al3+, while Cu2+ was one order of magnitude more substantial than the other cations. Finally, it was reported that the trivalent cations Fe3+, Ce3+, Pr3+, and Nd3+ could establish cross-links in pectin [35]. In that work, the formation of isolated rhamnogalacturonan II dimers was reported. Still, there were no reports regarding what type of structure the pectin chains formed in the presence of these trivalent metals.
Apart from monovalent or trivalent cations, divalent cations such as Ca2+ are the most used cross-linking agents for pectin. The gelation of LM pectin with calcium ions occurs by forming junction zones according to the so-called “egg-box” model, which has been initially described for alginates [36]. The lower the pectin esterification degree, the more divalent the cation binding sites that are available on pectin, and the ability to absorb divalent cations increases. The following section will discuss the egg-box model from a historical perspective.
Many works have proven the effective cross-linking of LM pectin using Ca2+ [37][38][39][40][41][42][43][44][45]. In addition to Ca2+, the cross-linking of LM pectin with Cd2+ and Cu2+ was reported for drug delivery applications [46]; Pectin was also cross-linked with Sr2+ and Zn2+ to form aerogels for diclofenac sodium controlled release [47]. Pectin cross-linked by Cu2+ ions was reported to be used as a scaffold for gold nanoparticles [48]. Moreover, pectin was cross-linked with a 2% (w/v) ZnCl2 solution [49] to evaluate the potential healthcare applicability, and Das and coworkers [50] used LM pectin cross-linked with Zn2+ to encapsulate a colon-specific drug delivery microsphere.
On the other hand, other molecules have been used to induce the gelation of LM pectin. Yoshimura and coworkers [22] described the cross-linking of the LM pectin with ethylene glycol diglycidyl ether and glutaraldehyde or Ca2+. They found that glutaraldehyde produced pectin gels, whereas the ethylene glycol diglycidyl ether did not show an apparent gel formation. Ullah and coworkers [51] reported the cross-link of pectin with methylene bisacrylamide and ammonium persulfate as initiators. Pre-saponified pectin was also cross-linked with adipic acid [52] or adipic acid dihydrazide, which was previously involved in the oxidation of pectin [53]. Chen and coworkers compared the cross-link using adipic acid dihydrazide and the classic Ca2+ cross-link, reporting that the adipic acid dihydrazide cross-link significantly improved cell adhesion. Another example of a non-ionic gelling agent is laccase, which was employed to cross-link sugar beet pectin, as performed by Jung and Wicker [54], where they noticed an increase in the molecular weight of pectin by chromatographic techniques.

3. Structural Models of Cross-Linked LM Pectin and Alginates

3.1. Historical Perspective of Structural Models Induced by Calcium

In 1973, Grant and coworkers [36] first proposed the interaction between a polysaccharide chain (such as alginate or pectin) and divalent cations in terms of the egg-box model. Initially, this model introduced alginates: a non-branched polysaccharide composed of two monosaccharides, α–L guluronic and β-D mannuronic acids [55]. Using circular dichroism, the authors suggested a cooperative inter-chain mechanism of ion binding involving two chains associated with the coulomb interaction between the active sites and the metal ion. More specifically, two opposite α-L-guluronate sequences paired together to form a structure with cavities, within which Ca2+ ions were accommodated through specific coordination interactions between two free carboxyl groups.
Five years later, Morris et al. [56] studied alginate chain gelation by X-ray diffraction coupled with circular dichroism using different calcium concentrations. Their results showed that the primary inter-chain association mechanism was regularly dimerizing alginate or poly (guluronate) chain segments. This structure results in cross-linked dimers with a geometric design that resembles an egg box. Compared with the first proposal of the “egg-box” model, the dimers can further aggregate laterally into multimers, as revealed by small-angle X-ray scattering. In addition, the sugar ring belonging to guluronic acid and the polymer chain adopts a characteristic zigzag shape. Studying the lateral aggregation of dimers is essential to understanding the properties of these biomaterials in terms of their structure to explain what happens during a water remediation process. On the other hand, knowing the correct features and operating principle of the “egg-box” model helps to understand the possible pollutants of adsorption models from wastewater.
In 1981, Walkinshaw and Arnott [57] added, for the first time, the concept of the junction zone for the “egg-box” model for poly(galacturonic) acid and a high methoxylation degree of calcium pectate (i.e., pectin cross-linked with metal ions). In this new model, for poly(galacturonic) acid, chain–chain interactions were stabilized by intermolecular hydrogen bonds formed between several adjacent D-galacturonic acid units and hydrophobic bonding between methyl esters. Similarly, hydrophobic binding from segments of methyl groups and specific intermolecular hydrogen bonds stabilized calcium pectate. The main interactions between pairs of chains could be the bridges formed by calcium ions, which incorporate into their coordination shells two-polyanion oxygen atoms from one chain and three from the other. In addition, in 1982, Powell et al. [58] investigated rhamnose distribution for the formation of stable poly(galacturonic) acid inter-chain junctions for high and low degrees of methyl(esterification). Their results confirm the Walkinshaw and Arnott hypothesis [57]. They also indicate that the length of the poly(galacturonic) acid sequences between rhamnose interruptions (i.e., hairy regions) is approximately constant (i.e., they prove the existence of the junction zone).
Later, Kohn [59] studied the interaction between carboxyl groups and divalent cations in pectin fragments, resulting in a complex formation. The interaction exhibits different affinities towards divalent cations: Ca2+, Sr2+, and Zn2+ ions are bound according to the well-known “egg-box” model, whereas Cd2+, Cu2+, and Pb2+ form intra- and inter-molecular complexes.
At this point, and to bring all the above studies together, a book entitled the chemistry and technology of pectin [60] was first published in 1991. This event changed scientists’ attitudes toward pectin research because of the exponential growth of publications in the following years.
During the first ten years of the 2000s, scientists began to study the “egg-box” model using various approaches, and different reconsiderations of the model have been proposed. A molecular simulation study [61], made on pairs of galacturonate oligomers, found a shift along the chain axis in the association of associated galacturonate chains. Therefore, they proposed the “shifted egg-box” model to explain the gelation mechanism of galacturonate chains, where two shifted antiparallel pectin chains are necessary to form the egg-box dimer.
In 2007, the “shifted egg-box” model for Ca-alginate gels was re-examined [62] and confirmed using X-ray diffraction measurements. Unlike previous “egg-box” models developed for pectin, in this case, the authors considered the helical conformation of the main chain, which had not been considered before. These results suggest that instead of a 2/1 helical conformation, as previously proposed for the egg-box model [63], a 3/1 helical conformation for the junction zones was energetically more favorable for these materials. In the same year, Fang et al. [64] studied the Ca2+-alginate gel of high (long polymer chain) and low molecular weight (short polymer chain) using isothermal titration calorimetry. They proposed a three-step binding behavior of calcium to alginate to form “egg-box” dimers: (i) the interaction of Ca2+ with a single guluronate unit forming mono-complexes (i.e., a carboxylic group with Ca2+); (ii) the propagation and formation of egg-box dimers via the pairing of these mono-complexes; and (iii) the lateral association of the egg-box dimers, generating multimers. The short chains are quite rigid, and inter-cluster association is the only possible way for the “egg-box” dimers to aggregate laterally. This aggregation increases in molecular size. However, long chains are more flexible and have smaller clusters. The intra-cluster association, in this case, results in a reduction in molecular size.
In 2008, Fang and coworkers [65] compared pectin with alginate for calcium-binding behavior. LM pectin is similar to alginate, but a two-step mechanism is involved. Whereas step I can be attributed to a mono-complex between Ca2+ ions and a polygalacturonate chain, step II is the formation of egg-box dimers through the pairing of the mono-complexes. In addition to steps I and II, the binding of alginate with Ca2+ includes a third step that is assigned to the lateral association of egg-box dimers.
In 2010, Gohil [66] studied the structural reorganization of pectin and alginate films after calcium binding by X-Ray diffraction and dynamic mechanical measurements. His results suggest that the structural reorganization of molecular network structure, after binding with Ca2+, destroyed some existing pectin crystalline tie points, resulting in its amorphization, as described by the “fringe-micellar” structure. In the same year, Fraeye and coworkers [41] summarized in detail how the characteristics of the final pectin gel are affected by different “intrinsic” and “extrinsic” parameters such as the amount and distribution of methyl ester, chain length, pectin concentration, amidation, and acetylation degrees, molecular weight, calcium content, temperature, and pH.
Ventura and coworkers [67] published a SAXS analysis of Ca2+–LM pectin gels with different calcium concentrations, coupled with molecular dynamics simulation studies. They proposed a new “egg-box” model considering semi-flexible and no linear chains. This chain flexibility was not considered in any of the previous models developed. They showed [67] that rod-like and point-like cross-links between neighboring pectin molecules could occur. With the increased Ca2+ concentration, the number of road-like cross-links decreased while the number of point-like cross-links increased.
Wang et al. [68] studied the high and low methoxyl pectin self-assembly molecules, which were regulated by calcium ions using atomic force microscopy (AFM) in the following years. The addition of calcium ions increased the viscosity of the low-methoxyl pectin solution. Otherwise, the viscosity of the high-methoxyl pectin solution remained stable. AFM confirmed that the esterification degree and calcium concentration cause different binding methods with calcium ions between pectin governed by the branched nature of pectin [67]. The formation of dimers by lateral aggregation was observed only for the LM-pectin. However, for the HM-pectin, the large number of esterified galacturonic acid residues limits pectin fiber aggregation due to the nonspecific hydrophobic interaction and hydrogen bonds.

3.2. Recent Evolution of Structural Models for Pectin Is Cross-Linked with Other Metals

As mentioned above, pectin is susceptible to cross-links with other cations apart from calcium. Specifically, in 2015, Assifaoui et al. [69] published the structural differences found in LM pectin when cross-linked with Ca2+ and Zn2+. These results show that Ca2+ cations only interact with carboxyl groups and form more homogeneous pectin network fibers (i.e., the well-known “egg-box” dimers). In contrast, Zn2+ also interacts with hydroxyl groups, resulting in a less homogeneous cross-linked pectin network.
In 2016, Huynh et al. [70] investigated the binding mechanism using different cationic metals (Zn2+, Ca2+, Ba2+, Mg2+) by isothermal titration calorimetry, viscosity measurements, and molecular dynamics simulations in poly (galacturonic) acid. In this framework, a monodentate interaction meant the interaction between a carboxyl group of galacturonic acid and a cation. In contrast, a bidentate interaction means the binding of two oxygen atoms of the galacturonic acid (from two different pectin chains) and a cation. They reported that the interaction between divalent cations and poly (galacturonic) acid is monodentate for Mg2+ and Zn2+ and bidentate for Ba2+ and Ca2+. Moreover, the binding mechanism for the divalent cations, Zn2+, Ca2+, and Ba2+, can be associated with mono-complexation and point-like cross-links or related to the appearance of the dimmers formation, depending on the molar ratio between the divalent cation and galacturonic acid (R = Metal2+/Gal). This conclusion was based on the number of water molecules coordinated with the different cations. In the case of Zn2+, Ca2+, and Ba2+, the coordination with water was weaker in strength than those of Mg2+. During the cross-link, the metal ions lose one water molecule for the Zn2+ and two for Ca2+ and Ba2+, respectively. However, Mg2+ strongly interacts with water and remains weakly bound to poly(galacturonic) acid by sharing water molecules from its coordination shell with the carboxylate groups. Finally, two years later, Huynh et al. [71] studied the gelation kinetics of pectin induced from divalent cations (Zn2+, Ca2+, Ba2+, Mg2+) by viscoelastic and turbidity measurements. Their results confirmed that the cations’ diffusion kinetics was lower for Zn2+ than for Ca2+ and Ba2+. However, for the Mg2+, the gel was not formed, supporting a previous study [70].

4. Properties of Pectin That Influence the Absorption of Metal Ions

Cataldo et al. [72] prepared pectate and poly(galacturonate) calcium gel beads for mercury (II) removal. Based on a different pH value study, they found that the best pH range for Hg2+ removal was between 3 and 3.6. Celus et al. [73] studied the ability of citrus pectin to adsorb Fe2+, emphasizing two structural properties: the degree of methyl esterification (DM) and the degree of blockiness (ratio of non-methyl esterified GalA units present in blocks to the total amount of GalA units) of citrus pectin. They found that the DM and DBabs influenced pectin-Fe2+ interactions: the higher the DBabs or lower DM, the higher the Fe2+ binding capacity.
In another study, pectin micro-gel particles were also used to remove methylene blue (MB) [74]. The authors found very high absorption ratios, even with time shorter than 20 min. Additionally, LM pectin was used to clean mineral soils polluted with Cu2+ [75]. Recently, a sweet potato residue modified by high hydrostatic pressure (HHP)-assisted pectinase was prepared [76]. This material was used for Pb2+ removal. They showed that the modified sweet potato pectin exhibited more excellent adsorption performances for Pb2+ and Cu2+ than natural pectin.
LM Pectin was also mixed with other biopolymers, such as chitosan, to prepare chitosan-pectin gel beads, which were synthesized via a green method [77] to remove a collection of heavy metals (Cu2+, Cd2+, Hg2+, and Pb2+). They analyzed different variables such as the effect of pH, contact time, heavy metals concentration, temperature, and adsorption mechanism. They found that adding pectin increased the adsorption capacities, porosity, and stability of the adsorbents. The infrared analysis of the adsorbents indicates that the interaction between heavy metals and chitosan-pectin gel beads is due to the complexation with functional groups such as carboxyl, hydroxyl, amine, and amide.
Mata et al. [78] prepared sugar-beet pectin xerogels using residues of the sugar industry. Their study is in conditions of continuous biosorption and not on stationary conditions as the above presented. In this case, the adsorbent was used for metal recovery from effluents in continuous systems. They studied different experimental conditions: feed flow rate and bed height (amount of biosorbent) and found that the best conditions for Cu2+ sorption in column reactors were: 3 g of biomass, 25 mg/L metal, 2 mL/min feed flow rate, and a reverse feeding system.
It is necessary to remark that there was a substantial dispersion of results obtained in the literature in relation to the adsorption capacity of pectin-based adsorbents. This is because the adsorption capacity of pectin strongly depended on its chemical characteristics and relied on both the source on which pectin was obtained and the extraction method. For instance, an early work by Kartel et al. [79] showed that apple pectin had the highest affinity for Co2+ ions, whereas, in the case of beet pectin, the affinity was better for Cu2+ and Cd2+ ions. In contrast, citrus pectin showed a different preference for Ni2+, Zn2+, and Pb2+ ions. These different affinities depend on the structural differences of pectin. For instance, the galacturonic acid content for apple pomace is between 20 and 44% [80], whereas, for orange peel, it has been between 66 and 70% on a dried basis. In addition, apple pectin obtained from the different extraction procedures was highly methylated (from 54.5% to 79.5%) [9], but the methoxylation degree was much lower for orange peels. Thus, the source of pectin and the way it is extracted determines the structural properties of pectin, which, in turn, defines the adsorption capacity of heavy metals.
Among the different structural properties of pectin, the more critical parameter related to heavy metals binding is the methoxylation degree (DM) [2][81]. When DM is small, more carboxyl groups are available on pectin, which can interact with the metal ion, increasing the adsorption capacity. In addition, it has been reported that the pattern of methyl esterification is also crucial for determining adsorption capacities [82].
The influence of DM on the adsorption of metal ions has been studied for the adsorption of Zn2+ [83], Fe3+ [73], Ca2+, Zn2+, Fe2+, and Mg2+ [84]. The influence on other structural parameters of pectin, such as the degree of acetylation (DA), chain length, or the branched domains, have been less studied concerning the adsorption capacity.
The molecular weight of pectin also influenced the adsorption capacity. The lower the molecular weight, the higher the adsorption capacity [85]. This effect occurs because small pectin chains can reveal additional active sites that can result in a more significant electrostatic attraction to capture more heavy metal ions [86]. In addition, a longer pectin chain can form a firmer gel in which the metal ions cannot penetrate.
In addition, due to the high degree of heterogeneity regarding chemical composition and physicochemical properties such as molecular weight, the degree of esterification, dispersion, and galacturonic acid content of the pectin, predictions of the metal-binding capacity of different pectin batches are quite tricky.

References

  1. Ciriminna, R.; Chavarría-Hernández, N.; Inés Rodríguez Hernández, A.; Pagliaro, M. Pectin: A new perspective from the biorefinery standpoint. Biofuels Bioprod. Biorefining 2015, 9, 368–377.
  2. Voragen, A.G.J.; Coenen, G.-J.; Verhoef, R.P.; Schols, H.A. Pectin, a versatile polysaccharide present in plant cell walls. Struct. Chem. 2009, 20, 263–275.
  3. Gawkowska, D.; Cybulska, J.; Zdunek, A. Structure-Related Gelling of Pectins and Linking with Other Natural Compounds: A Review. Polymers 2018, 10, 762.
  4. Dranca, F.; Oroian, M. Optimization of Pectin Enzymatic Extraction from Malus domestica ‘Fălticeni’ Apple Pomace with Celluclast 1.5L. Molecules 2019, 24, 2158.
  5. Cui, J.; Zhao, C.; Feng, L.; Han, Y.; Du, H.; Xiao, H.; Zheng, J. Pectins from fruits: Relationships between extraction methods, structural characteristics, and functional properties. Trends Food Sci. Technol. 2021, 110, 39–54.
  6. Guo, X.; Han, D.; Xi, H.; Rao, L.; Liao, X.; Hu, X.; Wu, J. Extraction of pectin from navel orange peel assisted by ultra-high pressure, microwave or traditional heating: A comparison. Carbohydr. Polym. 2012, 88, 441–448.
  7. Yeoh, S.; Shi, J.; Langrish, T.A.G. Comparisons between different techniques for water-based extraction of pectin from orange peels. Desalination 2008, 218, 229–237.
  8. Ravn, H.C.; Meyer, A.S. Chelating agents improve enzymatic solubilization of pectinaceous co-processing streams. Process. Biochem. 2014, 49, 250–257.
  9. Garna, H.; Mabon, N.; Robert, C.; Cornet, C.; Nott, K.; Legros, H.; Wathelet, B.; Paquot, M. Effect of Extraction Conditions on the Yield and Purity of Apple Pomace Pectin Precipitated but Not Washed by Alcohol. J. Food Sci. 2007, 72, C001–C009.
  10. Gavahian, M.; Mathad, G.N.; Pandiselvam, R.; Lin, J.; Sun, D.-W. Emerging technologies to obtain pectin from food processing by-products: A strategy for enhancing resource efficiency. Trends Food Sci. Technol. 2021, 115, 42–54.
  11. Hu, W.; Cassard, A.-M.; Ciocan, D. Pectin in Metabolic Liver Disease. Nutrients 2023, 15, 157.
  12. Li, D.-Q.; Li, J.; Dong, H.-L.; Li, X.; Zhang, J.-Q.; Ramaswamy, S.; Xu, F. Pectin in biomedical and drug delivery applications: A review. Int. J. Biol. Macromol. 2021, 185, 49–65.
  13. Houron, C.; Ciocan, D.; Trainel, N.; Mercier-Nomé, F.; Hugot, C.; Spatz, M.; Perlemuter, G.; Cassard, A.-M. Gut Microbiota Reshaped by Pectin Treatment Improves Liver Steatosis in Obese Mice. Nutrients 2021, 13, 3725.
  14. Munarin, F.; Guerreiro, S.G.; Grellier, M.A.; Tanzi, M.C.; Barbosa, M.A.; Petrini, P.; Granja, P.L. Pectin-Based Injectable Biomaterials for Bone Tissue Engineering. Biomacromolecules 2011, 12, 568–577.
  15. Freitas, C.; Coimbra, J.; Souza, V.; Sousa, R. Structure and Applications of Pectin in Food, Biomedical, and Pharmaceutical Industry: A Review. Coatings 2021, 11, 922.
  16. Moslemi, M. Reviewing the recent advances in application of pectin for technical and health promotion purposes: From la-boratory to market. Carbohydr. Polym. 2021, 254, 117324.
  17. Reichembach, L.H.; Petkowicz, C.L.D.O. Pectins from alternative sources and uses beyond sweets and jellies: An overview. Food Hydrocoll. 2021, 118, 106824.
  18. Oliva-Moreno, E.E.; Encinas, A. Addition of Pine Rosin to Pectin bioplastic films for improved water resistance. Mater. Lett. 2021, 290, 129488.
  19. Espitia, P.J.P.; Du, W.-X.; Avena-Bustillos, R.d.J.; Soares, N.d.F.F.; McHugh, T.H. Edible films from pectin: Physical-mechanical and antimicrobial properties—A review. Food Hydrocoll. 2014, 35, 287–296.
  20. Huang, S.; Tu, Z.; Sha, X.; Hu, Y.; Chen, N.; Wang, H. Fabrication and performance evaluation of pectin–fish gelatin–resveratrol preservative films. Food Chem. 2021, 361, 129832.
  21. Makaremi, M.; Pasbakhsh, P.; Cavallaro, G.; Lazzara, G.; Aw, Y.K.; Lee, S.M.; Milioto, S. Effect of Morphology and Size of Halloysite Nanotubes on Functional Pectin Bionanocomposites for Food Packaging Applications. ACS Appl. Mater. Interfaces 2017, 9, 17476–17488.
  22. Yoshimura, T.; Sengoku, K.; Fujioka, R. Pectin-based surperabsorbent hydrogels crosslinked by some chemicals: Synthesis and characterization. Polym. Bull. 2005, 55, 123–129.
  23. Chen, B.; McClements, D.J.; Gray, D.A.; Decker, E.A. Stabilization of Soybean Oil Bodies by Enzyme (Laccase) Cross-Linking of Adsorbed Beet Pectin Coatings. J. Agric. Food Chem. 2010, 58, 9259–9265.
  24. Groult, S.; Buwalda, S.; Budtova, T. Pectin hydrogels, aerogels, cryogels and xerogels: Influence of drying on structural and release properties. Eur. Polym. J. 2021, 149, 110386.
  25. Allwyn, S.R.A. A Review on Pectin: Chemistry due to General Properties of Pectin and its Pharmaceutical Uses. Sci. Rep. 2012, 1, 550–553.
  26. Voragen, F.; Schols, H.; Visser, M. Advances in Pectin and Pectinase Research; Springer: Dordrecht, The Netherlands, 2003.
  27. Chen, R.; Ratcliffe, I.; Williams, P.A.; Luo, S.; Chen, J.; Liu, C. The influence of pH and monovalent ions on the gelation of pectin from the fruit seeds of the creeping fig plant. Food Hydrocoll. 2020, 111, 106219.
  28. Wang, H.; Wan, L.; Chen, D.; Guo, X.; Liu, F.; Pan, S. Unexpected gelation behavior of citrus pectin induced by monovalent cations under alkaline conditions. Carbohydr. Polym. 2019, 212, 51–58.
  29. Jonassen, H.; Treves, A.; Kjøniksen, A.-L.; Smistad, G.; Hiorth, M. Preparation of Ionically Cross-Linked Pectin Nanoparticles in the Presence of Chlorides of Divalent and Monovalent Cations. Biomacromolecules 2013, 14, 3523–3531.
  30. Ström, A.; Schuster, E.; Goh, S.M. Rheological characterization of acid pectin samples in the absence and presence of mon-ovalent ions. Carbohydr. Polym. 2014, 113, 336–343.
  31. Yoo, S.-H.; Fishman, M.L.; Savary, B.J.; Hotchkiss, A.T. Monovalent Salt-Induced Gelation of Enzymatically Deesterified Pectin. J. Agric. Food Chem. 2003, 51, 7410–7417.
  32. Wehr, J.B.; Blamey, F.P.C.; Hanna, J.V.; Kopittke, P.M.; Kerven, G.L.; Menzies, N.W. Hydrolysis and Speciation of Al Bound to Pectin and Plant Cell Wall Material and Its Reaction with the Dye Chrome Azurol S. J. Agric. Food Chem. 2010, 58, 5553–5560.
  33. McKenna, B.A.; Nicholson, T.M.; Wehr, J.B.; Menzies, N. Effects of Ca, Cu, Al and La on pectin gel strength: Implications for plant cell walls. Carbohydr. Res. 2010, 345, 1174–1179.
  34. Wu, X.; Sun, H.; Qin, Z.; Che, P.; Yi, X.; Yu, Q.; Zhang, H.; Sun, X.; Yao, F.; Li, J. Fully physically crosslinked pectin-based hydrogel with high stretchability and toughness for biomedical application. Int. J. Biol. Macromol. 2020, 149, 707–716.
  35. Pellerin, P.; O’Neill, M.A. The interaction of the pectic polysaccharide Rhamnogalacturonan II with heavy metals and lan-thanides in wines and fruit juices. Analusis 1998, 26, 32–36.
  36. Grant, G.T.; Morris, E.R.; Rees, D.A.; Smith, P.J.; Thom, D. Biological interactions between polysaccharides and divalent cations: The egg-box model. FEBS Lett. 1973, 32, 195–198.
  37. Chaichi, M.; Badii, F.; Mohammadi, A.; Hashemi, M. Water resistance and mechanical properties of low methoxy-pectin nanocomposite film responses to interactions of Ca2+ ions and glycerol concentrations as crosslinking agents. Food Chem. 2019, 293, 429–437.
  38. Li, D.-Q.; Wang, J.; Guo, Z.-G.; Li, J.; Shuai, J. Pectin gels cross-linked by Ca 2+: An efficient material for methylene blue removal. J. Mol. Liq. 2017, 238, 36–42.
  39. Huang, T.; Qin, Y.; Li, M.; Gao, S.; Shen, C. Preparation and characterization of deacetylated konjac glucomannan/pectin composite films crosslinked with calcium hydroxide. J. Polym. Res. 2022, 29, 238.
  40. John, J.; Deshpande, A.P.; Varughese, S. Morphology control and ionic crosslinking of pectin domains to enhance the toughness of solvent cast PVA/pectin blends. J. Appl. Polym. Sci. 2021, 138, 50360.
  41. Fraeye, I.; Colle, I.; Vandevenne, E.; Duvetter, T.; Van Buggenhout, S.; Moldenaers, P.; Van Loey, A.; Hendrickx, M. Influence of pectin structure on texture of pectin–calcium gels. Innov. Food Sci. Emerg. Technol. 2010, 11, 401–409.
  42. Ngouémazong, D.E.; Kabuye, G.; Fraeye, I.; Cardinaels, R.; Van Loey, A.; Moldenaers, P.; Hendrickx, M. Effect of debranching on the rheological properties of Ca2+–pectin gels. Food Hydrocoll. 2012, 26, 44–53.
  43. Yang, X.; Nisar, T.; Liang, D.; Hou, Y.; Sun, L.; Guo, Y. Low methoxyl pectin gelation under alkaline conditions and its rhe-ological properties: Using NaOH as a pH regulator. Food Hydrocoll. 2018, 79, 560–571.
  44. Capel, F.; Nicolai, T.; Durand, D.; Boulenguer, P.; Langendorff, V. Calcium and acid induced gelation of (amidated) low methoxyl pectin. Food Hydrocoll. 2006, 20, 901–907.
  45. Yuliarti, O.; Hoon, A.L.S.; Chong, S.Y. Influence of pH, pectin and Ca concentration on gelation properties of low-methoxyl pectin extracted from Cyclea barbata Miers. Food Struct. 2017, 11, 16–23.
  46. Gołębiowski, A.; Kowalkowski, T.; Buszewski, B. Molecular parameters of low methoxylated pectin affected by gelation with copper and cadmium cations. Bioact. Carbohydrates Diet. Fibre 2020, 21, 100211.
  47. Tkalec, G.; Knez, Ž.; Novak, Z. PH sensitive mesoporous materials for immediate or controlled release of NSAID. Microporous Mesoporous Mater. 2016, 224, 190–200.
  48. Kohila rani, K.; Liu, Y.-X.; Devasenathipathy, R.; Yang, C.; Wang, S.-F. Simple preparation of gold nanoparticle-decorated copper cross-linked pectin for the sensitive determination of hydrogen peroxide. Ionics 2019, 25, 309–317.
  49. Nešić, A.; Onjia, A.; Davidović, S.; Dimitrijević, S.; Errico, M.E.; Santagata, G.; Malinconico, M. Design of pectin-sodium alginate based films for potential healthcare application: Study of chemico-physical interactions between the components of films and assessment of their antimicrobial activity. Carbohydr. Polym. 2017, 157, 981–990.
  50. Das, S.; Ng, K.-Y.; Ho, P.C. Design of a pectin-based microparticle formulation using zinc ions as the cross-linking agent and glutaraldehyde as the hardening agent for colonic-specific delivery of resveratrol: In vitro and in vivo evaluations. J. Drug Target. 2011, 19, 446–457.
  51. Ullah, K.; Sohail, M.; Buabeid, M.A.; Murtaza, G.; Ullah, A.; Rashid, H.; Khan, M.A.; Khan, S.A. Pectin-based (LA-co-MAA) semi-IPNS as a potential biomaterial for colonic delivery of oxaliplatin. Int. J. Pharm. 2019, 569, 118557.
  52. Li, F.T.; Yang, H.; Zhao, Y.; Xu, R. Novel modified pectin for heavy metal adsorption. Chin. Chem. Lett. 2007, 18, 325–328.
  53. Chen, S.; Cui, S.; Zhang, H.; Pei, X.; Hu, J.; Zhou, Y.; Liu, Y. Cross-Linked Pectin Nanofibers with Enhanced Cell Adhesion. Biomacromolecules 2018, 19, 490–498.
  54. Jung, J.; Wicker, L. Laccase mediated conjugation of sugar beet pectin and the effect on emulsion stability. Food Hydrocoll. 2012, 28, 168–173.
  55. Kuo, C.K.; Ma, P.X. Ionically crosslinked alginate hydrogels as scaffolds for tissue engineering: Part 1. Structure, gelation rate and mechanical properties. Biomaterials 2001, 22, 511–521.
  56. Morris, E.R.; Rees, D.A.; Thom, D.; Boyd, J. Chiroptical and stoichiometric evidence of a specific, primary dimerisation process in alginate gelation. Carbohydr. Res. 1978, 66, 145–154.
  57. Walkinshaw, M.; Arnott, S. Conformations and interactions of pectins: II. Models for junction zones in pectinic acid and calcium pectate gels. J. Mol. Biol. 1981, 153, 1075–1085.
  58. Powell, D.A.; Morris, E.R.; Gidley, M.J.; Rees, D.A. Conformations and interactions of pectins: II. Influence of residue se-quence on chain association in calcium pectate gels. J. Mol. Biol. 1982, 155, 517–531.
  59. Kohn, R. Binding of divalent cations to oligomeric fragments of pectin. Carbohydr. Res. 1987, 160, 343–353.
  60. Walter, R.H. The Chemistry and Technology of Pectin; Academic Press: London, UK, 1991.
  61. Braccini, I.; Pérez, S. Molecular Basis of Ca2+-Induced Gelation in Alginates and Pectins: The Egg-Box Model Revisited. Biomacromolecules 2001, 2, 1089–1096.
  62. Li, L.; Fang, Y.; Vreeker, R.; Appelqvist, I.; Mendes, E. Reexamining the Egg-Box Model in Calcium−Alginate Gels with X-ray Diffraction. Biomacromolecules 2007, 8, 464–468.
  63. Mackie, W. Conformations of crystalline alginic acids and their salts. Biochem. J. 1971, 125, 89P.
  64. Fang, Y.; Al-Assaf, S.; Phillips, G.O.; Nishinari, K.; Funami, T.; Williams, P.A.; Li, L. Multiple Steps and Critical Behaviors of the Binding of Calcium to Alginate. J. Phys. Chem. B 2007, 111, 2456–2462.
  65. Fang, Y.; Al-Assaf, S.; Phillips, G.O.; Nishinari, K.; Funami, T.; Williams, P.A. Binding behavior of calcium to polyuronates: Comparison of pectin with alginate. Carbohydr. Polym. 2008, 72, 334–341.
  66. Gohil, R.M. Synergistic blends of natural polymers, pectin and sodium alginate. J. Appl. Polym. Sci. 2011, 120, 2324–2336.
  67. Ventura, I.; Jammal, J.; Bianco-Peled, H. Insights into the nanostructure of low-methoxyl pectin–calcium gels. Carbohydr. Polym. 2013, 97, 650–658.
  68. Wang, H.; Fei, S.; Wang, Y.; Zan, L.; Zhu, J. Comparative study on the self-assembly of pectin and alginate molecules regulated by calcium ions investigated by atomic force microscopy. Carbohydr. Polym. 2019, 231, 115673.
  69. Assifaoui, A.; Lerbret, A.; Uyen, H.T.D.; Neiers, F.; Chambin, O.; Loupiac, C.; Cousin, F. Structural behaviour differences in low methoxy pectin solutions in the presence of divalent cations (Ca2+ and Zn2+): A process driven by the binding mechanism of the cation with the galacturonate unit. Soft Matter 2014, 11, 551–560.
  70. Huynh, U.T.D.; Lerbret, A.; Neiers, F.; Chambin, O.; Assifaoui, A. Binding of Divalent Cations to Polygalacturonate: A Mechanism Driven by the Hydration Water. J. Phys. Chem. B 2016, 120, 1021–1032.
  71. Huynh, U.T.D.; Chambin, O.; du Poset, A.M.; Assifaoui, A. Insights into gelation kinetics and gel front migration in cation-induced polysaccharide hydrogels by viscoelastic and turbidity measurements: Effect of the nature of divalent cations. Carbohydr. Polym. 2018, 190, 121–128.
  72. Cataldo, S.; Gianguzza, A.; Pettignano, A.; Villaescusa, I. Mercury(II) removal from aqueous solution by sorption onto alginate, pectate and polygalacturonate calcium gel beads. A kinetic and speciation based equilibrium study. React. Funct. Polym. 2013, 73, 207–217.
  73. Celus, M.; Kyomugasho, C.; Kermani, Z.J.; Roggen, K.; Van Loey, A.M.; Grauwet, T.; Hendrickx, M.E. Fe2+ adsorption on citrus pectin is influenced by the degree and pattern of methylesterification. Food Hydrocoll. 2017, 73, 101–109.
  74. Yu, L.-L.; Jiang, L.-N.; Wang, S.-Y.; Sun, M.-M.; Li, D.-Q.; Du, G.-M. Pectin microgel particles as high adsorption rate material for methylene blue: Performance, equilibrium, kinetic, mechanism and regeneration studies. Int. J. Biol. Macromol. 2018, 112, 383–389.
  75. Wang, R.-H.; Zhu, X.-F.; Qian, W.; Yu, Y.-C.; Xu, R.-K. Effect of pectin on adsorption of Cu(II) by two variable-charge soils from southern China. Environ. Sci. Pollut. Res. 2015, 22, 19687–19694.
  76. Arachchige, M.P.M.; Mu, T.; Ma, M. Effect of high hydrostatic pressure-assisted pectinase modification on the Pb2+ adsorption capacity of pectin isolated from sweet potato residue. Chemosphere 2021, 262, 128102.
  77. Shao, Z.; Lu, J.; Ding, J.; Fan, F.; Sun, X.; Li, P.; Fang, Y.; Hu, Q. Novel green chitosan-pectin gel beads for the removal of Cu(II), Cd(II), Hg(II) and Pb(II) from aqueous solution. Int. J. Biol. Macromol. 2021, 176, 217–225.
  78. Mata, Y.; Blázquez, M.; Ballester, A.; González, F.; Muñoz, J. Optimization of the continuous biosorption of copper with sugar-beet pectin gels. J. Environ. Manag. 2009, 90, 1737–1743.
  79. Kartel, M.; Kupchik, L.A.; Veisov, B.K. Evaluation of pectin binding of heavy metal ions in aqueous solutions. Chemosphere 1999, 38, 2591–2596.
  80. Wang, X.; Chen, Q.; Lü, X. Pectin extracted from apple pomace and citrus peel by subcritical water. Food Hydrocoll. 2014, 38, 129–137.
  81. Ralet, M.-C.; Crépeau, M.-J.; Buchholt, H.-C.; Thibault, J.-F. Polyelectrolyte behaviour and calcium binding properties of sugar beet pectins differing in their degrees of methylation and acetylation. Biochem. Eng. J. 2003, 16, 191–201.
  82. Löfgren, C.; Guillotin, S.; Evenbratt, H.; Schols, H.; Hermansson, A.-M. Effects of Calcium, pH, and Blockiness on Kinetic Rheological Behavior and Microstructure of HM Pectin Gels. Biomacromolecules 2005, 6, 646–652.
  83. Khotimchenko, M.; Kovalev, V.; Khotimchenko, Y. Equilibrium studies of sorption of lead(II) ions by different pectin com-pounds. J. Hazard. Mater. 2007, 149, 693–699.
  84. Kyomugasho, C.; Gwala, S.; Christiaens, S.; Jamsazzadeh Kermani, Z.; Van Loey, A.M.; Grauwet, T.; Hendrickx, M.E. Pectin nanostructure influences pectin-cation interactions and in vitro-bioaccessibility of Ca2+, Zn2+, Fe2+ and Mg2+-ions in model systems. Food Hydrocoll. 2017, 62, 299–310.
  85. Cao, L.; Lu, W.; Mata, A.; Nishinari, K.; Fang, Y. Egg-box model-based gelation of alginate and pectin: A review. Carbohydr. Polym. 2020, 242, 116389.
  86. Khozhaenko, E.; Kovalev, V.; Podkorytova, E.; Khotimchenko, M. Removal of the metal ions from aqueous solutions by na-noscaled low molecular pectin isolated from seagrass Phyllospadix iwatensis. Sci. Total Environ. 2016, 565, 913–921.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , ,
View Times: 747
Entry Collection: Wastewater Treatment
Revisions: 2 times (View History)
Update Date: 22 Mar 2023
1000/1000