Next Article in Journal
Calcium-Based Antimicrobial Peptide Compounds Attenuate DNFB-Induced Atopic Dermatitis-Like Skin Lesions via Th-Cells in BALB/c Mice
Next Article in Special Issue
Physiological and Molecular Mechanisms of Plant Responses to Copper Stress
Previous Article in Journal
Functional and Expressional Analyses Reveal the Distinct Role of Complement Factor I in Regulating Complement System Activation during GCRV Infection in Ctenopharyngodon idella
Previous Article in Special Issue
A Hormetic Spatiotemporal Photosystem II Response Mechanism of Salvia to Excess Zinc Exposure
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Metallophenolomics: A Novel Integrated Approach to Study Complexation of Plant Phenolics with Metal/Metalloid Ions

1
Research Institute of Biology, Oles Honchar Dnipro National University, 72 Gagarin Avenue, 49010 Dnipro, Ukraine
2
Department of Agriculture, Food and Environment, University of Pisa, Via del Borghetto, 80I-56124 Pisa, Italy
3
Ukrainian Association for Haemophilia and Haemostasis “Factor D”, Topola-3, 20/2/81, 49041 Dnipro, Ukraine
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(19), 11370; https://doi.org/10.3390/ijms231911370
Submission received: 1 September 2022 / Revised: 21 September 2022 / Accepted: 22 September 2022 / Published: 26 September 2022
(This article belongs to the Special Issue Molecular Mechanisms of Metal Toxicity and Plant Tolerance)

Abstract

:
Plant adaptive strategies have been shaped during evolutionary development in the constant interaction with a plethora of environmental factors, including the presence of metals/metalloids in the environment. Among adaptive reactions against either the excess of trace elements or toxic doses of non-essential elements, their complexation with molecular endogenous ligands, including phenolics, has received increasing attention. Currently, the complexation of phenolics with metal(loid)s is a topic of intensive studies in different scientific fields. In spite of the numerous studies on their chelating capacity, the systemic analysis of phenolics as plant ligands has not been performed yet. Such a systematizing can be performed based on the modern approach of metallomics as an integral biometal science, which in turn has been differentiated into subgroups according to the nature of the bioligands. In this regard, the present review summarizes phenolics–metal(loid)s’ interactions using the metallomic approach. Experimental results on the chelating activity of representative compounds from different phenolic subgroups in vitro and in vivo are systematized. General properties of phenolic ligands and specific properties of anthocyanins are revealed. The novel concept of metallophenolomics is proposed, as a ligand-oriented subgroup of metallomics, which is an integrated approach to study phenolics–metal(loid)s’ complexations. The research subjects of metallophenolomics are outlined according to the methodology of metallomic studies, including mission-oriented biometal sciences (environmental sciences, food sciences and nutrition, medicine, cosmetology, coloration technologies, chemical sciences, material sciences, solar cell sciences). Metallophenolomics opens new prospects to unite multidisciplinary investigations of phenolic–metal(loid) interactions.

1. Introduction

The life processes of plants have evolved in coordination with environmental factors. In addition, intensified anthropogenic load on ecosystems has led to increasing levels of chemical contamination and resulted in the emergence of new pollutants, namely xenobiotics. To understand the peculiarities of plant–environment interactions, it is essential to take into account the environmental and ecological aspects of the problem [1]. Hazardous metals and metalloids are among the major and widespread pollutants due to their high toxicity to the biosphere and the amplitude of their contamination in the natural environment [2]. Mining and metal extraction, fossil fuel combustion, agricultural application of fertilizers, sewage sludge, metal-containing pesticides, wastewater irrigation, and atmospheric deposition are the main anthropic sources of metal(loid)s [3].
A special feature of plants regarding their relations to metal(loid)s is that a certain amount of trace metals is necessary for a number of biologically essential processes (metalloenzymes, mineral nutrition, photosynthesis, prooxidant/antioxidant systems, etc.), whilst both the overdoses of essential and toxic doses of non-essential elements negatively affect plant metabolism. Therefore, to optimize those processes, plants have evolved multiple regulatory and defence mechanisms to counteract metal(loid) toxicity [4,5,6].
Among others, plants detoxify metal(loid)s via their biotransformation into metabolically inactive compounds [7,8]. Among biotransformation reactions, the in vivo chelating of metal ions is pivotal, which is accomplished by endogenous chelators: glutathione (GSH), phytochelatins (PCs), metallothioneins (MTs), organic acids, nicotinamine, amino acids [6,9,10]. An important feature of those bioligands is their capacity to bind various metals [10]. Such a property of binding both essential and non-essential metal(loid)s has been confirmed in vivo for flavonoid pigments of anthocyanins (ACNs) [11,12,13]. Those results allow postulating a hypothesis about the participation of phenolic compounds (PCs) in metal(loid) detoxification in plants [12]. However, a comprehensive systemic analysis of the role of PCs as endogenous chelators in plant metal tolerance has not yet been performed.
It is noteworthy that the current level of metal(loid) tolerance in plants is characterized by the extensive use of the integrated “omics” approach [14,15,16,17,18]. The “omics” approach is aimed at the studying of the organism as a holistic system, based on the integrative analysis of and interrelations among major biological processes [19]. The “omics” research object related to the behaviour of metals in living organism is referred to as the “metallome” [20] and the corresponding research field as “metallomics” [21]. (These terms will be evaluated in detail in Section 2.1.)
In recent years, the key significance of PCs was confirmed as pivotal and versatile plant defensive compounds against abiotic stresses, including metal(loid) tolerance [5,22,23,24]. Systematizing extensive experimental data resulted in the hypothesis of a universal dominant tendency to increased accumulation of PCs as components of the antioxidant defence system, which ensures the balance between the production and detoxification of reactive oxygen species (ROSs) under metal(loid) exposure [25]. Beside their ROS scavenging prerogative, other possible roles of those secondary metabolites in the metallomics context remain a poorly investigated issue.
This review outlines (i) the basic concepts of metallomic studies, (ii) their differentiation into subgroups, (iii) the chemistry of the complexation of phenolic ligands with metal(loid)s, and (iv) the introduction to the novel integrated approach of metallophenolomics to study metal(loid)–phenolic interrelations.

2. Metallomics as a Scientific Approach

The need for the systemic evaluation of metal behaviours in living organisms has led to the development of new approaches for studying this problem. For plant–metal(loid) interactions, the most relevant is the concept of metallomics, which is reviewed in detail in the following subsections.

2.1. Basic Concepts of Metallomics

The new scientific field, namely “metallomics”, as an integrated biometal science was introduced by Haraguchi [24]; the historical aspect of the origin and development of metal-related omics approaches was systematized by the author in the review [26]. The definitions of the key terms and basic concepts of this scientific approach were described in [27].
The term “metallome” can be described as the entirety of metal and metalloid species present in a biological system, defined as their identity and/or quantity [27]. Firstly, the metallome is the distribution of metal ions in a cellular compartment. Secondly, the metallome is related to the definition of the total element concentrations, metallocomplexes with different ligands, or all the species of a certain element (in their free forms or included in eobiotic or xenobiotic molecules). To define the metal-bound compounds in living organisms, in addition to the term “bioligands”, the terms “metallobiomolecules” [28], “chelating agents” [29], and “chelators” [9] are used.
The entirety of the scientific branches studying the metallome is included in the broader definition of “metallomics”. The main characteristics of those studies are: (i) the focus on metals or metalloids in a biological context; (ii) the statistical, functional, or structural link between the set of the element concentration or element speciation and the genome; (iii) a systematic or comprehensive approach [27]. It should be noted that the main feature of metallomics is not only the identification of metals, but also the determination of their roles and effects in biological processes. A peculiarity of metallomics is the studies of the distribution of metal ions, their quantities, the multivariance of their interaction with different bioligands, and the spatial and temporal characteristics of this process during the development of the living organism, as well as the impact of genetic variability. The complicated nature of this science was characterized by [29] as “metametallomics”.
Metallomics is the modern scientific branch, which has demonstrated an active development in the recent decade; therefore, this concept has received much attention and was extensively evaluated by many authors in a number of reviews and monographs [29,30,31,32,33,34,35]. Metallomics is an interdisciplinary topic and merges different spheres of research to build a global and systemic understanding of metal-assisted functions in biological systems; it includes various scientific fields and research subjects [26]. For the systemic analysis of the functions of biological systems, researchers consider integrating metallomics with different other omics (metabolomics, genomics, transcriptomics) [36].
To characterize completely the metal-related omics approaches, it should be noted that, in addition to metallomics, other scientific directions have been proposed, which have related terminological definitions. Staring from 2003, when the concept of the ionome was firstly introduced by Lahner et al. [37], ionomics has been developed in parallel with metallomics, the science involving “quantitative and simultaneous measurements of the elemental composition of living organisms, and changes in this composition in response to physiological stimuli, developmental state, and genetic modifications” [38]. The further development of ionomics and its connection with other omics has recently been reviewed [18,39,40].
To expand the area of the omics approach and encompass all chemical elements present in living organisms, including non-metals, elementomics was introduced, as the “study of elements of interest and element species, and their interactions, transformations, and functions in biological systems” [41]. Of note, the omics study of a specific global effect of a particular metal is also defined; for example, As-induced stress in plants is defined as arsenomics, an integrated approach associated with transcriptomics, proteomics, and metabolomics [42,43,44]. The term “metallometabolomics” was introduced to characterize the whole entirety of metallo-metabolites or to identify some of the dominant metal-complexing metabolites, e.g., in metal-hyperaccumulating plants [31]. Zhang et al. [45] used the term “elemental metabolomics” for the “quantification and characterization of total concentration of chemical elements in biological samples and monitoring of their changes.” Another aspect in metal-related studies is defined as “metallome homeostasis”—the understanding of “how the individual metal homeostasis systems overlap and interact so that all required essential metals are obtained and routed to the correct locations” [46].

2.2. Differentiation of Metallomic Studies

The extensive development of the metallomic approaches in different scientific fields and research subjects has resulted in the accumulation of a vast volume of experimental data. Due to the need to systematize the available literature accumulating on the matter in different directions, a problem arises from the division of the separated subgroups in metallomic studies [47]. Metallomics could be divided into subgroups in several ways, depending on the criteria for classification, with the focus on various aspects of the problem.
Specific experimental criteria: Szpunar [30] suggested to distinguish qualitative metallomics, which is based on identifying individual metal species, and quantitative metallomics, which deals with metal concentrations. For the monitoring of metallome changes in time and under environmental factors, the term “comparative metallomics” can be used [30]. Structural metallomics studies the metal coordination environment in biological systems [29]. Functional metallomics investigates the role of metal ions in the functions of proteins (catalytic, structural, regulatory) and, in particular, enzymes (oxidoreductases, transferases, hydrolases, liases, isomerases, ligases) [29].
Scientific field: For plant sciences, the following directions/subgroups have been proposed. In general, for plant metallomics, the term phytometallomics is used [48]. The metallomic study of plants is viewed as a specific aspect of the directions with a broader scope, which have been proposed as independent interdisciplines. Thus, to study the problems related to metal(loid)s in agricultural science, the agrometallomics concept was created [49]. The environmental branch of metallomic study is termed environmental metallomics [50,51] or envirometallomics [52].
Form of metal(loid) species: Some subgroups within metallomics can be distinguished based on the different natures of various metal species. Thus, studies devoted to the behaviour of different metal isotopes are defined as isotope metallomics [53,54]. The metabolism and behaviours of radioactive elements are related to radiometallomics [55]. The extensive development of nanotechnology in recent years has resulted in the emergence of a new branch of metallomics—nanometallomics—which aims at quantitation, distribution, structural changes, metabolism, the elucidation of reactions and mechanisms of metal-related nanomaterials in biological systems, and specific nano-scaled metal(loid)-assisted function science in different fields [56]. Wang et al. [57] extended this approach to incorporate all nanomaterials and metal-biomolecular homeostasis processes and proposed comparative nanometallomics as a new tool for nanosafety evaluations.
Nature of chemical elements: Depending on the given element, the specific metallome subgroups were outlined, such as for iron [58,59], nickel [60], copper [61,62], zinc [63], and manganese [64].
Structure of bioligand: The best-studied subgroup within metallomics is metalloproteomics, which studies bioligands of a proteic nature [30,50,65,66,67]. Depending on the type of metal(loid) associated with the proteins, specific subgroups can be distinguished within metalloproteomics, devoted to better understanding the biological role of specific element, such as selenoproteomics for selenium [65]. The subgroup of metallomics that globally studies the thiol peptides and their metal complexes is defined as metallothiolomics [68,69]. Codd [70] introduced the term metalloglycomics for studying competitive metal-carbohydrate binding. This approach is associated with the search for new metallodrugs in bioinorganic chemistry for the systematic study of the interactions between metal ions and coordination compounds with oligosaccharides [71,72]. Currently, however, taking into account the modern data on oligosaccharides’ participation as high-molecular weight ligands in metal homeostasis in plants [73], metalloglycomics should be considered as a subgroup of metallomics.
The subgroup of metallomic studies dealing with the interaction between metals and lipids is designated as metallolipidomics [74].
Thus, the current development of metallomics is characterized by a high differentiation of this approach. However, the differentiation by the bioligand structure does not encompass a specific branch for plant chelators of a phenolic nature yet.

2.3. Potential Role of Phenolic Chelators in Plant Metallomic Studies

To establish the potential role of phenolic compounds (PCs), two important concepts of the metallomic approach must be considered. Firstly, according to this approach, the studies of the metal(loid)s’ behaviour in plants include the following aspects: (i) mobilization of low-soluble metals from soil; (ii) translocation within the plant; (iii) sequestration of metal ions in the cytosol or in cellular compartments [31]. In this regard, the interest from the researchers in low-molecular-weight metal(loid)-containing metabolites is increasing, due to the following reasons: (i) the uptake and bioavailability of essential elements, in particular Fe and Zn, are crucial for optimal elemental content in fruit and vegetables for human consumption; (ii) the plant’s ability to accumulate metals is the basis for phytoremediation technologies and for the screening of hyperaccumulators, which are able to accumulate high levels of elements from the environment; (iii) plant tolerance mechanisms toward the stress effects of metal(oid)s include the induction of endogenous chelators [31]. However, phenolics as metallo-molecules have not been considered yet in the simplified model of biological systems within the context of omics science [26]. Secondly, the note by Lobinsky et al. [27] is of crucial importance: “the description of metallome can never be complete”, which results from the multivariance of the process of metal complex formation with already-established or potential bioligands. Such an explication opens the prospect for the further development of the metallomics concept, through the investigation of metal binding properties in the metabolites, which previously had not been considered in the context of the metallome, e.g., PCs.

3. Plant Phenolics as Ligands for Metal(loid)s

In our opinion, the systematization of the available information on the chelating capacity of plant PCs should be performed in two consecutive stages. In the first stage, it is necessary to analyse the binding of PCs with metal(loid)s’ ions in vitro to establish the structure of the metallocomplexes formed and the key criteria of such a binding. In the second stage, using the identified criteria of binding, it is possible to systematize the experimental results about the PC’s chelation with metal(loid) ions (Men+) in plants in vivo.

3.1. Complexing In Vitro

In the studies of PC–Men+ chelation, two main directions can be distinguished:
(1)
Evaluation of the complexation of individual PCs with Men+, based on the features of the ligands, which are modified due to chelation;
(2)
Assessment of metal chelating ability toward PCs and plant extracts based on the alterations in the absorption of metallochromic indicators.

3.1.1. Individual Phenolic Compounds

Various aspects related to the synthesis, identification of the structure, biological activity, and application of PC–Men+ complexes have been systematized in numerous reviews [75,76,77,78,79,80,81,82]. However, some aspects of this problem remain unclear due to the scarce attention given to the involvement of PCs in the modulation of the metallome
In this regard, we analysed the available data on the ability of natural phenolic metabolites to form metallocomplexes or identifying the binding of individual compounds to Men+ in vitro in order to answer the following questions:
(1)
How are metal binding properties manifested for the natural compounds from different PC subclasses, which are formed in the process of plant phenolic metabolism?
(2)
Which structural fragments of PCs are crucial for the complexation?
(3)
Can PCs be considered as universal ligands for multiple Men+?
We systematized available experimental data, and the main results are presented in Table 1 for individual representatives of various plant PC subgroups. The structural formulas of the ligands are shown in Figure 1 with their division into separate subgroups (phenolic acids 1–12, coumarins 14–16, chalcones 17, dihydrochalcones 18, flavanones 19–23, flavanonols 24, 25, flavonols 26–37, flavan-3-ols 38–43, flavones 44–51, isoflavones 52–54, anthocyanidins 55–57, xanthonoids 58, stilbenes 59, curcuminoids 60, lignans 61, flavonolignans 62, lignins, tannins 63–65). For some flavonoids, the data on complexation are combined with their derivatives. The number of Men+ ions, for which the formation of metal complexes is confirmed, is presented regardless of the compounds with different stoichiometric ligand:Men+ ratios, or for one ligand with different Men+ ions (heterometallic complexes), or for one Men+ with different ligands (mixed complexes). For some ligands, radiolabelled complexes are included. Chemical elements, for which the complexation with PC ligands has been established, are provided in Figure 2. It should be noted that we analysed only the data on metal complexes with natural PCs; currently, however, numerous studies are being performed on synthetic ligands of a phenolic nature, obtained by structural modification of the binding sites of natural PCs in order to create new effective biologically active substances [77].
Phenolic acids, as the first structural subgroup of the plant PC metabolic pathway, could be divided into hydroxybenzoic and hydroxycinnamic acids depending on the direction of their biosynthesis [186]. Among natural hydroxybenzoic acids, the complexes with Men+ have been identified for protocatechuic acid 1, vanillic acid 2, gallic acid 3, and syringic acid 4. Protocatechuic acid 1, depending on the pH, coordinates with Al(III) and U(VI) ions via the carboxyl group or the ortho-dihydroxyl group [83,84]. For vanillic acid 2, the complexation with 17 Men+ ions has been identified (Table 1). For gallic acid 3, the coordination with Men+ can involve the carboxylate and neighbouring phenolic hydroxyl groups [91]. The largest number of complexes with Men+ (22 ions) among hydroxybenzoic acids was identified for syringic acid 4.
Transformation of cinnamic acid 5 in the shikimate pathway results in the formation of different hydroxycinnamic acids (p-coumaric acid 6, caffeic acid 7, ferulic acid 8, isoferulic acid 9, sinapic acid 10, chlorogenic acid 11, rosmarinic acid 12, chicoric acid 13) [186]. Cinnamic acid 5 forms complexes with 19 Men+ ions using its carboxyl group. In the binding of p-coumaric acid 6 with Men+, its hydroxyl group could additionally be involved. In caffeic acid 7, its o-dihydroxyl group as an additional chelating site increases the ability of this molecule for complexation. The ability to form metallocomplexes has been confirmed for their methoxy derivatives (ferulic acid 8, isoferulic acid 9, sinapic acid 10). The esters of caffeic acid with quinic acid (chlorogenic acid 11), dihydroxyphenyl-lactic acid (rosmarinic acid 12), and tartaric acid (chicoric acid 13) retain the capacity for complexation with Men+.
For the following subgroups, the complexation with Men+ has been exemplified by their representative compounds: coumarins—coumarin 14, umbellipherone 15, daphnetin 16; chalcones—butein 17; dihydrochalcones—phloretin 18 (Table 1, Figure 1).
Flavanones form metallocomplexes in the form of both aglycons (naringenin 19, eriodictyol 21, hesperitin 22) and glycosides (naringin 20, hesperidin 23). For flavanonols, the metal chelating capacity has been confirmed for their derivatives with catechol (taxifolin 24) and gallic (dihydromyricetin 25) fragments.
Among the most-studied PC bioligands are flavonols (kaempferol 26, quercetin 27, rutin 28, quercitrin 29, isoquercitrin 30, isorhamnetin 31, tamarixetin 32, fisetin 33, morin 34, myricetin 35, myricitrin 36, galangin 37) (Table 1, Figure 1). It is noteworthy that the greatest amount of coordinated metals (43 different Men+ ions) has been identified for quercetin 27 and its glycosides (rutin 28, quercitrin 29, isoquercitrin 30). This pronounced capacity of quercetin to chelate metals is associated with its structural features, which determine the possibility of different variants for the interaction with Men+. Thus, the quercetin molecule contains three potential binding sites (Figure 3): (1) between the 3-hydroxy and 4-carbonyl groups in the C ring; (2) between the 5-hydroxy (in A ring) and 4-carbonyl groups (in the C ring); (3) between the 3’- and 4’-hydroxy groups in the B ring [78].
Complexation of flavan-3-ols with Men+ is carried out by catechol and gallic binding sites ((+)-catechin 38, its stereoisomer (-)-epicatechin 39, (+)-epigallocatechin 40, esters with gallic acid–(-)-epicatechin 3-gallate 41, (-)-epigallocatechin 3-gallate 42). In theaflavin 43, Men+ binding may also involve its tropolone moiety [155].
For flavones (primuletin 44, chrysin 45, apigenin 46, luteolin 47, tricetin 48, baicalein 49, baicalin 50, acacetin 51) without 3-hydroxy groups in the C ring, the complexation with Men+ may involve the binding sites between 5-hydroxy (in A ring) and 4-carbonyl (in the C ring) or the catechol and gallic moieties. In this subgroup, the greatest number of metallocomplexes was identified for chrysin 45 and luteolin 47 (each binds 10 various Men+ ions).
Isoflavone ligands are represented by daidzein 52, genistein 53, and its O-methylated derivative biochanin A 54.
Metal chelating capacity has been demonstrated for anthocyanins and their glycosides with two or three hydroxyl groups in the B ring: cyanidin 55, delphinidin 56, petunidin 57. In contrast to other flavonoids, a specific peculiarity of ACNs is a pH-dependent dynamic equilibrium of aqueous solutions between several structural forms, which are capable of Men+ binding [13]. Among these ligands, the greatest number of metal complexes was identified for cyanidin 55 and its glycosides (27 Men+, in cationic and anionic forms).
For xanthonoids, metallocomplex formation was exemplified by mangiferin 58 (glucosylxanthone) and for stilbenes by resveratrol 59.
Among curcuminoids, the most comprehensively studied ligand is curcumin 60, which may bind 28 various Men+ due to its capacity of keto-enol tautomerism.
The ability of lignans for complexation has been confirmed for secoisolariciresinol diglucoside 61 and of flavonolignans for silibinin 62 (10 Men+ ions each).
The metal binding capacity of lignin as a polymeric phenol was studied for a ligno-cellulosic substrate with Mn(II), Cu(II), and Fe(III) ions (Merdy et al., 2003).
The presence of a great number of hydroxy groups in the structure of tannins (oligomeric and polymeric phenols) determines their high capacity for complexation with Men+. This fact has been established for their different forms: condensed tannins (proanthocyanidins), oenothein B 63 (dimeric macrocyclic ellagitannin), ellagic acid 64, tannic acid 65. The latter is one of the most-studied PC ligands (21 Men+).
Thus, our attempt at systematizing the available experimental results revealed that metallocomplexes can be formed by numerous representative ligands from 18 subgroups of plant PCs, and they are capable of binding 69 different Men+ ions (63 chemical elements) in total (Figure 2).

3.1.2. Metal Chelating Ability

The metal chelating ability is recognized as a generally accepted integrated indicator of the complexing capacity of PCs; it is used as one of the indicators in antioxidant assays [187]. The main aspects of this approach were summarized in the reviews [130,188]. The approach is based on the ability of selected metallochromic indicators to form complexes with Men+, which absorb light in the visible wavelengths range. Upon the addition of the tested PC ligand, competitive binding with Men+ occurs, with a subsequent decrease in the absorption, which is expressed as equivalents of standard chelators or the percentage metal chelating [130]. The binding ability of ligands could also be evaluated by stability constants [188]. For example, in Fe chelation, ferrozine and 2,2′-bipyridine are used as metallochromic indicators and EDTA and deferoxamine as standard metal chelators [130,189]. This approach enables the evaluation of the dependence between the structure of the PC ligand and its metal chelating activity; thus, a comparative analysis of this indicator extracts of medicinal plants is possible [189,190].

3.2. Chelating Effects In Vivo

In the studies of in vivo binding between phenolic metabolites and Men+, two aspects should be highlighted: (1) production of blue anthocyanins (ACNs) in blue flowers and (2) the defensive role of chelation in plant tolerance to toxic metal exposure. It is noteworthy that the vast majority of the in vivo studies on this topic are devoted to ACNs as metal chelators. This is due to the fact of the availability of non-destructive methods for the binding identification based on the spectral characteristics of ACN-Men+ complexes in plant tissues [13]. Blue flower coloration is associated with copigmentation of ACNs and the formation of pigment–copigment–Men+ complexes (Yoshida et al., 2009). Copigmentation can be performed with and without the participation of metal ions [191]. Such studies could be systematized according to two directions, which differ in their levels of elucidation of the content and structural organization of the pigment complex. The first direction is the evaluation of various aspects of the formation of non-stoichiometric ACNs’ metallocomplexes, which are stabilized due to copigmentation with caffeoyl or coumaroyl derivatives of quinic acids or glycosylated flavonoids [192]. For Hydrangea macrophylla, Phacelia campanularia, and Tulipa gesneriana flowers, the major pigments of those complexes are the delphinidin glucosides, while the pigments of Meconopsis grandis flowers are primarily composed of cyanidin glucosides [192]. The binding with metal ions (Fe3+, Al3+, Mg2+) is considered as a necessary condition for the formation of those pigment complexes [192].
Another direction is systemic studies, which have resulted in the establishment of the unique structure of metalloanthocyanins. According to the term’s definition, metalloanthocyanin is a self-assembled, supramolecular complex metal-containing pigment, which comprises 6 ACN molecules, 6 flavone molecules, and 2 metal ions [192]. During blue flowers’ colour formation, three major mechanisms can be implemented, i.e., self-association, copigmentation, and metal complexation [192]. To date, the following metallochelates have been isolated and identified from blue flowers: protocyanin (Centaurea cyanus), commelinin (Commelina communis), protodelphin (Salvia paterns), cyanosalvianin (Salvia uliginosa), nemophilin (Nemophila menziesii) [192]. The constitutive components of those pigments are the ACNs having a chelating centre with two (cyanydin) or three (delphinidin) hydroxyls, flavonoid apigenin derivatives, and Mg2+ and Fe2+ ions [192]. In protocyanin, an additional coordination link of Ca2+ ions with flavone molecules has been established [192]. The advantages of the supramolecular structure for plants are the stability of the pigment complex at physiological pH and the increased tolerance to UV radiation, which play an important role in the implementation of the main function of ACNs during plant blooming under sun irradiation. The simultaneous presence of non-associated and chelated ACN molecules explains the phenomenon of purple coloration due to the mixing of the two colour stimuli, red and blue [13]. Different ratios between those ACN forms, when present in vivo, create various superpositions of their colour stimuli, thus resulting in colour variability with different hues of purple plant coloration, a feature that has an important evolutionary significance, as it allows a wide diversity of plant colours and better alignment with pollinators [193,194]. One peculiarity of metallo-anthocyanins is the ability to replace coordinated biogenic Men+ with abiogenic Me ions, while the spectral characteristics of the metallocomplexes are retained. Thus, commelinin-like pigments can be formed by replacing Mg2+ with Cd2+, Zn2+, Co2+, Ni2+, and Mn2+ [192].
The ACNs’ capability of binding various Men+ ions during the formation of pigment complexes in flowers allows hypothesizing that the chelating properties could be engaged for a different purpose—to decrease the toxicity of endogenous metals, thus increasing plant metal tolerance [195]. This hypothesis was confirmed using maize as a metal-excluder plant; the in vivo chelating effect of cyanidin-3-glucoside (Cya-3-glu) in maize root tissues was found for nine exogenous Men+ (Mg2+, Fe2+, Cd2+, Ni2+, Pb2+, Al3+, VO3−, MoO42−, Cr2O72−) [11,196]. The reversible nature of Cya-3-glu–Pb2+ binding was found in maize roots, which can be controlled by manipulating the pH in the root solution [13]. An increase in the Pb2+ concentration in the root nutrient solution resulted in the increased formation of Cya-3-glu–Pb2+ complexes in maize roots in a dose-dependent manner [196].
The formation of ACN–metal complexes in the hypocotyls of Brassica plants was found upon their treatment with MoO42− and WO42− ion solutions [197,198].
The capability of binding Men+ was also demonstrated for other PCs localized in various plant tissues. Thus, the study of ACNs’ distribution over the roots of Lotus pedunculatus Cav. confirmed the hypothesis about metal binding and detoxifying by proanthocyanins in plant vacuoles [199]. Al(III) metallocomplexes with epigallocatechin gallate and proanthocyanins were identified in the leaves, stems, and roots of Camellia sinensis [179] and an oenothein B (dimeric macrocycle ellagitannin) in the roots of Eucalyptus camaldulensis [180]. Cd2+ binding by polymerized polyphenols was demonstrated in the leaves of water plants [200]. According to Rocha et al. [201], the reduction of mercury toxicity in plants can be associated with the chelating activity of gallic acid.
The confirmation of the role of PC ligands in plant–metal homeostasis is the identification of the complexes of Cu(II) with quercetin, luteolin, and syringic acid in the berries of Euterpe oleraceae and Vaccinum myrtyllus [202].
Aluminium stimulates maize plants to secrete into the rhizosphere various endogenous PCs (catechin, catechol, quercetin) capable of complexing with Al3+, thus implementing one of the mechanisms of plant tolerance to the metal excess in the root nutrition medium [203]. The role of root-secreted coumarins was shown in iron-deficient plants by the acquisition of Fe through reduction and chelation [204,205]. It is noteworthy that the binding effects/capacity of the chelators with different structural groups (including PCs) by trace elements are considered as one of the mechanisms of the soil–plant interface [206]. In this relation, it should be highlighted that metal(loid)-induced accumulation of PCs by plants is associated with their protecting role in plant metal tolerance [25].

3.3. Properties of Phenolic Chelators

Summarizing the data on PC–Men+ binding presented in Section 3.1 and Section 3.2 of this review, we can outline both general and specific properties of phenolic chelators. The general properties of different PC subgroups are the following:
1.
Presence of Men+ binding sites with O atoms of carbonyl, hydroxyl, or carboxylate groups: Depending on the phenolic subgroup and a number of OH- groups as substituents, different chelation variants are possible. Thus, for quercetin, Men+ chelation can take place at three binding sites (Figure 3). For ACNs, Men+ binding occurs due to two or three hydroxyl substitutions in the B ring [13]. The formation of chelate structures with an unsaturated cycle with two or three coordinated O atoms defines the stability of such metallocomplexes. However, the coordination of Men+ with two O atoms of the carboxylate groups of PCs is also possible, e.g., in p-coumaric acid [207];
2.
Universal affinity of PCs in relation to different Men+ in both cationic and anionic forms: During the systematizing of the experimental and literature data, we revealed that PC ligands from various subgroups form complexes with 69 Men+ ions (63 chemical elements; Table 1, Figure 1). Binding to PCs is a characteristic of/universal for the elements differing in their roles in plants:
  • Essential macronutrients (Ca, Mg, K) and essential micronutrients (Fe, Mn, Zn, Ni, Cu, B, Mo), which are necessary for the plant life cycle, cannot be substituted by other elements and are directly involved in plant metabolism [208];
  • Beneficial elements (Al, Co, Na, Se) that promote the growth of various plant species, but are not essentially required for the completion of the plant life cycle [209,210];
  • Non-essential metal(loid)s that are not involved in primary plant metabolism;
  • Rare earth elements (REEs), which include the lanthanide group with 15 elements (Figure 2).
The complexation of PC ligands with 11 elements essential for human life (Na, Mg, K, Ca, Mn, Fe, Co, Cu, Zn, Mo, Se) has also been confirmed [211]. Such universalism of PC ligands is due to the multi-elemental composition of the plant metallome. Thus, according to Watanabe et al. [212], in the leaves of species from different families of terrestrial plants, 42 chemical elements have been found. An additional argument for a close PC–Men+ interrelation is the correlation between the plant accumulation of essential, beneficial, and non-essential elements with the content of total phenolics and total flavonoids [213];
3.
Formation of complexes with different Men+:ligand molar ratios, depending on the nature of the metal ion, the structure of the phenolic chelator(s), and the pH [77,78];
4.
The phenolic ligands’ capacity to form complexes with multiple different Men+. The examples of such chelates are the heterobimetallic complexes of quercetin–Cu–Sn2 and quercetin–Zn–Sn2 [214];
5.
The capability of certain PCs to form mixed ligand complexes and multiligand metal–phenolic assemblies. Thus, mixed ligand complexes have been obtained for Pt(II)–naringin–caffeic acid, Pt(II)–naringin–sinaptic acid, and complexes of V(V) with those ligands [106]. Porkodi and Raman [215] synthesized mixed complexes of curcumin and quercetin derivatives with Co(II), Ni(II), Cu(II), and Zn(II). This capability was employed for the fabrication of hybrid functional materials using metal–phenolic networks based on the polyphenol components of green tea infusions [216];
6.
Complexation of PC ligands with different metal(loid)s’ species (ions, oxides, isotopes, nanoparticles): In addition to Men+, solid-phase chelation of flavonoids has been confirmed for Al2O3, SiO2, and TiO2 [217,218,219]. Radiolabelled complexes have been synthesized for 99mTc with curcumin, rutin, and luteolin [139,220,221], and 68Ga with curcumin [139]. The chelating capacities of PCs are used in the “green chemistry” synthesis of biocompatible nanomaterials [222];
7.
The formation of metabolites with chelating capacity at all stages of the phenolic biosynthetic pathway: Such a property is manifested regardless of the features of a particular biosynthetic pathway, which are determined by the complexity of the interrelated and consequent transformations of metabolites from various subgroups and the specificity of the dominant metabolite accumulated in a particular plant species. According to the systematization of Men+ complexation reactions undertaken, we identified 18 subgroups of PCs capable of metal(loid) binding (Table 1). This enables multivariant scenarios of Men+ binding in plants, thus confirming the universal nature of phenolic chelators;
8.
Metal ions’ binding has been demonstrated for PCs localized in plant tissues with various functions and for endogenous metabolites secreted by plants into the rhizosphere. Thus, binding with Men+ has been confirmed for PCs in roots [11,180,196,199], hypocotyls [197,198], stems [179], leaves [200], flowers [192], berries [202], and the rhizosphere [203,204,205];
9.
The chelating capacity of PCs is manifested in different physiological processes of plants. Phenolic metallocomplexes are engaged in photoreception and photoprotection [12], plant–pollinator interactions [12], antioxidant and prooxidant mechanisms [188], metal detoxifying in plant tissues and the rhizosphere [11,196,197,198,199,200,203], vacuolar sequestration [12], the mobilization and phytoavailability of deficient elements [204,205];
10.
The ability of plant PCs to modulate the metallome in animals: Thus, the use of hesperidin, naringin, and quercetin as dietary supplements results in an altered element profile [223,224].
Besides the aforementioned general properties inherent to the chelators from different phenolic subclasses, the specific ACN properties should be distinguished, which are due to the chemical structure of those compounds:
  • The phenomenon of pH-dependent dynamic equilibrium between different structural forms of ACNs: Both unbound and chelated ACN forms could be present simultaneously in plant tissue, thus creating its colour variation: red–purple–blue [13];
  • Reversible nature of ACN binding with metal ions in plant tissue; the unbound ACN form can be regenerated by varying the pH: Such a peculiarity has been established for the binding of Pb2+ with Cy-3-glu in maize roots [13]. This feature is based on the pH-dependent transformation of the ACN structure;
  • ACNs and copigmentation: According to Trouillas et al. [191], copigmentation is defined as the formation (in the presence or absence of metal ions) of non-covalent complexes involving anthocyanin or anthocyanin-derived pigment(s) and a copigment(s), with the resulting changes in the optical properties of the pigment complex. Major natural copigments are the phenolic metabolites: hydrolysable tannins, flavonols, flavones, dihydroflavonols, flavanones, phenolic acids, and derivatives thereof [191]. ACNs themselves can act as copigments due to the self-association of the two molecules [191].
Specific features of ACNs define their unique properties as plant chelators due to the following points:
(i)
The ACN chromophore within the metallocomplex absorbs light in the UV and visible region, which provides the UV protection of plant tissue and attracts pollinators to flower petals. Binding with Men+ increases can resistance to solar radiation;
(ii)
ACN copigmentation occurs with the participation of both PCs and chelators from other groups (organic acids, amino acids). Moreover, some ACNs and phenolic copigments contain malonyl, succinyl, and quinic fragments in their molecules [192]. Such structural features of ACNs and copigments increase the number of binding sites and modify the chelation activity;
(iii)
The ACNs’ capacity to biotransform biogenic (nutrient) elements and engage them in the processes of photoprotection and plant–pollinator interactions, which was formed during natural evolution, can be used by plants for the other functional role—detoxication of abiogenic metals as pollutants resulting from man-induced activities.
Therefore, the performed analysis of in vitro and in vivo complexing, along with the revealed general capacity of chelation of PCs from different subclasses and the specific properties of ACNs allow postulating that phenolic compounds comprise a separate group of plant chelators. The preceding systematization of the experimental results enables assessing the role of phenolic ligands in the context of metallomics.

4. Concept of Metallophenolomics

4.1. Definitions

According to the main concepts of metallomics, the analysis of the metallome, in addition to the study of metal(loid) content and their distribution between the cell compartments, also includes the determination of the complex formation process, which involves metal ions and a bioligand(s) [65]. In this respect, one of the variants of the differentiation of metallomics into its subgroups is based on the nature of the bioligand [65]. Our systematization of the data on the chelating capacities of PCs (Section 3.3 enables, analogous to the metalloproteome, the separation of the metallophenolome; the object of the latter is the interactions between phenolic ligands and metal ions. Accordingly, the corresponding branch related to the study of the metallophenolome should be defined as metallophenolomics. Metallophenolomics is a ligand-oriented subgroup of metallomics; thus, it represents an integrated approach to study the complexation of plant phenolics with metals/metal(loid)s in different research subjects.
This term we firstly introduced in previous publications [225,226] to describe the interactions between metals and plant PCs. It should also be noted that in the context of metabolomics, the following terms had already been proposed previously: phenol-omics [227], phenol metabolomics [228], lignomics [229].

4.2. Research Subjects

According to the main concepts of metallomics [26], metallophenolomics is an interdisciplinary area of research and relates to both basic (chemistry, botany, biology, medicine, pharmacy, agriculture), as well as applied science (food science, nutritional science, toxicology, health science, environmental/green science). It is reasonable to analyse the research area of metallophenolomics in relation to the research subjects that were defined by Haraguchi [26] for metallomics. Research subjects for metallophenolomics are provided in Figure 4:
(1)
Quantitative distribution and imaging analysis of elements in plant tissues. The use of the chelating effect of PCs is exemplified by the determination of Al localization in plant tissues by complexing to morin as the fluorochrome with subsequent detection by confocal laser microscopy [230,231]. This technique complements histochemical assays, where chelating dyes are used, and metal localization is examined by light microscopy [232] or visible reflectance and tristimulus colorimetry [233]. Tannic acid can be employed as a natural chelator for labile iron imaging in the prevention and treatment of iron-associated cancer or other iron-overload disorders [234];
(2)
Speciation of elements in plants: To establish the chemical forms of metals in which they are bound to plant chelators, the general analytical approaches of metallomics can be employed, along with a set of specific methods based on the structural peculiarities of these bioligands.
General approaches include universal techniques for analysing the metal-containing biomolecules, i.e., hyphenated techniques (e.g., HPLC-ICPMS/ESI-MS) [235]. Hydrangea blue complexes composed of 3-O-glycosyldelphinidin, Al3+, and 5-O-acilquinic acid were investigated by electrospray-ionization mass spectrometry [236]. However, some supplementary non-destructive methods for phenol bioligands’ investigation are available due to their following specific features. Firstly, some of those metabolites (e.g., anthocyanins) are in most cases localized in surface plant tissues. Secondly, the chromophore system in some phenolic metal chelating molecules defines selective light absorption in the visible region, thus enabling the use of non-destructive methods (e.g., reflectance spectroscopy, tristimulus colorimetry) based on the interaction of the light beam with pigmented plant tissues. Thirdly, the differences in spectral characteristics between the unbound and chelated forms of bioligands are the markers for identifying the in vivo binding. Such specific analytical techniques could be exemplified by identifying ACN binding to Men+ in flowers and roots [13];
(3)
Structural analysis of metal binding by phenolic compounds: This research field includes investigating the structural features of PCs from various subclasses along with their chelation sites [77], the structure–activity relations (antioxidant and metal chelating properties) [237], in silico prediction of the binding sites’ structure, and the potential biological activities of metallocomplexes [238];
(4)
Elucidation of the reaction mechanisms of the metallophenolome using model phenol–metal complexes: Phenolic–Men+ complexation can alter the antioxidant activity of PCs due to the reduction of transition metal ions and the induction of Fenton reactions [237]. Therefore, the iron or copper chelating properties of PCs are considered to be a variant of the antioxidative effect [188,189,239]. The antioxidant activity of some phenolic–metal complexes is superior to that of the parent ligands [77]. The effect of metal binding on the antioxidant activity of the molecule was studied in ferrous flavonoid mixtures [148]. An important element of redox processes in biological systems is also the prooxidant activity of phenolics, which is stimulated by Men+ [240,241]. Flavonoid–metal complexes may exhibit superoxide dismutase activity, and their radical-scavenging activity is superior to the unbound flavonoid [242]. Phenolic–metal complexes are used in elucidating the reaction mechanisms with biologically important molecules such as DNA [243,244], proteins [243], pectins [245], and lipids [77]. Model phenolic metallocomplexes are employed in biomimetic studies of the mechanisms of action of metalloenzymes, which use flavonoids as a substrate [77];
(5)
Identification of unknown metallophenolics: An important problem in this research field remains the identification of stoichiometric and non-stoichiometric anthocyanin metallocomplexes in blue flowers [192] and phenolic–Men+ complexes in plant tissues and the rhizosphere under the toxic impact of metals [13,203]. An example of such a novel approach is the identification of metal binding PC ligands in wine [246];
(6)
Targeted analysis of metal chelators in phenolic metabolism: Metabolomic profiling enables identifying the effects of various metals on the qualitative composition and quantitative content of individual phenolic metabolites as potential chelators. Thus, Cu’s effect on Cucumber sativus significantly upregulates 4-hydroxycinnamic acid compared to other phenolic metabolites [247]. The treatment of Helianthus annuus with Cr enabled identifying ten isocoumarin derivatives as target metal chelating compounds [248]. Different changes in the metabolic profiles of flavonols and hydroxicinnamic acids were identified for two metallicolous populations of Arabidopsis halleri, which demonstrated different mechanisms of Cr tolerance [249]. The higher tolerance of a red- versus a green-leafed cultivar of sweet basil against boron toxicity was hypothesized to be partially related to the capability of anthocyanins to act as B chelators [250,251,252,253];
(7)
Medical diagnosis of health and diseases related to trace metals: The chelating capacities of PCs are employed in numerous medical diagnostic techniques. Tannic acid may be applied for chelation and imaging of labile iron in iron-associated cancer or other iron-overload disorders [234]. The complex of morin with 68Ga was proposed as a novel radiopharmaceutical for diagnostic purposes and kidney cancer cell labelling [254]. The morin metal complexes with DNA were confirmed as an effective tool for the discrimination of anticancer drugs’ binding mechanism to DNA [255];
(8)
Metallodrug design: PC–Men+ complexes demonstrate a broad spectrum of biological activities (anti-/pro-oxidant, antimicrobial, antiviral, anti-inflammatory, anti-diabetic, anticonvulsant, anticancer) [77,78,79,80]. An option to use various ligands differing in their binding sites and the universal property of chelating multiple Men+ corroborates the prospects of PCs in metallodrug design. Flavonoid–Men+ complexes are viewed as a novel class of therapeutical agents [76]. The complexation of PCs with Men+ enables obtaining metallocomplexes with improved biological activity compared to their parent ligands [77,79]. An essential advantage of those complexes is the use of non-toxic natural chelators, thus reducing the potential toxicity of metallodrugs in chemotherapy [256];
(9)
Chemical evolution of the living systems and organisms on Earth: The evolution and sophistication of mechanisms for chelating different elements are deemed as a factor of the evolutionary development of organisms [257]. Plant systems’ evolution involves metal homeostasis networks [4,258,259]. One of the important evolutionary aspects of elemental hyperaccumulation is based on the involvement of multiple kinds of chelators in plant tolerance mechanisms [260]. The important evolutionary role of flavonoids with antioxidant and chelating capacities was shown in the adaptation of metallicolous populations, wherein divergent strategies were revealed for Cd uptake, translocation, and detoxifying in different genetic units of Arabidopsis halleri [249]. The chelation mechanism, which was formed during evolutionary development toward essential elements, is utilized in plants to detoxify non-essential metal(loid)s as pollutants from anthropogenic sources [13]. The universality of this tolerance mechanism was demonstrated in relation to a new class of contaminants—metal-containing nanoparticles (NPs) [222]. ACN–Men+ complexation is the important factor of the colour evolution of flowers, which enhances plant polychroism in plant–pollinator interactions [192];
(10)
Mission-oriented biometal sciences. Environmental science: One of the adaptive mechanisms in plant tolerance strategies against metal(loid) toxicity is deemed to be the chelation process involving different kinds of chelators [261,262]. The chelating properties of phenolic metabolites are integrated in the omics approach to study plant responses to metal stress (transcriptomics, proteomics, metabolomics, ionomics) [263]. To improve plant metal tolerance, various techniques are utilized, including pre-treatment by phenolic-rich plant extracts with chelating capacity [264,265]. Chelation processes are crucial in phytoremediation technologies for metal-contaminated areas [266]. The phytoremediation potency of plants depends on various factors including the content of phenolic chelators in plant tissues [267]. Chelating capacity determines the efficacy of phenolics as natural removing agents (biosorbents) from contaminated soils and wastewaters [268,269,270,271]. The defensive role of phenolic metabolites in plant responses to metal-containing NPs as a new class of contaminants has already been demonstrated [272,273,274].
Food science and nutrition: The chelating capacity of anthocyanins is utilized for their stabilization by metal ions and creating blue food colorants with health-promoting effects [275,276]. To enhance the photostability of ACNs as colorants, they are incorporated into mesoporous silica granules containing various metal ions [277]. ACN–Fe3+ complexation with coating was efficient in preserving anthocyanin pigments in thermally processed fruit products [278]. PCs with chelating and/or antioxidative effects are deemed as important food ingredients capable of preventing the accumulation of non-essential metals in the human body [279]. Due to this, phenol-rich plant species are extensively studied as potential sources of phytochemicals with high antioxidant and chelating activities [280,281,282]. Experimental elaborations of this issue comprise the basement for dietary strategies for the treatment of metal toxicity [283]. The human-health-promoting effects of PCs are used in food antioxidant applications [284], global flavonoid intake [285], food metabolome [286], foodomics [287], and nutrigenomics [288]. Ferric anthocyanin chelators increase the pH sensitivity of indicator films for monitoring freshness and preparing intelligent packaging of food [289].
Medicine: A broad spectrum of pharmacological activities determines the significance of phenolic–metal complexes as a novel class of therapeutic agents [76,79]. A key feature of the production of those drugs is the use of non-toxic natural chelators [256]. This peculiarity asserts the prospects of phenolic ligands versus synthetic chelators for metal chelation therapy [290,291]. Phenolics as food components with antioxidant and chelating activities are employed in developing preventive and therapeutic strategies for metal intoxication [292,293,294]. The chelating activity of flavonoids is appreciated as an important factor in multitarget-directed ligand strategies in the management of Alzheimer’s disease [295]. The novel direction of nanomedicine is associated with biocompatible flavonoid-mediated metal-containing nanomaterials with a dual function as both nano-carriers and nano-drugs in numerous medicinal applications [222].
Cosmetology: Owing to the antioxidant and chelating properties, plant phenolics are utilized as bioactive ingredients in the cosmetic industry [296]. A multicomponent powder of polyphenol-rich extracts was efficient in counteracting skin damages induced by metal deposition from air pollution in the environment [297]. Plant anthocyanins with chelating effects are used as renewable hair dyes that are free of any toxic effects of synthetic dyes [298].
Coloration technology: Metal chelating flavonoids are important natural colorants in textile dying [77]. To ensure affinity between textile materials and pigments, metal salts (Sn, Al, Fe, Cu, Cr) are needed as mordants [299]. The flavonoid–metal complexation in the dying process improved pigment stability and altered the colour characteristics [77]. By varying the flavonoid composition, combining different plant species, metal ions, and conditions of dying, striking colour diversity can be achieved [300]. The use of natural pigments represents an alternative to synthetic dyes in the development of sustainable and eco-friendly processes [301].
Chemical science: Flavonoid–metal complexation is utilized in preparative, analytical, and synthetic approaches. Thus, an ACN purification method was proposed, which is based on the ligand exchange mechanism and uses cationic resins charged with Fe(III) [302]. In quantitative analysis, flavonoid–metal chelation could be used in two ways. On the one hand, flavonoids are the chromogenic reagents for the determination of various metals by spectrophotometric or fluorometric detection [77,135]. On the other hand, flavonoid content in plant samples can be determined using metal ions as analytical reagents [77,303,304]. The capacity of PCs to reduce metal ions and stabilize them into nanoparticles by, i.e., chelation mechanism, comprises the basement for the green synthesis of metal-containing nanomaterials [305,306].
Geochemistry: An important role of phenolics in the complexation and reduction of Fe in the dissolution and transport of terrestrial iron to aquatic ecosystems was confirmed, being a key link in the global coupled iron and carbon cycles [307].
Corrosion protection: The capacity of flavonoids with catechol fragments to complex with Fe(III) plays a significant role in the corrosion inhibition mechanism [308]. PCs as ingredients of plant extracts determine the efficacy of environmentally sustainable and “green” corrosion inhibitors for metals and alloys [309].
Material science: Phenol–metal chelation is important for the creation of novel hybrid multifunctional biomaterials for chemical, biomedical, and environmental applications. Thus, direct gelation between tannic acid and metal ions produces metallogels, films, and capsules [310]. A versatile platform for the development of functional hybrid materials is the synthesis of coordination-driven assemblies of metal–phenolic networks [311]. To create new biopolymers, a formaldehyde-assisted metal–ligand crosslinking strategy was proposed for the synthesis of metal–phenolic coordination spheres based on the principles of sol–gel chemistry [312]. A nanostructured porous carbon monolith was obtained based on phenolic–metal interactions (tannic acid–Zn chloride) as sorbent multi-scale molecules. The resulting material provides versatile adsorption behaviours ranging from small gas molecules to larger molecules such as dyes, oils, and organic solvents [313]. A natural antibiotic system was developed using tannic acid–metal coordination coating of curcumin nanoparticles [314].
Solar cell science: One of the most promising renewable and sustainable technologies is the dye-sensitized solar cell technology, an environmentally friendly method for producing electricity [315,316]. Among natural dyes from various plants, flavonoids and, in particular, ACNs are crucial, due to their ability to chelate TiO2 NPs [317,318].
Therefore, the research area of metallophenolomics includes major research subjects that have been defined for metallomics [21,26] and corroborates the multidisciplinary nature of this scientific direction [32,211].

5. Conclusions

The polyfunctionality of phenolic compounds (PCs) determines their important adaptive role in plant responses to metal(loid) stress and tolerance. Binding sites in the molecules of phenolics from various subclasses underlie their chelating capacity toward different metal(loid)s, metal(loid) oxides, and metal(loid)-containing nanoparticles. The chelating effect is exhibited by phenolic compounds localized in various plant tissues and organs (roots, leaves, hypocotyls, flowers), as well as the rhizosphere. Phenolic–metal(loid) complexation is pivotal for multiple plant processes: photoreception and photoprotection, plant–pollinator interactions, antioxidant and prooxidant mechanisms, metal detoxifying in plant tissues and the rhizosphere, vacuolar sequestration, and mobilization of elements and phytoavailability. Among the PCs, the structural features of anthocyanins determine their specific properties as phenolic chelators, which is associated with the dynamic equilibrium between their different structural forms and copigmentation effects. An integrated approach studying metal(loid)–phenolic complexation is described as metallophenolomics, a ligand-oriented subgroup of the field of metallomics. The concept of metallophenolomics opens a novel scientific route of research, which enables joining the efforts of the scientific community in multidisciplinary investigations of phenolic–metal interactions and metal-assisted areas of applications.

Author Contributions

Conceptualization, S.A.S., V.S.F., and M.L; writing—original draft preparation, S.A.S. and V.S.F., writing—revision S.A.S., V.S.F., and M.L supervision, M.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Richards, C.L.; Hanzawa, Y.; Katari, M.S.; Ehrenreich, I.M.; Engelmann, K.E.; Purugganan, M.D. Perspective on ecological and evolutionary systems biology. Annu. Rev. Plant Biol. 2009, 35, 331–351. [Google Scholar] [CrossRef]
  2. Hill, M.K. Understanding Environmental Pollution, 3rd ed.; Cambridge University Press: New York, NY, USA, 2010; 534p, ISBN 978-0-521-73669-5. [Google Scholar]
  3. Hasanuzzaman, M.; Fujita, M. Heavy metals in the environment: Current status, toxic effects on plants and phytoremediation. In Phytotechnologies: Remediation of Environmental Contaminants; Anjum, N.A., Pereira, M.E., Ahmad, I., Duarte, A.C., Umar, S., Khan, N.A., Eds.; CRC Press: Boca Raton, FL, USA, 2013; pp. 7–73. ISBN 1439875189. [Google Scholar]
  4. Sharma, A.; Kapoor, D.; Gautam, S.; Landi, M.; Kandhol, N.; Araniti, F.; Ramakrishnan, M.; Satish, L.; Singh, V.P.; Sharma, P.; et al. Heavy metal induced regulation of plant biology: Recent insights. Physiol. Plant. 2022, 174, e13688. [Google Scholar] [CrossRef] [PubMed]
  5. Angulo-Bejarano, P.I.; Puente-Rivera, J.; Cruz-Ortega, R. Metal and metalloid toxicity in plants: An overview on molecular aspects. Plants 2021, 10, 635. [Google Scholar] [CrossRef] [PubMed]
  6. Corso, M.; de la Torre, V.S.G. Biomolecular approaches to understanding metal tolerance and hyperaccumulation in plants. Metallomics 2020, 12, 840–859. [Google Scholar] [CrossRef]
  7. Hall, J.Á. Cellular mechanisms for heavy metal detoxification and tolerance. J. Exp. Bot. 2002, 53, 1–11. [Google Scholar] [CrossRef] [PubMed]
  8. Pilon-Smits, E. Phytoremediation. Annu. Rev. Plant Biol. 2005, 56, 15–39. [Google Scholar] [CrossRef]
  9. Rauser, W.E. Structure and function of metal chelators produced by plants. Cell Biochem. Biophys. 1999, 31, 19–48. [Google Scholar] [CrossRef]
  10. Anjum, N.A.; Hasanuzzaman, M.; Hossain, M.A.; Thangavel, P.; Roychoudhury, A.; Gill, S.S.; Rodrigo, M.A.M.; Adam, V.; Fujita, M.; Kizek, R.; et al. Jacks of metal/metalloid chelation trade in plants—An overview. Front. Plant Sci. 2015, 6, 192. [Google Scholar] [CrossRef] [PubMed]
  11. Fedenko, V.S. Cyanidin as endogenous chelator of metal ions in maize seedling roots. Ukr. Biochem. J. 2008, 80, 102–106. [Google Scholar]
  12. Landi, M.; Tattini, M.; Gould, K.S. Multiple functional roles of anthocyanins in plant-environment interactions. Environ. Exp. Bot. 2015, 119, 4–17. [Google Scholar] [CrossRef]
  13. Fedenko, V.S.; Shemet, S.A.; Landi, M. UV–vis spectroscopy and colorimetric models for detecting anthocyanin-metal complexes in plants: An overview of in vitro and in vivo techniques. J. Plant Physiol. 2017, 212, 13–28. [Google Scholar] [CrossRef] [PubMed]
  14. Jones, O.A.; Dias, D.A.; Callahan, D.L.; Kouremenos, K.A.; Beale, D.J.; Roessner, U. The use of metabolomics in the study of metals in biological systems. Metallomics 2015, 7, 29–38. [Google Scholar] [CrossRef] [PubMed]
  15. Pirzadah, T.B.; Malik, B.; Hakeem, K.R. Integration of “omic” approaches to unravel the heavy metal tolerance in plants. In Essentials of Bioinformatics; Hakeem, K., Shaik, N., Banaganapalli, B., Elango, R., Eds.; Springer: Cham, Switzerland, 2019; Volume III, pp. 79–92. ISBN 978-3-030-19318-8. [Google Scholar] [CrossRef]
  16. Hanus-Fajerska, E.; Wiszniewska, A.; Kamińska, I.; Koźmińska, A. Metallomic approach to enhance agricultural application of halophytes. In Handbook of Halophytes: From Molecules to Ecosystems towards Biosaline Agriculture; Grigore, M.N., Ed.; Springer: Cham, Switzerland, 2021; pp. 1953–1969. ISBN 978-3-030-57635-6. [Google Scholar] [CrossRef]
  17. Jamla, M.; Khare, T.; Joshi, S.; Patil, S.; Penna, S.; Kumar, V. Omics approaches for understanding heavy metal responses and tolerance in plants. Curr. Plant Biol. 2021, 27, 100213. [Google Scholar] [CrossRef]
  18. Raza, A.; Tabassum, J.; Zahid, Z.; Charagh, S.; Bashir, S.; Barmukh, R.; Khan, R.S.A.; Barbosa, F., Jr.; Zhang, C.; Chen, H.; et al. Advances in “omics” approaches for improving toxic metals/metalloids tolerance in plants. Front. Plant Sci. 2022, 12, 794373. [Google Scholar] [CrossRef]
  19. Lay, J.O., Jr.; Liyanage, R.; Borgmann, S.; Wilkins, C.L. Problems with the “omics”. TrAC Trends Anal. Chem. 2006, 25, 1046–1056. [Google Scholar] [CrossRef]
  20. Williams, R.J.P. Chemical selection of elements by cells. Coord. Chem. Rev. 2001, 216, 583–595. [Google Scholar] [CrossRef]
  21. Haraguchi, H. Metallomics as integrated biometal science. J. Anal. At. Spectrom. 2004, 19, 5–14. [Google Scholar] [CrossRef]
  22. Berni, R.; Luyckx, M.; Xu, X.; Legay, S.; Sergeant, K.; Hausman, J.F.; Lutts, S.; Cai, G.; Guerriero, G. Reactive oxygen species and heavy metal stress in plants: Impact on the cell wall and secondary metabolism. Environ. Exp. Bot. 2019, 161, 98–106. [Google Scholar] [CrossRef]
  23. Sharma, A.; Shahzad, B.; Rehman, A.; Bhardwaj, R.; Landi, M.; Zheng, B. Response of phenylpropanoid pathway and the role of polyphenols in plants under abiotic stress. Molecules 2019, 24, 2452. [Google Scholar] [CrossRef] [Green Version]
  24. Anjitha, K.S.; Sameena, P.P.; Puthur, J.T. Functional aspects of plant secondary metabolites in metal stress tolerance and their importance in pharmacology. Plant Stress 2021, 2, 100038. [Google Scholar] [CrossRef]
  25. Fedenko, V.S.; Shemet, S.A.; Guidi, L.; Landi, M. Metal/metalloid-induced accumulation of phenolic compounds in plants. In Metal Toxicity in Higher Plants; Landi, M., Shemet, S.A., Fedenko, V.S., Eds.; Nova Science Publishers: New York, NY, USA, 2020; pp. 67–115. ISBN 978-1-53616-790-0. [Google Scholar]
  26. Haraguchi, H. Metallomics research―Good luck on new publication. Met. Res. 2021, 1, rev-1–rev-13. [Google Scholar] [CrossRef]
  27. Lobinski, R.; Becker, J.S.; Haraguchi, H.; Sarkar, B. Metallomics: Guidelines for terminology and critical evaluation of analytical chemistry approaches (IUPAC Technical Report). Pure Appl. Chem. 2010, 82, 493–504. [Google Scholar] [CrossRef]
  28. Jakubowski, N.; Lobinski, R.; Moens, L. Metallobiomolecules. The basis of life, the challenge of atomic spectroscopy. J. Anal. At. Spectrom. 2004, 19, 1–4. [Google Scholar] [CrossRef]
  29. Maret, W. Metallomics: A Primer of Integrated Biometal Sciences; Imperial College Press: London, UK, 2016; 156p, ISBN 178326830. [Google Scholar]
  30. Szpunar, J. Metallomics: A new frontier in analytical chemistry. Anal. Bioanal. Chem. 2004, 378, 54–56. [Google Scholar] [CrossRef]
  31. Mounicou, S.; Szpunar, J.; Lobinski, R. Metallomics: The concept and methodology. Chem. Soc. Rev. 2009, 38, 1119–1138. [Google Scholar] [CrossRef]
  32. Maret, W. Metallomics: The Science of Biometals and Biometalloids. In Metallomics: The Science of Biometals; Arruda, M., Ed.; Springer: Cham, Switzerland, 2018; pp. 1–20. ISBN 978-3-319-90143-5. [Google Scholar] [CrossRef]
  33. Maret, W. An Appraisal of the Field of Metallomics and the Roles of Metal Ions in Biochemistry and Cell Signaling. Appl. Sci. 2021, 11, 10846. [Google Scholar] [CrossRef]
  34. Metallomics: Recent Analytical Techniques and Applications; Ogra, Y.; Hirata, T. (Eds.) Springer: Tokyo, Japan, 2017; p. 372. ISBN 4431567933. [Google Scholar] [CrossRef]
  35. Metallomics: The Science of Biometals; Arruda, M.A.Z. (Ed.) Springer: Cham, Switzerland, 2018; p. 279. ISBN 978-3-319-90143-5. [Google Scholar] [CrossRef]
  36. Singh, V.; Verma, K. Metals from cell to environment: Connecting metallomics with other omics. Open J. Plant Sci. 2018, 3, 001–014. [Google Scholar]
  37. Lahner, B.; Gong, J.; Mahmoudian, M.; Smith, E.L.; Abid, K.B.; Rogers, E.E.; Guerinot, M.L.; Harper, J.F.; Ward, J.M.; McIntyre, L.; et al. Genomic scale profiling of nutrient and trace elements in Arabidopsis thaliana. Nat. Biotechnol. 2003, 21, 1215–1221. [Google Scholar] [CrossRef]
  38. Salt, D.E.; Baxter, I.; Lahner, B. Ionomics and the study of the plant ionome. Annu. Rev. Plant Biol. 2008, 59, 709–733. [Google Scholar] [CrossRef]
  39. El-Esawi, M.A.; Sinha, R.P.; Chauhan, D.K.; Tripathi, D.K.; Pathak, J. Role of ionomics in plant abiotic stress tolerance. In Plant Life Under Changing Environment: Responses and Management; Tripathi, D.K., Chauhan, D.K., Prasad, S.M., Ramawat, N., Singh, V.P., Sharma, S., Dubey, N.K., Eds.; Academic Press: London, UK, 2020; pp. 835–860. ISBN 978-0-12-818204-8. [Google Scholar] [CrossRef]
  40. Singh, A.; Jaiswal, A.; Singh, A.; Tomar, R.S.; Kumar, A. Plant ionomics: Toward high-throughput nutrient profiling. In Plant Nutrition and Food Security in the Era of Climate Change; Kumar, V., Srivastava, A.K., Suprasanna, P., Eds.; Academic Press: London, UK, 2022; pp. 227–254. ISBN 978-0-12-822916-3. [Google Scholar] [CrossRef]
  41. Li, Y.F.; Chen, C.; Qu, Y.; Gao, Y.; Li, B.; Zhao, Y.; Chai, Z. Metallomics, elementomics, and analytical techniques. Pure Appl. Chem. 2008, 80, 2577–2594. [Google Scholar] [CrossRef]
  42. Tripathi, R.D.; Tripathi, P.; Dwivedi, S.; Dubey, S.; Chakrabarty, D. Arsenomics: Omics of arsenic metabolism in plants. Front. Physiol. 2012, 3, 275. [Google Scholar] [CrossRef]
  43. Saeed, M.; Quraishi, U.M.; Malik, R.N. Arsenic uptake and toxicity in wheat (Triticum aestivum L.): A review of multi-omics approaches to identify tolerance mechanisms. Food Chem. 2021, 355, 129607. [Google Scholar] [CrossRef] [PubMed]
  44. Martínez-Castillo, J.I.; Saldaña-Robles, A.; Ozuna, C. Arsenic stress in plants: A metabolomic perspective. Plant Stress 2022, 3, 100055. [Google Scholar] [CrossRef]
  45. Zhang, P.; Georgiou, C.A.; Brusic, V. Elemental metabolomics. Briefings Bioinf. 2018, 19, 524–536. [Google Scholar] [CrossRef] [PubMed]
  46. Outten, F.W.; Twining, B.S. Metal homeostasis. In Wiley Encyclopedia of Chemical Biology; Tadhg, P., Begley, T.P., Eds.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2008; pp. 1–10. ISBN 9780471754770. [Google Scholar] [CrossRef]
  47. Li, Y.F.; Sun, H. Metallomics in Multidisciplinary Research and the Analytical Advances. At. Spectrosc. 2021, 42, 227–230. [Google Scholar] [CrossRef]
  48. Van der Ent, A.; Harris, H.H. Phytometallomics. Metallomics 2020, 12, 324–325. [Google Scholar] [CrossRef] [PubMed]
  49. Li, X.; Liu, T.; Chang, C.; Lei, Y.; Mao, X. Analytical Methodologies for Agrometallomics: A Critical Review. J. Agric. Food Chem. 2021, 69, 6100–6118. [Google Scholar] [CrossRef]
  50. López-Barea, J.; Gómez-Ariza, J.L. Environmental proteomics and metallomics. Proteomics 2006, 6, S51–S62. [Google Scholar] [CrossRef]
  51. Blasco, J.; Rodríguez-Moro, G.; Callejón-Leblic, B.; Ramírez-Acosta, S.; Arellano-Beltrán, F.; Arias-Borrego, A.; García-Barrera, T.; Gómez-Ariza, J.L. Environmental metallomics and metabolomics in free-living and model organisms: An approach for unraveling metal exposure mechanisms. In Environmental Metabolomics; Álvarez-Muñoz, D., Farré, M., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 91–119. ISBN 978-0-12-818196-6. [Google Scholar] [CrossRef]
  52. Chen, B.; Hu, L.; He, B.; Luan, T.; Jiang, G. Environmetallomics: Systematically investigating metals in environmentally relevant media. TrAC Trends Anal. Chem. 2020, 126, 115875. [Google Scholar] [CrossRef]
  53. Albarède, F.; Télouk, P.; Balter, V. Medical applications of isotope metallomics. Rev. Mineral. Geochem. 2017, 82, 851–885. [Google Scholar] [CrossRef]
  54. Mahan, B.; Chung, R.S.; Pountney, D.L.; Moynier, F.; Turner, S. Isotope metallomics approaches for medical research. Cell. Mol. Life Sci. 2020, 77, 3293–3309. [Google Scholar] [CrossRef] [PubMed]
  55. Liang, Y.; Liu, Y.; Li, H.; Bai, X.; Yan, X.; Zhao, J.; Gao, Y. Advances of Synchrotron Radiation-based Radiometallomics for the Study of Uranium. At. Spectrosc. 2021, 42, 254–261. [Google Scholar] [CrossRef]
  56. Li, Y.F.; Zhao, J.; Gao, Y.; Chen, C. Nanometallomics: New Approach on Analyzing Biological Effects of Metal-Related Nanomaterials. In Toxicology of Nanomaterials; Zhao, Y., Zhang, Z., Feng, W., Eds.; Wiley-VCH Verlag: Weinheim, Germany, 2016; pp. 220–232. ISBN 978-3-527-68913-2. [Google Scholar] [CrossRef]
  57. Wang, L.; Zhao, J.; Cui, L.; Li, Y.F.; Li, B.; Chen, C. Comparative nanometallomics as a new tool for nanosafety evaluation. Metallomics 2021, 13, mfab013. [Google Scholar] [CrossRef] [PubMed]
  58. Lindahl, P.A.; Holmes-Hampton, G.P. Biophysical probes of iron metabolism in cells and organelles. Curr. Opin. Chem. Biol. 2011, 15, 342–346. [Google Scholar] [CrossRef] [PubMed]
  59. Dlouhy, A.C.; Outten, C.E. The iron metallome in eukaryotic organisms. In Metallomics and the Cell. Metal Ions in Life Sciences; Banci, L., Ed.; Springer: Dordrecht, The Netherlands, 2013; Volume 12, pp. 241–278. ISBN 978-94-007-5560-4. [Google Scholar] [CrossRef]
  60. Sydor, A.M.; Zamble, D.B. Nickel metallomics: General themes guiding nickel homeostasis. In Metallomics and the Cell. Metal Ions in Life Sciences; Banci, L., Ed.; Springer: Dordrecht, The Netherlands, 2013; Volume 12, pp. 375–416. ISBN 978-94-007-5560-4. [Google Scholar] [CrossRef]
  61. Rensing, C.; McDevitt, S.F. The Copper Metallome in Prokaryotic Cells. In Metallomics and the Cell. Metal Ions in Life Sciences; Banci, L., Ed.; Springer: Dordrecht, The Netherlands, 2013; Volume 12, pp. 417–450. ISBN 978-94-007-5560-4. [Google Scholar] [CrossRef]
  62. Vest, K.E.; Hashemi, H.F.; Cobine, P.A. The copper metallome in eukaryotic cells. In Metallomics and the Cell. Metal Ions in Life Sciences; Banci, L., Ed.; Springer: Dordrecht, The Netherlands, 2013; Volume 12, pp. 451–478. ISBN 978-94-007-5560-4. [Google Scholar] [CrossRef]
  63. Colvin, R.A.; Stork, C.J.; Li, Y.V.; Lai, B. Exploring the zinc metallome of cultured cortical neurons using synchroton radiation X-ray fluorescence microscopy. In Metal Ion in Stroke; Li, Y., Zhang, J., Eds.; Springer: New York, NY, USA, 2012; pp. 227–237. ISBN 978-1-4419-9662-6. [Google Scholar] [CrossRef]
  64. Park, J.; McCormick, S.P.; Chakrabarti, M.; Lindahl, P.A. Insights into the iron-ome and manganese-ome of Δmtm1 Saccharomyces cerevisiae mitochondria. Metallomics 2013, 5, 656–672. [Google Scholar] [CrossRef]
  65. Szpunar, J. Advances in analytical methodology for bioinorganic speciation analysis: Metallomics, metalloproteomics and heteroatom-tagged proteomics and metabolomics. Analyst 2005, 130, 442–465. [Google Scholar] [CrossRef]
  66. Shi, W.; Chance, M.R. Metallomics and metalloproteomics. Cell. Mol. Life Sci. 2008, 65, 3040–3048. [Google Scholar] [CrossRef]
  67. Zeng, X.; Cheng, Y.; Wang, C. Global mapping of metalloproteomes. Biochemistry 2021, 60, 3507–3514. [Google Scholar] [CrossRef]
  68. Wesenberg, D.; Krauss, G.J.; Schaumlöffel, D. Metallo-thiolomics: Investigation of thiol peptide regulated metal homeostasis in plants and fungi by liquid chromatography-mass spectrometry. Int. J. Mass Spectrom. 2011, 307, 46–54. [Google Scholar] [CrossRef]
  69. Schaumlöffel, D. The position of metallomics within other omics fields. In Metallomics: Analytical Techniques and Speciation Methods; Michalke, B., Ed.; Wiley-VCH Verlag: Weinheim, Germany, 2016; pp. 1–16. ISBN 978-3-527-33969-3. [Google Scholar] [CrossRef]
  70. Codd, R. Metalloglycomics: A new perspective upon competitive metal–carbohydrate binding using EPR spectroscopy. Chem. Commun. 2004, 23, 2653–2655. [Google Scholar] [CrossRef] [PubMed]
  71. Farrell, N.P.; Gorle, A.K.; Peterson, E.J.; Berners-Price, S.J. Metalloglycomics. In Metallo-Drugs: Development and Action of Anticancer Agents. Metal Ions in Life Sciences; Sigel, A., Sigel, H., Freisinger, E., Sigel, R.K.O., Eds.; Walter de Gruyter GmbH: Berlin, Germany, 2018; Volume 18, pp. 109–140. ISBN 978-3-11-046984-4. [Google Scholar] [CrossRef]
  72. Gorle, A.K.; Rajaratnam, P.; Chang, C.W.; von Itzstein, M.; Berners-Price, S.J.; Farrell, N.P. Glycans as ligands in bioinorganic chemistry. Probing the interaction of a trinuclear platinum anticancer complex with defined monosaccharide fragments of heparan sulfate. Inorg. Chem. 2019, 58, 7146–7155. [Google Scholar] [CrossRef] [PubMed]
  73. Carrillo, J.T.; Borthakur, D. Methods for metal chelation in plant homeostasis. Plant Physiol. Biochem. 2021, 163, 95–107. [Google Scholar] [CrossRef] [PubMed]
  74. Qu, Y.; Li, Y.-F.; Bai, R.; Chen, C.; Chai, Z. Introduction. In Nuclear Analytical Techniques for Metallomics and Metalloproteomics; Chen, C., Chai, Z., Gao, Y., Eds.; Royal Society of Chemistry: Cambridge, UK, 2010; pp. 1–43. ISBN 1847559018. [Google Scholar] [CrossRef]
  75. Grazul, M.; Budzisz, E. Biological activity of metal ions complexes of chromones, coumarins and flavones. Coord. Chem. Rev. 2009, 253, 2588–2598. [Google Scholar] [CrossRef]
  76. Selvaraj, S.; Krishnaswamy, S.; Devashya, V.; Sethuraman, S.; Krishnan, U.M. Flavonoid–metal ion complexes: A novel class of therapeutic agents. Med. Res. Rev. 2014, 34, 677–702. [Google Scholar] [CrossRef] [PubMed]
  77. Kasprzak, M.M.; Erxleben, A.; Ochocki, J. Properties and applications of flavonoid metal complexes. RSC Adv. 2015, 5, 45853–45877. [Google Scholar] [CrossRef]
  78. Samsonowicz, M.; Regulska, E. Spectroscopic study of molecular structure, antioxidant activity and biological effects of metal hydroxyflavonol complexes. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2017, 173, 757–771. [Google Scholar] [CrossRef] [PubMed]
  79. Uivarosi, V.; Munteanu, A. Flavonoid complexes as promising anticancer metallodrugs. In Flavonoids-from Biosynthesis to Human Health; Justino, G.C., Ed.; InTech: Rijeka, Croatia, 2017; pp. 305–337. ISBN 978-953-51-3423-7. [Google Scholar] [CrossRef]
  80. Borowska, S.; Brzoska, M.M.; Tomczyk, M. Complexation of bioelements and toxic metals by polyphenolic compounds–implications for health. Curr. Drug Targets 2018, 19, 1612–1638. [Google Scholar] [CrossRef] [PubMed]
  81. Khater, M.; Ravishankar, D.; Greco, F.; Osborn, H.M. Metal complexes of flavonoids: Their synthesis, characterization and enhanced antioxidant and anticancer activities. Future Med. Chem. 2019, 11, 2845–2867. [Google Scholar] [CrossRef] [PubMed]
  82. Malacaria, L.; Corrente, G.A.; Beneduci, A.; Furia, E.; Marino, T.; Mazzone, G. A Review on Coordination Properties of Al (III) and Fe (III) toward Natural Antioxidant Molecules: Experimental and Theoretical Insights. Molecules 2021, 26, 2603. [Google Scholar] [CrossRef] [PubMed]
  83. Rossberg, A.; Reich, T.; Bernhard, G. Complexation of uranium (VI) with protocatechuic acid—Application of iterative transformation factor analysis to EXAFS spectroscopy. Anal. Bioanal. Chem. 2003, 376, 631–638. [Google Scholar] [CrossRef] [PubMed]
  84. Cornard, J.P.; Lapouge, C.; André, E. pH influence on the complexation site of Al (III) with protocatechuic acid. A spectroscopic and theoretical approach. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2013, 108, 280–287. [Google Scholar] [CrossRef] [PubMed]
  85. Kula, A. Thermal analysis of lanthanide (III) and Y(III) complexes with 4-hydroxy-3-methoxybenzoic acid. J. Therm. Anal. Calorim. 2005, 81, 381–385. [Google Scholar] [CrossRef]
  86. Vulpius, D.; Geipel, G.; Baraniak, L.; Bernhard, G. Complex formation of neptunium (V) with 4-hydroxy-3-methoxybenzoic acid studied by time-resolved laser-induced fluorescence spectroscopy with ultra-short laser pulses. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2006, 63, 603–608. [Google Scholar] [CrossRef] [PubMed]
  87. Oke, I.M.; Ramorobi, L.M.; Mashele, S.S.; Bonnet, S.L.; Makhafola, T.J.; Eze, K.C.; Noreljaleel, A.E.M.; Chukwuma, C.I. Vanillic acid–Zn (II) complex: A novel complex with antihyperglycaemic and anti-oxidative activity. J. Pharm. Pharmacol. 2021, 73, 1703–1714. [Google Scholar] [CrossRef]
  88. Fazary, A.E.; Taha, M.; Ju, Y.H. Iron complexation studies of gallic acid. J. Chem. Eng. Data 2009, 54, 35–42. [Google Scholar] [CrossRef]
  89. Taha, M.; Khan, I.; Coutinho, J.A. Complexation and molecular modeling studies of europium (III)–gallic acid–amino acid complexes. J. Inorg. Biochem. 2016, 157, 25–33. [Google Scholar] [CrossRef]
  90. Motloung, D.M.; Mashele, S.S.; Matowane, G.R.; Swain, S.S.; Bonnet, S.L.; Noreljaleel, A.E.; Oyedemi, S.O.; Chukwuma, C.I. Synthesis, characterization, antidiabetic and antioxidative evaluation of a novel Zn (II)-gallic acid complex with multi-facet activity. J. Pharm. Pharmacol. 2020, 72, 1412–1426. [Google Scholar] [CrossRef]
  91. Frešer, F.; Hostnik, G.; Tošović, J.; Bren, U. Dependence of the Fe (II)-Gallic Acid Coordination Compound Formation Constant on the pH. Foods 2021, 10, 2689. [Google Scholar] [CrossRef]
  92. Iwan, M.; Kula, A.; Rzączyńska, Z.; Pikus, S.; Flisiuk, D.; Gomoła, M. Synthesis and properties of lanthanide (III) complexes with 4-hydroxy-3, 5-dimethoxybenzoic acid. Chem. Pap. 2007, 61, 376–382. [Google Scholar] [CrossRef]
  93. Świsłocka, R. Experimental (FT-IR, FT-Raman, 1H, 13C NMR) and theoretical study of alkali metal syringates. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2013, 111, 290–298. [Google Scholar] [CrossRef] [PubMed]
  94. Kumar, A.; Kumar, D.; Kumari, K.; Mkhize, Z.; Seru, L.K.; Bahadur, I.; Singh, P. Metal-ligand complex formation between ferrous or ferric ion with syringic acid and their anti-oxidant and anti-microbial activities: DFT and molecular docking approach. J. Mol. Liq. 2021, 322, 114872. [Google Scholar] [CrossRef]
  95. Allan, J.R.; Carson, B.R.; Gerrard, D.L.; Hoey, S. Thermal, spectral and magnetic studies of some compounds of cobalt (II), nickel (II) and copper (II) with cinnamic acid. Thermochim. Acta 1989, 154, 315–322. [Google Scholar] [CrossRef]
  96. Kalinovskaya, I.V.; Karasev, V.E.; Zadorozhnaya, A.N.; Lifar, L.I. Luminescence spectral properties of europium (III) and terbium (III) complexes with cinnamic acid. Russ. J. Coord. Chem. 2001, 27, 516–519. [Google Scholar] [CrossRef]
  97. Ferrer, E.G.; Salinas, M.V.; Correa, M.J.; Vrdoljak, F.; Williams, P.A. ALP inhibitors: Vanadyl (IV) complexes of ferulic and cinnamic acid. Z. Naturforsch. B 2005, 60, 305–311. [Google Scholar] [CrossRef]
  98. Kalinowska, M.; Świsłocka, R.; Lewandowski, W. The spectroscopic (FT-IR, FT-Raman and 1H, 13C NMR) and theoretical studies of cinnamic acid and alkali metal cinnamates. J. Mol. Struct. 2007, 834, 572–580. [Google Scholar] [CrossRef]
  99. Kalinowska, M.; Lewandowski, W.; Świsłocka, R.; Regulska, E. The FT-IR, FT-Raman, 1H and 13C NMR study on molecular structure of sodium (I), calcium (II), lanthanum (III) and thorium (IV) cinnamates. Spectroscopy 2010, 24, 277–281. [Google Scholar] [CrossRef]
  100. Kalinowska, M.; Świsłocka, R.; Lewandowski, W. Zn (II), Cd (II) and Hg (I) complexes of cinnamic acid: FT-IR, FT-Raman, 1H and 13C NMR studies. J. Mol. Struct. 2011, 993, 404–409. [Google Scholar] [CrossRef]
  101. Graminha, A.E.; Honorato, J.; Dulcey, L.L.; Godoy, L.R.; Barbosa, M.F.; Cominetti, M.R.; Menezes, A.C.; Batista, A.A. Evaluation of the biological potential of ruthenium (II) complexes with cinnamic acid. J. Inorg. Biochem. 2020, 206, 111021. [Google Scholar] [CrossRef] [PubMed]
  102. Chukwuma, C.I.; Mashele, S.S.; Swain, S.S. Antidiabetic and Antioxidative Properties of Novel Zn (II)-cinnamic Acid Complex. Med. Chem. 2021, 17, 913–925. [Google Scholar] [CrossRef] [PubMed]
  103. Świsłocka, R.; Kowczyk-Sadowy, M.; Kalinowska, M.; Lewandowski, W. Spectroscopic (FT-IR, FT-Raman, 1H and 13C NMR) and theoretical studies of p-coumaric acid and alkali metal p-coumarates. Spectroscopy 2012, 27, 35–48. [Google Scholar] [CrossRef]
  104. Koç, S.; Köse, D.A.; Avcı, E.; Köse, K. Synthesis and Thermal Characterization of p-Coumaric Acid Complexes of CoII, NiII, CuII and ZnII Metal Cations and Biological Applications. Hittite J. Sci. Eng. 2016, 3, 15–22. [Google Scholar] [CrossRef]
  105. Khvan, A.M.; Kristallovich, E.L.; Abduazimov, K.A. Complexation of caffeic and ferulic acids by transition-metal ions. Chem. Nat. Compd. 2001, 37, 72–75. [Google Scholar] [CrossRef]
  106. Fazary, A.E.; Ju, Y.H.; Al-Shihri, A.S.; Bani-Fwaz, M.Z.; Alfaifi, M.Y.; Alshehri, M.A.; Saleh, K.A.; Elbehairi, S.E.I.; Fawy, K.F.; Abd-Rabboh, H.S. Platinum and vanadate bioactive complexes of glycoside naringin and phenolates. Open Chem. 2017, 15, 189–199. [Google Scholar] [CrossRef] [Green Version]
  107. Singh, K.; Kumar, A. Kinetics of complex formation of Fe (III) with caffeic acid: Experimental and theoretical study. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2019, 211, 148–153. [Google Scholar] [CrossRef] [PubMed]
  108. Arciszewska, Ż.; Gama, S.; Kalinowska, M.; Świderski, G.; Świsłocka, R.; Gołębiewska, E.; Naumowicz, M.; Worobiczuk, M.; Cudowski, A.; Pietryczuk, A.; et al. Caffeic Acid/Eu (III) Complexes: Solution Equilibrium Studies, Structure Characterization and Biological Activity. Int. J. Mol. Sci. 2022, 23, 888. [Google Scholar] [CrossRef] [PubMed]
  109. Kalinowska, M.; Piekut, J.; Bruss, A.; Follet, C.; Sienkiewicz-Gromiuk, J.; Świsłocka, R.; Rzączyńska, Z.; Lewandowski, W. Spectroscopic (FT-IR, FT-Raman, 1H, 13C NMR, UV/VIS), Thermogravimetric and Antimicrobial Studies of Ca (II), Mn (II), Cu (II), Zn (II) and Cd (II) Complexes of Ferulic Acid. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2014, 122, 631–638. [Google Scholar] [CrossRef]
  110. Kalinowska, M.; Gołębiewska, E.; Mazur, L.; Lewandowska, H.; Pruszyński, M.; Świderski, G.; Wyrwas, M.; Pawluczuk, N.; Lewandowski, W. Crystal structure, spectroscopic characterization, antioxidant and cytotoxic activity of new Mg (II) and Mn (II)/Na (I) complexes of isoferulic acid. Materials 2021, 14, 3236. [Google Scholar] [CrossRef] [PubMed]
  111. Sengupta, P.; Pal, U.; Mondal, P.; Bose, A. Multi-spectroscopic and computational evaluation on the binding of sinapic acid and its Cu (II) complex with bovine serum albumin. Food Chem. 2019, 301, 125254. [Google Scholar] [CrossRef] [PubMed]
  112. Naso, L.G.; Valcarcel, M.; Roura-Ferrer, M.; Kortazar, D.; Salado, C.; Lezama, L.; González-Baró, A.C.; Williams, P.A.M.; Ferrer, E.G. Promising antioxidant and anticancer (human breast cancer) oxidovanadium (IV) complex of chlorogenic acid. Synthesis, characterization and spectroscopic examination on the transport mechanism with bovine serum albumin. J. Inorg. Biochem. 2014, 135, 86–99. [Google Scholar] [CrossRef] [PubMed]
  113. Kalinowska, M.; Bajko, E.; Matejczyk, M.; Kaczyński, P.; Łozowicka, B.; Lewandowski, W. The study of anti-/pro-oxidant, lipophilic, microbial and spectroscopic properties of new alkali metal salts of 5-o-caffeoylquinic acid. Int. J. Mol. Sci. 2018, 19, 463. [Google Scholar] [CrossRef] [PubMed]
  114. Kalinowska, M.; Sienkiewicz-Gromiuk, J.; Świderski, G.; Pietryczuk, A.; Cudowski, A.; Lewandowski, W. Zn (II) complex of plant phenolic chlorogenic acid: Antioxidant, antimicrobial and structural studies. Materials 2020, 13, 3745. [Google Scholar] [CrossRef] [PubMed]
  115. Palierse, E.; Przybylski, C.; Brouri, D.; Jolivalt, C.; Coradin, T. Interactions of Calcium with Chlorogenic and Rosmarinic Acids: An Experimental and Theoretical Approach. Int. J. Mol. Sci. 2020, 21, 4948. [Google Scholar] [CrossRef] [PubMed]
  116. Yang, R.; Tian, J.; Liu, Y.; Zhu, L.; Sun, J.; Meng, D.; Wang, Z.; Wang, C.; Zhou, Z.; Chen, L. Interaction mechanism of ferritin protein with chlorogenic acid and iron ion: The structure, iron redox, and polymerization evaluation. Food Chem. 2021, 349, 129144. [Google Scholar] [CrossRef]
  117. Świsłocka, R.; Regulska, E.; Karpińska, J.; Świderski, G.; Lewandowski, W. Molecular structure and antioxidant properties of alkali metal salts of rosmarinic acid. Experimental and DFT studies. Molecules 2019, 24, 2645. [Google Scholar] [CrossRef] [PubMed]
  118. Kola, A.; Hecel, A.; Lamponi, S.; Valensin, D. Novel Perspective on Alzheimer’s Disease Treatment: Rosmarinic Acid Molecular Interplay with Copper (II) and Amyloid β. Life 2020, 10, 118. [Google Scholar] [CrossRef]
  119. Świderski, G.; Jabłońska-Trypuć, A.; Kalinowska, M.; Świsłocka, R.; Karpowicz, D.; Magnuszewska, M.; Lewandowski, W. Spectroscopic, Theoretical and antioxidant study of 3D-transition metals (Co (II), Ni (II), Cu (II), Zn (II)) complexes with cichoric acid. Materials 2020, 13, 3102. [Google Scholar] [CrossRef]
  120. Manolov, I.; Kostova, I.; Netzeva, T.; Konstantinov, S.; Karaivanova, M. Cytotoxic activity of cerium complexes with coumarin derivatives. Molecular modeling of the ligands. Arch. Pharm. 2000, 333, 93–98. [Google Scholar] [CrossRef]
  121. Pi, J.; Zeng, J.; Luo, J.J.; Yang, P.H.; Cai, J.Y. Synthesis and biological evaluation of Germanium (IV)–polyphenol complexes as potential anti-cancer agents. Bioorg. Med. Chem. Lett. 2013, 23, 2902–2908. [Google Scholar] [CrossRef]
  122. Sulpizio, C.; Müller, S.T.; Zhang, Q.; Brecker, L.; Rompel, A. Synthesis, characterization, and antioxidant activity of Zn2+ and Cu2+ coordinated polyhydroxychalcone complexes. Monatsh. Chem. 2016, 147, 1871–1881. [Google Scholar] [CrossRef]
  123. Jin, G.; Zhao, Z.; Chakraborty, T.; Mandal, A.; Roy, A.; Roy, S.; Guo, Z. Decrypting the molecular mechanistic pathways delineating the chemotherapeutic potential of ruthenium-phloretin complex in colon carcinoma correlated with the oxidative status and increased apoptotic events. Oxid. Med.Cell. Longev. 2020, 2020, 7690845. [Google Scholar] [CrossRef] [PubMed]
  124. Shubina, V.S.; Shatalina, Y.V. Absorption spectroscopy study of acid-base and metal-binding properties of flavanones. J. Appl. Spectrosc. 2013, 80, 761–766. [Google Scholar] [CrossRef]
  125. Alexiou, A.D.; Decandio, C.C.; Almeida, S.D.N.; Ferreira, M.J.; Romoff, P.; Rocha, R.C. Metal-ligand coordination and antiradical activity of a trichromium (III) complex with the flavonoid naringenin. J. Coord. Chem. 2017, 70, 2148–2160. [Google Scholar] [CrossRef]
  126. Restrepo-Guerrero, A.G.; Goitia-Semenco, H.; Naso, L.G.; Rey, M.; Gonzalez, P.J.; Ferrer, E.G.; Williams, P.A. Antioxidant and Anticancer Activities and Protein Interaction of the Oxidovanadium (IV) Naringin Complex. Inorganics 2022, 10, 13. [Google Scholar] [CrossRef]
  127. Bijlsma, J.; de Bruijn, W.J.; Velikov, K.P.; Vincken, J.P. Unravelling discolouration caused by iron-flavonoid interactions: Complexation, oxidation, and formation of networks. Food Chem. 2022, 370, 131292. [Google Scholar] [CrossRef]
  128. Shi, S.; Zhang, Y.; Chen, X.; Peng, M. Investigation of flavonoids bearing different substituents on ring C and their Cu2+ complex binding with bovine serum albumin: Structure–affinity relationship aspects. J. Agric. Food Chem. 2011, 59, 10761–10769. [Google Scholar] [CrossRef]
  129. Lutoshkin, M.A.; Kuznetsov, B.N.; Levdansky, V.A. Spectrophotometric and quantum-chemical study of acid-base and complexing properties of (±)-taxifolin in aqueous solution. Heterocycl. Commun. 2017, 23, 395–400. [Google Scholar] [CrossRef]
  130. Gulcin, İ.; Alwasel, S.H. Metal Ions, Metal Chelators and Metal Chelating Assay as Antioxidant Method. Processes 2022, 10, 132. [Google Scholar] [CrossRef]
  131. Guo, Q.; Yuan, J.; Zeng, J.; He, X.; Li, D. Synthesis of dihydromyricetin–manganese (II) complex and interaction with DNA. J. Mol. Struct. 2012, 1027, 64–69. [Google Scholar] [CrossRef]
  132. Yao, Y.; Zhang, M.; He, L.; Wang, Y.; Chen, S. Evaluation of General Synthesis Procedures for Bioflavonoid–Metal Complexes in Air-Saturated Alkaline Solutions. Front. Chem. 2020, 8, 589. [Google Scholar] [CrossRef]
  133. Mucha, P.; Skoczyńska, A.; Małecka, M.; Hikisz, P.; Budzisz, E. Overview of the Antioxidant and Anti-Inflammatory Activities of Selected Plant Compounds and Their Metal Ions Complexes. Molecules 2021, 26, 4886. [Google Scholar] [CrossRef] [PubMed]
  134. Kuntić, V.S.; Malešev, D.L.; Radović, Z.V.; Kosanić, M.M.; Mioč, U.B.; Vukojević, V.B. Spectrophotometric Investigation of Uranil (II)− Rutin Complex in 70 Ethanol. J. Agric. Food Chem. 1998, 46, 5139–5142. [Google Scholar] [CrossRef]
  135. Pyrzynska, K.; Pękal, A. Flavonoids as analytical reagents. Crit. Rev. Anal. Chem. 2011, 41, 335–345. [Google Scholar] [CrossRef] [PubMed]
  136. Atabey-Ozdemir, B.; Demirkiran, O.; Yildiz, U.; Tekin, I.O.; Coban, B. Cytotoxicity and DNA binding of copper (II) and zinc (II) complexes of flavonoids: Quercitrin, myricitrin, rutin. Bulg. Chem. Commun. 2017, 49, 901–907. [Google Scholar]
  137. Catapano, M.C.; Tvrdý, V.; Karlíčková, J.; Migkos, T.; Valentová, K.; Křen, V.; Mladěnka, P. The stoichiometry of isoquercitrin complex with iron or copper is highly dependent on experimental conditions. Nutrients 2017, 9, 1193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Barbosa, V.T.; de Menezes, J.B.; Santos, J.C.C.; de Assis Bastos, M.L.; de Araújo-Júnior, J.X.; do Nascimento, T.G.; Basílio-Júnior, I.D.; Grillo, L.A.M.; Dornelas, C.B. Characterization and stability of the antimony-quercetin complex. Adv. Pharm. Bull. 2019, 9, 432. [Google Scholar] [CrossRef]
  139. Wongso, H. Natural product-based Radiopharmaceuticals: Focus on curcumin and its analogs, flavonoids, and marine peptides. J. Pharm. Anal. 2021, 12, 380–393. [Google Scholar] [CrossRef]
  140. Sahyon, H.A.; Althobaiti, F.; Ramadan, A.E.M.M.; Fathy, A.M. Quercetin-Based Rhodium (III) Complex: Synthesis, Characterization and Diverse Biological Potentials. J. Mol. Struct. 2022, 1257, 132584. [Google Scholar] [CrossRef]
  141. Lomozová, Z.; Catapano, M.C.; Hrubša, M.; Karlíčková, J.; Macáková, K.; Kučera, R.; Mladěnka, P. Chelation of iron and copper by quercetin B-ring methyl metabolites, isorhamnetin and tamarixetin, and their effect on metal-based Fenton chemistry. J. Agric. Food Chem. 2021, 69, 5926–5937. [Google Scholar] [CrossRef]
  142. Li, J.; Zhu, J.; Wu, H.; Li, W. Synthesis, in vitro, and in silico studies of fisetin and quercetin and their metal complexes as inhibitors of α-glucosidase and thrombin. J. Mol. Liq. 2022, 349, 118164. [Google Scholar] [CrossRef]
  143. Cruz, M.A.; Tovani, C.B.; Favarin, B.Z.; Soares, M.P.; Fukada, S.Y.; Ciancaglini, P.; Ramos, A.P. Synthesis of Sr–morin complex and its in vitro response: Decrease in osteoclast differentiation while sustaining osteoblast mineralization ability. J. Mater. Chem. B 2019, 7, 823–829. [Google Scholar] [CrossRef] [PubMed]
  144. Bodini, M.E.; Del Valle, M.A.; Tapia, R.; Leighton, F.; Berrios, P. Zinc catechin complexes in aprotic medium. Redox chemistry and interaction with superoxide radical anion. Polyhedron 2001, 20, 1005–1009. [Google Scholar] [CrossRef]
  145. Hynes, M.J.; Coinceanainn, M.Ó. The kinetics and mechanisms of the reaction of iron (III) with gallic acid, gallic acid methyl ester and catechin. J. Inorg. Biochem. 2001, 85, 131–142. [Google Scholar] [CrossRef]
  146. Inoue, M.B.; Inoue, M.; Fernando, Q.; Valcic, S.; Timmermann, B.N. Potentiometric and 1H NMR studies of complexation of Al3+ with (−)-epigallocatechin gallate, a major active constituent of green tea. J. Inorg. Biochem. 2002, 88, 7–13. [Google Scholar] [CrossRef]
  147. Ansari, A.A.; Sharma, R.K. Synthesis and characterization of a biologically active lanthanum (III)–catechin complex and DNA binding spectroscopic studies. Spectrosc. Lett. 2009, 42, 178–185. [Google Scholar] [CrossRef]
  148. Grzesik, M.; Namiesnik, J.; Bartosz, G.; Sadowska-Bartosz, I. Antioxidant properties of ferrous catechin complexes. Free Radic. Biol. Med. 2018, 120, S111. [Google Scholar] [CrossRef]
  149. Alasady, S.A.; Muhamad, Y.H.; Ahmed, R.S. Theoretical and Thermodynamics Studies of Complexes Formation between Natural Flavonoids and Hg (II) Ion. Syst. Rev. Pharm. 2020, 11, 2393–2404. [Google Scholar] [CrossRef]
  150. Fathima, A.; Manikandamathavan, V.M.; Jonnalagadda, R.R.; Nair, B.U. Chromium-catechin complex, synthesis and toxicity check using bacterial models. Heliyon 2020, 6, e04563. [Google Scholar] [CrossRef] [PubMed]
  151. Liu, L.; Xiao, X.; Li, K.; Li, X.; Shi, B.; Liao, X. Synthesis of Catechin-Rare Earth Complex with Efficient and Broad-Spectrum Anti-Biofilm Activity. Chem. Biodivers. 2020, 17, e1900734. [Google Scholar] [CrossRef]
  152. Navarro, R.E.; Santacruz, H.; Inoue, M. Complexation of epigallocatechin gallate (a green tea extract, egcg) with Mn2+: Nuclear spin relaxation by the paramagnetic ion. J. Inorg. Biochem. 2005, 99, 584–588. [Google Scholar] [CrossRef] [PubMed]
  153. Xie, W.; Guo, Z.; Zhao, L.; Wei, Y. Metal-phenolic networks: Facile assembled complexes for cancer theranostics. Theranostics 2021, 11, 6407. [Google Scholar] [CrossRef] [PubMed]
  154. O’Coinceanainn, M.; Astill, C.; Baderschneider, B. Coordination of aluminium with purpurogallin and theaflavin digallate. J. Inorg. Biochem. 2003, 96, 463–468. [Google Scholar] [CrossRef]
  155. O’Coinceanainn, M.; Bonnely, S.; Baderschneider, B.; Hynes, M.J. Reaction of iron (III) with theaflavin: Complexation and oxidative products. J. Inorg. Biochem. 2004, 98, 657–663. [Google Scholar] [CrossRef]
  156. Naso, L.G.; Martínez, V.R.; Ferrer, E.G.; Williams, P.A. Antimetastatic effects of VOflavonoid complexes on A549 cell line. J. Trace Elem. Med. Biol. 2021, 64, 126690. [Google Scholar] [CrossRef] [PubMed]
  157. Malacaria, L.; La Torre, C.; Furia, E.; Fazio, A.; Caroleo, M.C.; Cione, E.; Marino, T.; Plastina, P. Aluminum (III), iron (III) and copper (II) complexes of luteolin: Stability, antioxidant, and anti-inflammatory properties. J. Mol. Liq. 2022, 345, 117895. [Google Scholar] [CrossRef]
  158. Frański, R. Influence of iron redox abilities on the electrospray ionization collision induced dissociation of iron complexes with methoxylated flavonoids. Int. J. Mass Spectrom. 2019, 446, 116216. [Google Scholar] [CrossRef]
  159. Dowling, S.; Regan, F.; Hughes, H. The characterisation of structural and antioxidant properties of isoflavone metal chelates. J. Inorg. Biochem. 2010, 104, 1091–1098. [Google Scholar] [CrossRef]
  160. Fedenko, V.S. Cyanidin complexation with metal ions. Ukr. Biochem. J. 2006, 78, 149–153. [Google Scholar]
  161. Khaodee, W.; Aeungmaitrepirom, W.; Tuntulani, T. Effectively simultaneous naked-eye detection of Cu (II), Pb (II), Al (III) and Fe (III) using cyanidin extracted from red cabbage as chelating agent. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2014, 126, 98–104. [Google Scholar] [CrossRef] [PubMed]
  162. Ike, J.N.; Tyopine, A.A.; Okoye, C.O.B. Application of Cyanidin in Quantitative Estimation of Metals in Fish Samples. Am. J. Anal. Chem. 2019, 10, 621–628. [Google Scholar] [CrossRef]
  163. Torrini, F.; Renai, L.; Scarano, S.; Del Bubba, M.; Palladino, P.; Minunni, M. Colorimetric selective quantification of anthocyanins with catechol/pyrogallol moiety in edible plants upon zinc complexation. Talanta 2022, 240, 123156. [Google Scholar] [CrossRef] [PubMed]
  164. Mollaamin, F.; Mohammadian, N.T.; Najaflou, N.; Monajjemi, M. Iranian Qara Qat fruit (redcurrant) in Arasbaran forests as the resource of anthocyanin pigments in formation of [ACN-Mg2+/Al3+/Ga3+/Sn2+/Cr3+/Fe3+] chelation clusters. SN Appl. Sci. 2021, 3, 1–18. [Google Scholar] [CrossRef]
  165. Tang, P.; Giusti, M.M. Metal chelates of petunidin derivatives exhibit enhanced color and stability. Foods 2020, 9, 1426. [Google Scholar] [CrossRef] [PubMed]
  166. Andreu, G.P.; Delgado, R.; Velho, J.A.; Curti, C.; Vercesi, A.E. Iron complexing activity of mangiferin, a naturally occurring glucosylxanthone, inhibits mitochondrial lipid peroxidation induced by Fe2+-citrate. Eur. J. Pharmacol. 2005, 513, 47–55. [Google Scholar] [CrossRef] [PubMed]
  167. Nuñez-Selles, A.J.; Nuevas-Paz, L.; Martínez-Sánchez, G. Inhibition of Peroxidation Potential and Protein Oxidative Damage by Metal Mangiferin Complexes. Appl. Sci. 2022, 12, 2240. [Google Scholar] [CrossRef]
  168. Dias, K.; Nikolaou, S. Does the combination of resveratrol with Al (III) and Zn (II) improve its antioxidant activity? Nat. Prod. Commun. 2011, 6, 1673–1676. [Google Scholar] [CrossRef] [Green Version]
  169. Chiavarino, B.; Crestoni, M.E.; Fornarini, S.; Taioli, S.; Mancini, I.; Tosi, P. Infrared spectroscopy of copper-resveratrol complexes: A joint experimental and theoretical study. J. Chem. Phys. 2012, 137, 024307. [Google Scholar] [CrossRef] [PubMed]
  170. Priyadarsini, K.I. The chemistry of curcumin: From extraction to therapeutic agent. Molecules 2014, 19, 20091–20112. [Google Scholar] [CrossRef] [PubMed]
  171. Shakeri, A.; Panahi, Y.; Johnston, T.P.; Sahebkar, A. Biological properties of metal complexes of curcumin. BioFactors 2019, 45, 304–317. [Google Scholar] [CrossRef] [PubMed]
  172. Vergara, V.B.; Kalinich, J.F. Nutraceuticals as Potential Radionuclide Decorporation Agents. Nutrients 2021, 13, 2545. [Google Scholar] [CrossRef]
  173. Fucassi, F.; Heikal, A.; Mikhalovska, L.I.; Standen, G.; Allan, I.U.; Mikhalovsky, S.V.; Cragg, P.J. Metal chelation by a plant lignan, secoisolariciresinol diglucoside. J. Incl. Phenom. Macrocycl. Chem. 2014, 80, 345–351. [Google Scholar] [CrossRef]
  174. Borsari, M.; Gabbi, C.; Ghelfi, F.; Grandi, R.; Saladini, M.; Severi, S.; Borella, F. Silybin, a new iron-chelating agent. J. Inorg. Biochem. 2001, 85, 123–129. [Google Scholar] [CrossRef]
  175. Tvrdý, V.; Catapano, M.C.; Rawlik, T.; Karlíčková, J.; Biedermann, D.; Křen, V.; Mladěnka, P.; Valentová, K. Interaction of isolated silymarin flavonolignans with iron and copper. J. Inorg. Biochem. 2018, 189, 115–123. [Google Scholar] [CrossRef] [PubMed]
  176. Vimalraj, S.; Rajalakshmi, S.; Saravanan, S.; Preeth, D.R.; Vasanthi, R.L.; Shairam, M.; Chatterjee, S. Synthesis and characterization of zinc-silibinin complexes: A potential bioactive compound with angiogenic, and antibacterial activity for bone tissue engineering. Colloids Surf. B Biointerfaces 2018, 167, 134–143. [Google Scholar] [CrossRef] [PubMed]
  177. Merdy, P.; Guillon, E.; Frapart, Y.M.; Aplincourt, M. Iron and manganese surface complex formation with extracted lignin. Part 2: Characterisation of magnetic interaction between transition metal and quinonic radical by EPR microwave power saturation experiments. New J. Chem. 2003, 27, 577–582. [Google Scholar] [CrossRef]
  178. Zeng, X.; Du, Z.; Xu, Y.; Sheng, Z.; Jiang, W. Characterization of the interactions between apple condensed tannins and biologically important metal ions [Fe2+ (3d6), Cu2+ (3d9) and Zn2+ (3d10)]. LWT 2019, 114, 108384. [Google Scholar] [CrossRef]
  179. Fu, Z.; Jiang, X.; Li, W.W.; Shi, Y.; Lai, S.; Zhuang, J.; Yao, S.; Liu, Y.; Hu, J.; Gao, L.; et al. Proanthocyanidin–aluminum complexes improve aluminum resistance and detoxification of Camellia sinensis. J. Agric. Food Chem. 2020, 68, 7861–7869. [Google Scholar] [CrossRef] [PubMed]
  180. Tahara, K.; Hashida, K.; Otsuka, Y.; Ohara, S.; Kojima, K.; Shinohara, K. Identification of a hydrolyzable tannin, oenothein B, as an aluminum-detoxifying ligand in a highly aluminum-resistant tree, Eucalyptus camaldulensis. Plant Physiol. 2014, 164, 683–693. [Google Scholar] [CrossRef]
  181. Przewloka, S.R.; Shearer, B.J. The further chemistry of ellagic acid. II. Ellagic acid and water-soluble ellagates as metal precipitants. Holzforschung 2002, 56, 13–19. [Google Scholar] [CrossRef]
  182. Kraal, P.; Jansen, B.; Nierop, K.G.; Verstraten, J.M. Copper complexation by tannic acid in aqueous solution. Chemosphere 2006, 65, 2193–2198. [Google Scholar] [CrossRef] [PubMed]
  183. Guo, J.; Ping, Y.; Ejima, H.; Alt, K.; Meissner, M.; Richardson, J.J.; Yan, Y.; Peter, K.; von Elverfeldt, D.; Hagemeyer, C.E.; et al. Engineering multifunctional capsules through the assembly of metal–phenolic networks. Angew. Chem. Int. Ed. 2014, 53, 5546–5551. [Google Scholar] [CrossRef]
  184. Liu, T.; Zhang, M.; Liu, W.; Zeng, X.; Song, X.; Yang, X.; Zhang, X.; Feng, J. Metal ion/tannic acid assembly as a versatile photothermal platform in engineering multimodal nanotheranostics for advanced applications. ACS Nano 2018, 12, 3917–3927. [Google Scholar] [CrossRef] [PubMed]
  185. Fu, Z.; Chen, R. Study of Complexes of Tannic Acid with Fe (III) and Fe (II). J. Anal. Methods Chem. 2019, 2019, 3894571. [Google Scholar] [CrossRef] [PubMed]
  186. Kumar, N.; Goel, N. Phenolic acids: Natural versatile molecules with promising therapeutic applications. Biotechnol. Rep. 2019, 24, e00370. [Google Scholar] [CrossRef] [PubMed]
  187. Shahidi, F.; Zhong, Y. Measurement of antioxidant activity. J. Funct. Foods 2015, 18, 757–781. [Google Scholar] [CrossRef]
  188. Perron, N.R.; Brumaghim, J.L. A review of the antioxidant mechanisms of polyphenol compounds related to iron binding. Cell Biochem. Biophys. 2009, 53, 75–100. [Google Scholar] [CrossRef]
  189. Mladěnka, P.; Macáková, K.; Filipský, T.; Zatloukalová, L.; Jahodář, L.; Bovicelli, P.; Silvestri, I.P.; Hrdina, R.; Saso, L. In vitro analysis of iron chelating activity of flavonoids. J. Inorg. Biochem. 2011, 105, 693–701. [Google Scholar] [CrossRef]
  190. Nobahar, A.; Carlier, J.D.; Miguel, M.G.; Costa, M.C. A review of plant metabolites with metal interaction capacity: A green approach for industrial applications. BioMetals 2021, 34, 761–793. [Google Scholar] [CrossRef]
  191. Trouillas, P.; Sancho-García, J.C.; De Freitas, V.; Gierschner, J.; Otyepka, M.; Dangles, O. Stabilizing and modulating color by copigmentation: Insights from theory and experiment. Chem. Rev. 2016, 116, 4937–4982. [Google Scholar] [CrossRef]
  192. Yoshida, K.; Mihoko, M.; Kondo, T. Blue flower color development byanthocyanins: From chemical structure to cell physiology. Nat. Prod. Rep. 2009, 26, 857–964. [Google Scholar] [CrossRef]
  193. Iwashina, T. Contribution to flower colors of flavonoids including anthocyanins: A review. Nat. Prod. Commun. 2015, 10, 529–544. [Google Scholar] [CrossRef]
  194. Trunschke, J.; Lunau, K.; Pyke, G.H.; Ren, Z.X.; Wang, H. Flower color evolution and the evidence of pollinator-mediated selection. Front. Plant Sci. 2021, 12, 617851. [Google Scholar] [CrossRef] [PubMed]
  195. Fedenko, V.S.; Shemet, S.A.; Struzhko, V.S. Complexation of cyanidin with cadmium ions in solution. Ukr. Biochem. J. 2005, 77, 104–109. [Google Scholar]
  196. Fedenko, V.S. Dose effect of cyanidin interaction with lead ions in roots of maize seedlings. Ukr. Biochem. J. 2007, 79, 24–29. [Google Scholar]
  197. Hale, K.L.; McGrath, S.P.; Lombi, E.; Stack, S.M.; Terry, N.; Pickering, I.J.; George, G.N.; Pilon-Smits, E.A.H. Molybdenum sequestration in Brassica: A role for anthocyanins? Plant Physiol. 2001, 126, 1391–1402. [Google Scholar] [CrossRef] [PubMed]
  198. Hale, K.L.; Tufan, H.A.; Pickering, I.J.; George, G.N.; Terry, N.; Pilon, M.; Pilon-Smits, E.A.H. Anthocyanins facilitate tungsten accumulation in Brassica. Physiol. Plant. 2002, 116, 351–358. [Google Scholar] [CrossRef]
  199. Stoutjesdijk, P.A.; Sale, P.W.; Larkin, P.J. Possible involvement of condensed tannins in aluminium tolerance of Lotus pedunculatus. Funct. Plant Biol. 2001, 28, 1063–1074. [Google Scholar] [CrossRef]
  200. Lavid, N.; Schwartz, A.; Yarden, O.; Tel-Or, E. The involvement of polyphenols and peroxidase activities in heavy-metal accumulation by epidermal glands of the waterlily (Nymphaeaceae). Planta 2001, 212, 323–331. [Google Scholar] [CrossRef] [PubMed]
  201. Rocha, J.E.; Guedes, T.T.; Bezerra, C.F.; Costa, M.D.S.; Campina, F.F.; de Freitas, T.S.; Souza, A.K.; Souza, C.E.S.; de Matos, Y.M.L.S.; Pereira-Junior, F.N.; et al. Identification of the gallic acid mechanism of action on mercury chloride toxicity reduction using infrared spectroscopy and antioxidant assays. Int. Biodeterior. Biodegrad. 2019, 141, 24–29. [Google Scholar] [CrossRef]
  202. Wojcieszek, J.; Ruzik, L. Enzymatic extraction of copper complexes with phenolic compounds from Açaí (Euterpe oleracea Mart.) and bilberry (Vaccinium myrtillus L.) fruits. Food Anal. Methods 2016, 9, 2105–2114. [Google Scholar] [CrossRef]
  203. Kidd, P.S.; Llugany, M.; Poschenrieder, C.H.; Gunse, B.; Barcelo, J. The role of root exudates in aluminium resistance and silicon-induced amelioration of aluminium toxicity in three varieties of maize (Zea mays L.). J. Exp. Bot. 2001, 52, 1339–1352. [Google Scholar] [CrossRef] [PubMed]
  204. Tsai, H.H.; Schmidt, W. Mobilization of iron by plant-borne coumarins. Trends Plant Sci. 2017, 22, 538–548. [Google Scholar] [CrossRef] [PubMed]
  205. Tsai, H.H.; Rodríguez-Celma, J.; Lan, P.; Wu, Y.C.; Vélez-Bermúdez, I.C.; Schmidt, W. Scopoletin 8-hydroxylase-mediated fraxetin production is crucial for iron mobilization. Plant Physiol. 2018, 177, 194–207. [Google Scholar] [CrossRef] [PubMed]
  206. Antoniadis, V.; Levizou, E.; Shaheen, S.M.; Ok, Y.S.; Sebastian, A.; Baum, C.; Prasad, M.N.V.; Wenzel, W.W.; Rinklebe, J. Trace elements in the soil-plant interface: Phytoavailability, translocation, and phytoremediation—A review. Earth-Sci. Rev. 2017, 171, 621–645. [Google Scholar] [CrossRef]
  207. Kalinowska, M.; Laderiere, B.; Champagne, P.; Kowczyk-Sadowy, M.; Lewandowski, W. Mn (II), Cu (II) and Cd (II) p-coumarates: FT-IR, FT-Raman, 1H and 13C NMR and thermogravimetric studies. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2013, 103, 264–271. [Google Scholar] [CrossRef] [PubMed]
  208. Marschner’s Mineral Nutrition of Higher Plants, 3rd ed.; Marschner, H. (Ed.) Academic press: London, UK, 2011; 672p, ISBN 978-0-12384905-2. [Google Scholar]
  209. Vatansever, R.; Ozyigit, I.I.; Filiz, E. Essential and beneficial trace elements in plants, and their transport in roots: A review. Appl. Biochem. Biotechnol. 2017, 181, 464–482. [Google Scholar] [CrossRef] [PubMed]
  210. Piccolo, E.; Ceccanti, C.; Guidi, L.; Landi, M. Role of beneficial elements in plants: Implications for the photosynthetic process. Photosynthetica 2021, 59, 349–360. [Google Scholar] [CrossRef]
  211. Maret, W. The quintessence of metallomics: A harbinger of a different life science based on the periodic table of the bioelements. Metallomics 2022, 14, mfac051. [Google Scholar] [CrossRef] [PubMed]
  212. Watanabe, T.; Maejima, E.; Yoshimura, T.; Urayama, M.; Yamauchi, A.; Owadano, M.; Okada, R.; Osaki, M.; Kanayama, Y.; Shinano, T. The ionomic study of vegetable crops. PLoS ONE 2016, 11, e0160273. [Google Scholar] [CrossRef] [PubMed]
  213. Buruleanu, L.C.; Radulescu, C.; Georgescu, A.A.; Dulama, I.D.; Nicolescu, C.M.; Olteanu, L.R.; Stanescu, S.G. Chemometric assessment of the interactions between the metal contents, antioxidant activity, total phenolics, and flavonoids in mushrooms. Anal. Lett. 2019, 52, 1195–1214. [Google Scholar] [CrossRef]
  214. Tabassum, S.; Zaki, M.; Afzal, M.; Arjmand, F. New modulated design and synthesis of quercetin–CuII/ZnII–Sn2 IV scaffold as anticancer agents: In vitro DNA binding profile, DNA cleavage pathway and Topo-I activity. Dalton Trans. 2013, 42, 10029–10041. [Google Scholar] [CrossRef] [PubMed]
  215. Porkodi, J.; Raman, N. Synthesis, characterization and biological screening studies of mixed ligand complexes using flavonoids as precursors. Appl. Organomet. Chem. 2018, 32, e4030. [Google Scholar] [CrossRef]
  216. Rahim, M.A.; Björnmalm, M.; Bertleff-Zieschang, N.; Ju, Y.; Mettu, S.; Leeming, M.G.; Caruso, F. Multiligand metal–phenolic assembly from green tea infusions. ACS Appl. Mater. Interfaces 2017, 10, 7632–7639. [Google Scholar] [CrossRef] [PubMed]
  217. Kurepa, J.; Nakabayashi, R.; Paunesku, T.; Suzuki, M.; Saito, K.; Woloschak, G.E.; Smalle, J.A. Direct isolation of flavonoids from plants using ultra-small anatase TiO2 nanoparticles. Plant J. 2014, 77, 443–453. [Google Scholar] [CrossRef] [PubMed]
  218. Binkowska, I. Hesperidin: Synthesis and characterization of bioflavonoid complex. SN Appl. Sci. 2020, 2, 1–9. [Google Scholar] [CrossRef]
  219. Candela, R.G.; Lazzara, G.; Piacente, S.; Bruno, M.; Cavallaro, G.; Badalamenti, N. Conversion of Organic Dyes into Pigments: Extraction of Flavonoids from Blackberries (Rubus ulmifolius) and Stabilization. Molecules 2021, 26, 6278. [Google Scholar] [CrossRef] [PubMed]
  220. Widyasari, E.M.; Kusumawardhany, E.; Sugiharti, R.J.; Sriyani, M.E.; Marzuki, M. The Optimization Method for Synthesis of 99mTc-Rutin as Potential Radiotracer in The Development of Cancer Drugs from Flavonoid. Indones. J. Cancer Chemoprev. 2019, 10, 80–87. [Google Scholar] [CrossRef]
  221. El-Sharawy, D.M.; Khater, S.I.; Essam, H.M.; Sherif, N.H.; Hassan, H.M.; Elmaidomy, A.H. 99mTc-Luteolin: Radiolabeling, In Silico ADMET and Biological Evaluation as a Natural Tracer Tumor imaging. J. Radiat. Res. Appl. Sci. 2021, 14, 125–132. [Google Scholar] [CrossRef]
  222. Sathishkumar, P.; Gu, F.L.; Zhan, Q.; Palvannan, T.; Yusoff, A.R.M. Flavonoids mediated ‘Green’nanomaterials: A novel nanomedicine system to treat various diseases–Current trends and future perspective. Mater. Lett. 2018, 210, 26–30. [Google Scholar] [CrossRef]
  223. Pappas, A.C.; Zoidis, E.; Goliomytis, M.; Simitzis, P.E.; Sotirakoglou, K.; Charismiadou, M.A.; Nikitas, C.; Danezis, G.; Deligeorgis, S.G.; Georgiou, C.A. Elemental Metabolomics: Modulation of egg metallome with flavonoids, an exploratory study. Antioxidants 2019, 8, 361. [Google Scholar] [CrossRef]
  224. Zoidis, E.; Pappas, A.C.; Goliomytis, M.; Simitzis, P.E.; Sotirakoglou, K.; Tavrizelou, S.; Danezis, G.; Georgiou, C.A. Quercetin and Egg Metallome. Antioxidants 2021, 10, 80. [Google Scholar] [CrossRef] [PubMed]
  225. Fedenko, V.S.; Shemet, S.A. Plant metallomics: Role of phenolic chelators. In Proceedings of the 5th International Scientific Conference Restoration of Disturbed Natural Ecosystems, Donetsk, Ukraine, 12–15 May 2014; pp. 404–405. [Google Scholar]
  226. Fedenko, V.S. Phenolic chelators and plant metallomics. In Plant Physiology: Achievements and New Trends for Development; Morgun, V.V., Ed.; Logos: Kyiv, Ukraine, 2017; Volume 2, pp. 582–589. ISBN 978-616-7442-01-0. [Google Scholar]
  227. Santin, M.; Lucini, L.; Castagna, A.; Rocchetti, G.; Hauser, M.T.; Ranieri, A. Comparative “phenol-omics” and gene expression analyses in peach (Prunus persica) skin in response to different postharvest UV-B treatments. Plant Physiol. Biochem. 2019, 135, 511–519. [Google Scholar] [CrossRef] [PubMed]
  228. Gutsch, A.; Vandionant, S.; Sergeant, K.; Jozefczak, M.; Vangronsveld, J.; Hausman, J.F.; Cuypers, A. Systems biology of metal tolerance in plants: A case study on the effects of Cd exposure on two model plants. In Plant Metallomics and Functional Omics; Sablok, G., Ed.; Springer: Cham, Switzerland, 2019; pp. 23–37. ISBN 978-3-030-19103-0. [Google Scholar] [CrossRef]
  229. Guerriero, G.; Sergeant, K.; Hausman, J.F. Integrated-omics: A powerful approach to understanding the heterogeneous lignification of fibre crops. Int. J. Mol. Sci. 2013, 14, 10958–10978. [Google Scholar] [CrossRef] [PubMed]
  230. Eticha, D.; Staß, A.; Horst, W.J. Localization of aluminium in the maize root apex: Can morin detect cell wall-bound aluminium? J. Exp. Bot. 2005, 56, 1351–1357. [Google Scholar] [CrossRef] [PubMed]
  231. Zhu, X.F.; Lei, G.J.; Wang, Z.W.; Shi, Y.Z.; Braam, J.; Li, G.X.; Zheng, S.J. Coordination between apoplastic and symplastic detoxification confers plant aluminum resistance. Plant Physiol. 2013, 162, 1947–1955. [Google Scholar] [CrossRef]
  232. Gei, V.; Erskine, P.D.; Harris, H.H.; Echevarria, G.; Mesjasz-Przybyłowicz, J.; Barnabas, A.D.; Przybyłowicz, W.J.; Kopittke, P.M.; van der Ent, A. Tools for the discovery of hyperaccumulator plant species and understanding their ecophysiology. In Agromining: Farming for Metals; Van der Ent, A., Echevarria, G., Baker, A., Morel, J., Eds.; Springer: Cham, Switzerland, 2018; pp. 117–133. ISBN 978-3-319-61899-9. [Google Scholar] [CrossRef]
  233. Fedenko, V.S.; Landi, M.; Shemet, S.A. Detection of nickel in maize roots: A novel nondestructive approach by reflectance spectroscopy and colorimetric models. Ecol. Indic. 2017, 82, 463–469. [Google Scholar] [CrossRef]
  234. Phiwchai, I.; Yuensook, W.; Sawaengsiriphon, N.; Krungchanuchat, S.; Pilapong, C. Tannic acid (TA): A molecular tool for chelating and imaging labile iron. Eur. J. Pharm. Sci. 2018, 114, 64–73. [Google Scholar] [CrossRef] [PubMed]
  235. Feldmann, J.; Bluemlein, K.; Krupp, E.M.; Mueller, M.; Wood, B.A. Metallomics study in plants exposed to arsenic, mercury, selenium and sulphur. In Metallomics. Advances in Experimental Medicine and Biology; Arruda, M., Ed.; Springer: Cham, Switzerland, 2018; Volume 1055, pp. 67–100. ISBN 978-3-319-90143-5. [Google Scholar] [CrossRef]
  236. Ito, T.; Oyama, K.I.; Yoshida, K. Direct observation of hydrangea blue-complex composed of 3-O-glucosyldelphinidin, Al3+ and 5-O-acylquinic acid by ESI-mass spectrometry. Molecules 2018, 23, 1424. [Google Scholar] [CrossRef] [PubMed]
  237. Cherrak, S.A.; Mokhtari-Soulimane, N.; Berroukeche, F.; Bensenane, B.; Cherbonnel, A.; Merzouk, H.; Elhabiri, M. In vitro antioxidant versus metal ion chelating properties of flavonoids: A structure-activity investigation. PLoS ONE 2016, 11, e0165575. [Google Scholar] [CrossRef] [PubMed]
  238. Jabeen, E.; Janjua, N.K.; Ahmed, S.; Murtaza, I.; Ali, T.; Masood, N.; Rizvi, A.S.; Murtaza, G. DFT predictions, synthesis, stoichiometric structures and anti-diabetic activity of Cu (II) and Fe (III) complexes of quercetin, morin, and primuletin. J. Mol. Struct. 2017, 1150, 459–468. [Google Scholar] [CrossRef]
  239. Říha, M.; Karlíčková, J.; Filipský, T.; Macáková, K.; Rocha, L.; Bovicelli, P.; Silvestri, I.P.; Saso, L.; Jahodář, L.; Radomír Hrdina, R.; et al. In vitro evaluation of copper-chelating properties of flavonoids. RSC Adv. 2014, 4, 32628–32638. [Google Scholar] [CrossRef]
  240. Sakihama, Y.; Cohen, M.F.; Grace, S.C.; Yamasaki, H. Plant phenolic antioxidant and prooxidant activities: Phenolics-induced oxidative damage mediated by metals in plants. Toxicology 2002, 177, 67–80. [Google Scholar] [CrossRef]
  241. Eghbaliferiz, S.; Iranshahi, M. Prooxidant activity of polyphenols, flavonoids, anthocyanins and carotenoids: Updated review of mechanisms and catalyzing metals. Phytother. Res. 2016, 30, 1379–1391. [Google Scholar] [CrossRef] [PubMed]
  242. Kostyuk, V.A.; Potapovich, A.I.; Strigunova, E.N.; Kostyuk, T.V.; Afanas’ev, I.B. Experimental evidence that flavonoid metal complexes may act as mimics of superoxide dismutase. Arch. Biochem. Biophys. 2004, 428, 204–208. [Google Scholar] [CrossRef]
  243. Munteanu, A.C.; Badea, M.; Olar, R.; Silvestro, L.; Dulea, C.; Negut, C.D.; Uivarosi, V. Synthesis and structural investigation of new bio-relevant complexes of lanthanides with 5-hydroxyflavone: DNA binding and protein interaction studies. Molecules 2016, 21, 1737. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Jabeen, E.; Janjua, N.K.; Ahmed, S. Removal of metal ions using metal-flavonoid-DNA adduct protocol. J. Saudi Chem. Soc. 2019, 23, 118–126. [Google Scholar] [CrossRef]
  245. Chirug, L.; Okun, Z.; Ramon, O.; Shpigelman, A. Iron ions as mediators in pectin-flavonols interactions. Food Hydrocoll. 2018, 84, 441–449. [Google Scholar] [CrossRef]
  246. Fabjanowicz, M.; Płotka-Wasylka, J. Metals and metal-binding ligands in wine: Analytical challenges in identification. Trends Food Sci. Technol. 2021, 112, 382–390. [Google Scholar] [CrossRef]
  247. Zhao, L.; Huang, Y.; Paglia, K.; Vaniya, A.; Wancewicz, B.; Keller, A.A. Metabolomics Reveals the Molecular Mechanisms of Copper Induced Cucumber Leaf (Cucumis sativus) Senescence. Environ. Sci. Technol. 2018, 52, 7092–7100. [Google Scholar] [CrossRef]
  248. Gonzalez Ibarra, A.A.; Wrobel, K.; Yanez Barrientos, E.; Corrales Escobosa, A.R.; Gutierrez Corona, J.F.; Enciso Donis, I.; Wrobel, K. Changes of Metabolomic Profile in Helianthus annuus under Exposure to Chromium (VI) Studied by capHPLC-ESI-QTOF-MS and MS/MS. J. Anal. Methods Chem. 2017, 2017, 3568621. [Google Scholar] [CrossRef]
  249. Corso, M.; Schvartzman, M.; Guzzo, F.; Souard, F.; Malkowski, E.; Hanikenne, M.; Verbruggen, N. Contrasting cadmium resistance strategies in two metallicolous populations of Arabidopsis halleri. New Phytol. 2018, 218, 283–297. [Google Scholar] [CrossRef] [PubMed]
  250. Landi, M. Can anthocyanins be part of the metal homeostasis network in plant? Am. J. Agric. Biol. Sci. 2015, 10, 170–177. [Google Scholar] [CrossRef]
  251. Landi, M.; Pardossi, A.; Remorini, D.; Guidi, L. Antioxidant and photosynthetic response of a purple-leaved and a green-leaved cultivar of sweet basil (Ocimum basilicum) to boron excess. Environ. Exp. Bot. 2013, 85, 64–75. [Google Scholar] [CrossRef]
  252. Landi, M.; Remorini, D.; Pardossi, A.; Guidi, L. Purple versus green-leafed Ocimum basilicum: Which differences occur with regard to photosynthesis under boron toxicity? J. Plant Nutr. Soil Sci. 2013, 176, 942–951. [Google Scholar] [CrossRef]
  253. Landi, M.; Margaritopoulou, T.; Papadakis, I.E.; Araniti, F. Boron toxicity in higher plants: An update. Planta 2019, 250, 1011–1032. [Google Scholar] [CrossRef] [Green Version]
  254. Sentkowska, A.; Kilian, K.; Kopeć, M.; Pyrzyńska, K.; Cheda, Ł. Ga (III) complex with morin for kidney cancer cell labelling. Appl. Organomet. Chem. 2017, 31, e3882. [Google Scholar] [CrossRef]
  255. Jamali, A.A.; Tavakoli, A.; Dolatabadi, J.E.N. Analytical overview of DNA interaction with Morin and its metal complexes. Eur. Food Res. Technol. 2012, 235, 367–373. [Google Scholar] [CrossRef]
  256. Guzowska, M.K.; Kalinowska, M.; Lewandowski, W. Good fashion is evolution, not revolution”-methods to enhance existing anticancer medicines, primarily with use of transition metal. Anticancer Agents Med. Chem. 2017, 18, 476–487. [Google Scholar] [CrossRef] [PubMed]
  257. Williams, R.J.P.; Rickaby, R.E. Evolution’s Destiny: Co-Evolving Chemistry of the Environment and Life; Royal Society of Chemistry: Cambridge, UK, 2012; p. 319. ISBN 978-1-84973-558-2. [Google Scholar]
  258. Hanikenne, M.; Nouet, C. Metal hyperaccumulation and hypertolerance: A model for plant evolutionary genomics. Curr. Opin. Plant Biol. 2011, 14, 252–259. [Google Scholar] [CrossRef] [PubMed]
  259. Krämer, U. Conceptualizing plant systems evolution. Curr. Opin. Plant Biol. 2018, 42, 66–75. [Google Scholar] [CrossRef]
  260. Cappa, J.J.; Pilon-Smits, E.A. Evolutionary aspects of elemental hyperaccumulation. Planta 2014, 239, 267–275. [Google Scholar] [CrossRef] [PubMed]
  261. Regvar, M.; Vogel-Mikuš, K. Functional Significance of Metal Ligands in Hyperaccumulating Plants: What Do We Know? In Detoxification of Heavy Metals; Sherameti, I., Varma, A., Eds.; Springer: Berlin/Heidelberg, Germany, 2011; pp. 59–72. ISBN 978-3-642-21408-0. [Google Scholar] [CrossRef]
  262. Jalmi, S.K.; Bhagat, P.K.; Verma, D.; Noryang, S.; Tayyeba, S.; Singh, K.; Sharma, D.; Sinha, A.K. Traversing the Links between Heavy Metal Stress and Plant Signaling. Front. Plant Sci. 2018, 9, 12. [Google Scholar] [CrossRef] [PubMed]
  263. Singh, S.; Parihar, P.; Singh, R.; Singh, V.P.; Prasad, S.M. Heavy metal tolerance in plants: Role of transcriptomics, proteomics, metabolomics, and ionomics. Front. Plant Sci. 2016, 6, 1143. [Google Scholar] [CrossRef] [PubMed]
  264. Glińska, S.; Bartczak, M.; Oleksiak, S.; Wolska, A.; Gabara, B.; Posmyk, M.; Janas, K. Effects of anthocyanin-rich extract from red cabbage leaves on meristematic cells of Allium cepa L. roots treated with heavy metals. Ecotoxicol. Environ. Saf. 2007, 68, 343–350. [Google Scholar] [CrossRef] [PubMed]
  265. Stingu, A.; Volf, I.; Popa, V.I.; Gostin, I. New approaches concerning the utilization of natural amendments in cadmium phytoremediation. Ind. Crops Prod. 2012, 35, 53–60. [Google Scholar] [CrossRef]
  266. Yadav, K.K.; Gupta, N.; Kumar, A.; Reece, L.M.; Singh, N.; Rezania, S.; Khan, S.A. Mechanistic understanding and holistic approach of phytoremediation: A review on application and future prospects. Ecol. Eng. 2018, 120, 274–298. [Google Scholar] [CrossRef]
  267. Zhao, Y.H.; Jia, X.; Wang, W.K.; Liu, T.; Huang, S.P.; Yang, M.Y. Growth under elevated air temperature alters secondary metabolites in Robinia pseudoacacia L. seedlings in Cd-and Pb-contaminated soils. Sci. Total Environ. 2016, 565, 586–594. [Google Scholar] [CrossRef]
  268. Kuppusamy, S.; Thavamani, P.; Megharaj, M.; Naidu, R. Bioremediation potential of natural polyphenol rich green wastes: A review of current research and recommendations for future directions. Environ. Technol. Innov. 2015, 4, 17–28. [Google Scholar] [CrossRef]
  269. Vandenbossche, M.; Jimenez, M.; Casetta, M.; Traisnel, M. Remediation of heavy metals by biomolecules: A review. Crit. Rev. Environ. Sci. Technol. 2015, 45, 1644–1704. [Google Scholar] [CrossRef]
  270. Bacelo, H.A.; Santos, S.C.; Botelho, C.M. Tannin-based biosorbents for environmental applications—A review. Chem. Eng. J. 2016, 303, 575–587. [Google Scholar] [CrossRef]
  271. Sharma, S.; Rana, S.; Thakkar, A.; Baldi, A.; Murthy, R.S.R.; Sharma, R.K. Physical, chemical and phytoremediation technique for removal of heavy metals. J. Heavy Met. Toxic. Dis. 2016, 1, 3–10. [Google Scholar] [CrossRef]
  272. Hatami, M.; Kariman, K.; Ghorbanpour, M. Engineered nanomaterial-mediated changes in the metabolism of terrestrial plants. Sci. Total Environ. 2016, 571, 275–291. [Google Scholar] [CrossRef] [PubMed]
  273. Marslin, G.; Sheeba, C.J.; Franklin, G. Nanoparticles alter secondary metabolism in plants via ROS burst. Front. Plant Sci. 2017, 8, 832. [Google Scholar] [CrossRef] [PubMed]
  274. Ruotolo, R.; Maestri, E.; Pagano, L.; Marmiroli, M.; White, J.C.; Marmiroli, N. Plant response to metal-containing engineered nanomaterials: An omics-based perspective. Environ. Sci. Technol. 2018, 52, 2451–2467. [Google Scholar] [CrossRef]
  275. Cortez, R.; Luna-Vital, D.A.; Margulis, D.; Gonzalez de Mejia, E. Natural pigments: Stabilization methods of anthocyanins for food applications. Compr. Rev. Food Sci. Food Saf. 2017, 16, 180–198. [Google Scholar] [CrossRef]
  276. Sigurdson, G.T.; Tang, P.; Giusti, M.M. Natural colorants: Food colorants from natural sources. Annu. Rev. Food Sci. Technol. 2017, 8, 261–280. [Google Scholar] [CrossRef]
  277. Kohno, Y.; Kato, Y.; Shibata, M.; Fukuhara, C.; Maeda, Y.; Tomita, Y.; Kobayashi, K. Enhanced stability of natural anthocyanin incorporated in Fe-containing mesoporous silica. Microporous Mesoporous Mater. 2015, 203, 232–237. [Google Scholar] [CrossRef]
  278. Jung, J.; Cavender, G.; Simonsen, J.; Zhao, Y. Investigation of the mechanisms of using metal complexation and cellulose nanofiber/sodium alginate layer-by-layer coating for retaining anthocyanin pigments in thermally processed blueberries in aqueous media. J. Agric. Food Chem. 2015, 63, 3031–3038. [Google Scholar] [CrossRef]
  279. Manini, P.; Panzella, L.; Eidenberger, T.; Giarra, A.; Cerruti, P.; Trifuoggi, M.; Napolitano, A. Efficient binding of heavy metals by black sesame pigment: Toward innovative dietary strategies to prevent bioaccumulation. J. Agric. Food Chem. 2016, 64, 890–897. [Google Scholar] [CrossRef]
  280. Loizzo, M.R.; Tundis, R.; Bonesi, M.; Menichini, F.; Mastellone, V.; Avallone, L.; Menichini, F. Radical scavenging, antioxidant and metal chelating activities of Annona cherimola Mill.(cherimoya) peel and pulp in relation to their total phenolic and total flavonoid contents. J. Food Compos. Anal. 2012, 25, 179–184. [Google Scholar] [CrossRef]
  281. Islam, T.; Yu, X.; Xu, B. Phenolic profiles, antioxidant capacities and metal chelating ability of edible mushrooms commonly consumed in China. LWT-Food Sci. Technol. 2016, 72, 423–431. [Google Scholar] [CrossRef]
  282. Mladěnka, P.; Říha, M.; Martin, J.; Gorová, B.; Matějíček, A.; Spilková, J. Fruit extracts of 10 varieties of elderberry (Sambucus nigra L.) interact differently with iron and copper. Phytochem. Lett. 2016, 18, 232–238. [Google Scholar] [CrossRef]
  283. Zhai, Q.; Narbad, A.; Chen, W. Dietary strategies for the treatment of cadmium and lead toxicity. Nutrients 2014, 7, 552–571. [Google Scholar] [CrossRef]
  284. Carocho, M.; Morales, P.; Ferreira, I.C. Antioxidants: Reviewing the chemistry, food applications, legislation and role as preservatives. Trends Food Sci. Technol. 2018, 71, 107–120. [Google Scholar] [CrossRef] [Green Version]
  285. Escobar-Cévoli, R.; Castro-Espín, C.; Béraud, V.; Buckland, G.; Zamora-Ros, R.; Béraud, G.B.V. An overview of global flavonoid intake and its food sources. In Flavonoids-From Biosynthesis to Human Health; Justino, G.C., Ed.; Intech: Rijeka, Croatia, 2017; pp. 371–391. ISBN 978-953-51-3424-4. [Google Scholar] [CrossRef]
  286. Scalbert, A.; Brennan, L.; Manach, C.; Andres-Lacueva, C.; Dragsted, L.O.; Draper, J.; Rappaport, S.M.; van der Hooft, J.J.; Wishart, D.S. The food metabolome: A window over dietary exposur. Am. J. Clin. Nutr. 2014, 99, 1286–1308. [Google Scholar] [CrossRef]
  287. Olivas-Aguirre, F.J.; Rodrigo-García, J.; Martínez-Ruiz, N.D.R.; Cárdenas-Robles, A.I.; Mendoza-Díaz, S.O.; Álvarez-Parrilla, E.; González-Aguilar, G.A.; De la Rosa, L.A.; Ramos-Jiménez, A.; Wall-Medrano, A. Cyanidin-3-O-glucoside: Physical-chemistry, foodomics and health effects. Molecules 2016, 21, 1264. [Google Scholar] [CrossRef] [PubMed]
  288. Nunes, M.A.; Rodrigues, F.; Vinha, A.F.; Alves, R.C.; Oliveira, M.B.P. Nutrigenomics and polyphenols. In Polyphenols: Properties, Recovery, and Applications; Galanakis, C.M., Ed.; Woodhead Publishing: Duxford, UK, 2018; pp. 103–132. ISBN 978-0-12-813573-0. [Google Scholar] [CrossRef]
  289. Gao, R.; Hu, H.; Shi, T.; Bao, Y.; Sun, Q.; Wang, L.; Rena, Y.; Jin, W.; Yuan, L. Incorporation of gelatin and Fe2+ increases the pH-sensitivity of zein-anthocyanin complex films used for milk spoilage detection. Curr. Res. Food Sci. 2022, 5, 677–686. [Google Scholar] [CrossRef] [PubMed]
  290. Hatcher, H.C.; Singh, R.N.; Torti, F.M.; Torti, S.V. Synthetic and natural iron chelators: Therapeutic potential and clinical use. Future Med. Chem. 2009, 1, 1643–1670. [Google Scholar] [CrossRef] [PubMed]
  291. Chandra, A.; Saxena, D. Lead Toxicity and Flavonoids. In. Research Methodology in Chemical Sciences: Experimental and Theoretical Approach; Chakraborty, T., Ledwani, L., Eds.; Apple Academic Press: Oakville, ON, Canada, 2017; pp. 305–336. ISBN 978-1-4987-2860-7. [Google Scholar] [CrossRef]
  292. Brzóska, M.; Borowska, S.; Tomczyk, M. Antioxidants as a potential preventive and therapeutic strategy for cadmium. Curr. Drug Targets 2016, 17, 1350–1384. [Google Scholar] [CrossRef]
  293. Gomes de Moura, C.F.; Ribeiro, D.A. Are food compounds able to modulate noxious activities induced by cadmium exposure? Crit. Rev. Food Sci. Nutr. 2017, 57, 632–636. [Google Scholar] [CrossRef] [PubMed]
  294. Borowska, S.; Brzóska, M.M.; Gałażyn-Sidorczuk, M.; Rogalska, J. Effect of an Extract from Aronia melanocarpa L. Berries on the Body Status of Zinc and Copper under Chronic Exposure to Cadmium: An In Vivo Experimental Study. Nutrients 2017, 9, 1374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  295. Jalili-Baleh, L.; Babaei, E.; Abdpour, S.; Bukhari, S.N.A.; Foroumadi, A.; Ramazani, A.; Abdollahi, M.; Khoobi, M. A review on flavonoid-based scaffolds as multi-target-directed ligands (MTDLs) for Alzheimer’s disease. Eur. J. Med. Chem. 2018, 152, 570–589. [Google Scholar] [CrossRef] [PubMed]
  296. Rodrigues, F.; de la Luz Cádiz-Gurrea, M.; Nunes, M.A.; Pinto, D.; Vinha, A.F.; Linares, I.B.; Oliveira, M.B.P.P.; Carretero, A.S. Cosmetics. In Polyphenols: Properties, Recovery, and Applications; Galanakis, C.M., Ed.; Woodhead Publishing: Duxford, UK, 2018; pp. 393–427. ISBN 978-0-12-813573-0. [Google Scholar] [CrossRef]
  297. Giacomelli, L.; Togni, S.; Meneghin, M.; Eggenhöffner, R.; Maramaldi, G. In vivo validation of the multicomponent powder (Vitachelox®) against the deposition of polluting ions. Clin. Cosmet. Investig. Dermatol. 2018, 11, 109–113. [Google Scholar] [CrossRef] [PubMed]
  298. Rose, P.M.; Cantrill, V.; Benohoud, M.; Tidder, A.; Rayner, C.M.; Blackburn, R.S. Application of Anthocyanins from Blackcurrant (Ribes nigrum L.) Fruit Waste as Renewable Hair Dyes. J. Agric. Food Chem. 2018, 66, 6790–6798. [Google Scholar] [CrossRef]
  299. Vankar, P.S. Structure-mordant interaction, replacement by biomordants and enzymes. In Natural Dyes for Textiles: Sources, Chemistry and Applications; Woodhead Publishing: Duxford, UK, 2017; pp. 89–102. ISBN 978-0-08-101884-2. [Google Scholar] [CrossRef]
  300. Samanta, A.K.; Konar, A. Dyeing of textiles with natural dyes. In Natural Dyes; Kumbasar, E.A., Ed.; Intech: Rijeka, Croatia, 2011; pp. 29–56. ISBN 978-953-307-783-3. [Google Scholar] [CrossRef]
  301. Arora, J.; Agarwal, P.; Gupta, G. Rainbow of natural dyes on textiles using plants extracts: Sustainable and eco-friendly processes. Green Sustain. Chem. 2017, 7, 35–47. [Google Scholar] [CrossRef]
  302. Castañeda-Ovando, A.; Galán-Vidal, C.A.; Contreras-López, E.; Páez-Hernández, M. Purification of anthocyanins with o-dihydroxy arrangement by sorption in cationic resins charged with Fe (III). J. Chem. 2014, 2014, 367236. [Google Scholar] [CrossRef]
  303. Pękal, A.; Pyrzynska, K. Evaluation of aluminium complexation reaction for flavonoid content assay. Food Anal. Methods 2014, 7, 1776–1782. [Google Scholar] [CrossRef]
  304. Bernal, F.A.; Orduz-Diaz, L.L.; Coy-Barrera, E. Exploitation of the complexation reaction of ortho-dihydroxylated anthocyanins with aluminum (III) for their quantitative spectrophotometric determination in edible sources. Food Chem. 2015, 185, 84–89. [Google Scholar] [CrossRef] [PubMed]
  305. Marslin, G.; Siram, K.; Maqbool, Q.; Selvakesavan, R.K.; Kruszka, D.; Kachlicki, P.; Franklin, G. Secondary Metabolites in the Green Synthesis of Metallic Nanoparticles. Materials 2018, 11, 940. [Google Scholar] [CrossRef]
  306. Piacenza, E.; Presentato, A.; Turner, R.J. Stability of biogenic metal (loid) nanomaterials related to the colloidal stabilization theory of chemical nanostructures. Crit. Rev. Biotechnol. 2018, 38, 1–20. [Google Scholar] [CrossRef]
  307. Wan, X.; Xiang, W.; Wan, N.; Yan, S.; Bao, Z.; Wang, Y. Complexation and reduction of iron by phenolic substances: Implications for transport of dissolved Fe from peatlands to aquatic ecosystems and global iron cycling. Chem. Geol. 2018, 498, 128–138. [Google Scholar] [CrossRef]
  308. Veys-Renaux, D.; Reguer, S.; Bellot-Gurlet, L.; Mirambet, F.; Rocca, E. Conversion of steel by polyphenolic model molecules: Corrosion inhibition mechanism by rutin, esculin, esculetol. Corros. Sci. 2018, 136, 1–8. [Google Scholar] [CrossRef]
  309. Verma, C.; Ebenso, E.E.; Bahadur, I.; Quraishi, M.A. An overview on plant extracts as environmental sustainable and green corrosion inhibitors for metals and alloys in aggressive corrosive media. J. Mol. Liq. 2018, 266, 577–590. [Google Scholar] [CrossRef]
  310. Rahim, M.A.; Ejima, H.; Cho, K.L.; Kempe, K.; Müllner, M.; Best, J.P.; Caruso, F. Coordination-driven multistep assembly of metal–polyphenol films and capsules. Chem. Mater. 2014, 26, 1645–1653. [Google Scholar] [CrossRef]
  311. Halake, K.; Cho, S.; Kim, J.; Lee, T.; Cho, Y.; Chi, S.; Park, M.; Kim, K.; Lee, D.; Ju, H.; et al. Applications using the metal affinity of polyphenols with mussel-inspired chemistry. Macromol. Res. 2018, 26, 93–99. [Google Scholar] [CrossRef]
  312. Wei, J.; Wang, G.; Chen, F.; Bai, M.; Liang, Y.; Wang, H.; Zhao, D.; Zhao, Y. Sol-gel Synthesis of Metal-Phenolic Coordination Spheres and Their Derived Carbon Composites. Angew. Chem. Int. Ed. 2018, 130, 9986–9991. [Google Scholar] [CrossRef]
  313. Kim, J.H.; Oh, J.Y.; Lee, J.M.; Jeong, Y.C.; So, S.H.; Cho, Y.S.; Nam, S.; Park, C.R.; Yang, S.J. Macroscopically interconnected hierarchically porous carbon monolith by metal-phenolic coordination as an sorbent for multi-scale molecules. Carbon 2018, 126, 190–196. [Google Scholar] [CrossRef]
  314. Liao, Y.; Yao, Y.; Yu, Y.; Zeng, Y. Enhanced Antibacterial Activity of Curcumin by Combination with Metal Ions. Colloid Interface Sci. Commun. 2018, 25, 1–6. [Google Scholar] [CrossRef]
  315. Hug, H.; Bader, M.; Mair, P.; Glatzel, T. Biophotovoltaics: Natural pigments in dye-sensitized solar cells. Appl. Energy 2014, 115, 216–225. [Google Scholar] [CrossRef]
  316. Al-Alwani, M.A.; Mohamad, A.B.; Ludin, N.A.; Kadhum, A.A.H.; Sopian, K. Dye-sensitised solar cells: Development, structure, operation principles, electron kinetics, characterisation, synthesis materials and natural photosensitisers. Renew. Sustain. Energy Rev. 2016, 65, 183–213. [Google Scholar] [CrossRef]
  317. Ludin, N.A.; Mahmoud, A.A.A.; Mohamad, A.B.; Kadhum, A.A.H.; Sopian, K.; Karim, N.S.A. Review on the development of natural dye photosensitizer for dye-sensitized solar cells. Renew. Sustain. Energy Rev. 2014, 31, 386–396. [Google Scholar] [CrossRef]
  318. Kumara, N.T.R.N.; Lim, A.; Lim, C.M.; Petra, M.I.; Ekanayake, P. Recent progress and utilization of natural pigments in dye sensitized solar cells: A review. Renew. Sustain. Energy Rev. 2017, 78, 301–317. [Google Scholar] [CrossRef]
Figure 1. Structure of individual representative ligands capable of binding metal/metalloid ions from various plant phenolic subgroups.
Figure 1. Structure of individual representative ligands capable of binding metal/metalloid ions from various plant phenolic subgroups.
Ijms 23 11370 g001aIjms 23 11370 g001bIjms 23 11370 g001cIjms 23 11370 g001dIjms 23 11370 g001eIjms 23 11370 g001fIjms 23 11370 g001gIjms 23 11370 g001hIjms 23 11370 g001i
Figure 2. The elements confirmed to form phenolic ligand–Men+ complexes (highlighted in red).
Figure 2. The elements confirmed to form phenolic ligand–Men+ complexes (highlighted in red).
Ijms 23 11370 g002
Figure 3. Possible binding sites of quercetin according to [77,78].
Figure 3. Possible binding sites of quercetin according to [77,78].
Ijms 23 11370 g003
Figure 4. Research subjects of metallophenolomics.
Figure 4. Research subjects of metallophenolomics.
Ijms 23 11370 g004
Table 1. Complexes of plant phenolic ligands with metal(loid) ions.
Table 1. Complexes of plant phenolic ligands with metal(loid) ions.
Phenolic LigandMetal(loid) IonNumber of Metal IonsReferences
Phenolic acid
Hydroxybenzoic acids
Protocatechuic acid 1Al(III), U(VI)2[83,84]
Vanillic acid 2Zn(II), Y(III), La(III), Ce(III),
Pr(III), Nd(III), Sm(III), Eu(III), Gd(III), Tb(III), Dy(III), Ho(III), Er(III), Tm(III), Yb(III), Lu(III), Np(V)
17[85,86,87]
Gallic acid 3Fe(II), Zn(II), Fe(III), Eu(III)4[88,89,90,91]
Syringic acid 4Li(I), Na(I), K(I), Rb(I), Cs(I),
Fe(II), Fe(III), Y(III), La(III), Ce(III), Pr(III), Nd(III), Sm(III), Eu(III), Gd(III), Tb(III), Dy(III), Ho(III), Er(III), Tm(III), Yb(III), Lu(III)
22[92,93,94]
Hydroxycinnamic acids
Cinnamic acid 5Li(I), Na(I), K(I), Rb(I), Cs(I), Hg(I), Ca(II), Co(II), Ni(II), Cu(II), Zn(II),
Ru(II), Cd(II), La(III), Eu(III), Tb(III), VO(IV), Th(IV)
19[95,96,97,98,99,100,101,102]
p-Coumaric acid 6Li(I), Na(I), K(I), Rb(I), Cs(I), Co(II), Ni(II), Cu(II), Zn(II),Al(III)10[82,103,104]
Caffeic acid 7Li(I), Na(I), K(I), Rb(I), Cs(I), Cu(II), Pb(II), Pt(II), Al(III),
Fe(III), Cr(III), Eu(III)
12[82,105,106,107,108]
Ferulic acid 8Ca(II), Mn(II), Cu(II), Zn(II), Cd(II), Al(III), VO(IV), V(V)8[61,82,97,109]
Isoferulic acid 9Na(I), Mg(II), Mn(II)3[110]
Sinapic acid 10Cu(II), Pt(II), V(V)3[106,111]
Chlorogenic acid 11Li(I), Na(I), K(I), Rb(I), Cs(I),
Ca(II), Zn (II), Fe(III), VO(IV)
9[112,113,114,115,116]
Rosmarinic acid 12Li(I), Na(I), K(I), Rb(I), Cs(I), Ca(II), Cu(II)7[115,117,118]
Chicoric acid 13Co(II), Ni(II), Cu(II), Zn(II)4[119]
Coumarins
Coumarin 14La(III), Ce(III), Nd(III), Sm(III), Dy(III) 5[108]
Umbellipherone 15Ce(III)1[120]
Daphnetin 16Cu(II), Zn(II), Ge(IV)3[121]
Chalcones
Butein 17Cu(II), Zn(II)2[122]
Dihydrochalcones
Phloretin 18Ru(III)1[123]
Flavanones
Naringenin 19,
naringin 20
Fe(II), Cu(II), Ni(II), Zn(II), Pt(II), Fe(III), Cr(III), La(III), Y(III), Eu(III), Ce(IV), VO(IV), V(V)12[76,77,106,124,125,126]
Eriodictyol 21Fe(II), Fe(III)2[127]
Hesperitin 22,
hesperidin 23
Ni(II), Cu(II), Zn(II), Al(III), VO(IV),5[76]
Flavanonols
Taxifolin 24Fe(II), Ni(II), Cu(II), Zn(II),
Fe(III)
5[77,124,128,129,130]
Dihydromyricetin 25Mn(II), Fe(II), Co(II), Ni(II), Cu(II), Zn(II)6[131,132]
Flavonols
Kaempferol 26Fe(II), Cu(II), Zn(II), Pb(II), Fe(III), VO(IV)6[76,77,81,133]
Quercetin 27,
rutin 28,
quercitrin 29, isoquercitrin 30
Mg(II), Ca(II), Sc(II), Mn(II),
Fe(II), Co(II), Ni(II), Cu(II),
Zn(II), Mo(II), Pd(II), Cd(II),
Hg(II), Sn(II), Pb(II), Al(III),
Cr(III), Fe(III), Ga(III), Y(III), Rh(III), Sb(III), La(III), Pr(III), Nd(III), Eu(III), Gd(III), Tb(III), Dy(III), Tm(III), Au(III), Ge(IV), Zr(IV), Ru(IV), Sn(IV), Os(IV), Cr(VI), Mo(VI), W(VI), Tc(VII), Os(VIII), VO(IV), UO2(II),
43[76,77,78,81,134,135,136,137,138,139,140]
Isorhamnetin 31Fe(II), Cu(II)2[141]
Tamarixetin 32Fe(II), Cu(II)2[141]
Fisetin 33Fe(II), Cu(II), Zn(II), Fe(III), VO(IV)4[77,142]
Morin 34Mg(II), Ca(II), Mn(II), Co(II),
Ni(II), Cu(II), Zn(II), Sr(II), Pd(II),
Cd(II), Ba(II), Sn(II), Pt(II),
Al(III), Cr(III), Fe(III), Au(III), La(III), Eu(III), Gd(III), Lu(III), Zr(IV), VO(IV), Mo(VI), W(VI), Ti(COO)2 2+
26[76,77,78,143]
Myricetin 35,
myricitrin 36
Cu(II), Zn(II), Al(III), Fe(III)4[76,78,136]
Galangin 37Fe(II), Cu(II), Zn(II), Al(III)4[133]
Flavan-3-ols
(+)-Catechin 38,
(-)-epicatechin 39
Fe(II), Cu(II), Zn(II), Hg(II),
Al(III), Fe(III), Cr(III), La(III), Yb(III), Gd(III)
10[77,144,145,146,147,148,149,150,151]
(+)-Epigallocatechin 40Fe(II)1[148]
(-)-Epicatechin
3-gallate 41
Fe(II), Cu(II), Zn(II), Al(III),
Fe(III)
3[80,146]
(-)-Epigallocatechin
3-gallate 42
Fe(II), Mn(II), Cu(II), Zn(II),
Pt(II), Al(III), Fe(III)
7[80,146,148,152,153]
Theaflavin 43Al(III), Fe(III)2[154,155]
Flavones
Primuletin 44Zn(II), Cu(II); Pb(II), Al(III),
Fe(III)
5[77,133]
Chrysin 45Cu(II), Pd(II), Al(III), Fe(III), La(III), Ho(III), Er(III), Yb(III), Ce(IV), VO(IV)10[76,77,133]
Apigenin 46Cu(II), Pb(II), VO(IV)3[76,133,156]
Luteolin 47Mn(II), Fe(II), Cu(II), Al(III), Fe(III), Y(III), Ho(III), Yb(III), Lu(III), VO(IV)10[76,77,81,156,157]
Tricetin 48Fe(II), Fe(III)2[127]
Baicalein 49,
baicalin 50
Fe(II), Cu(II), Fe(III), VO(IV)4[76,77,156]
Acacetin 51Fe(III)1[158]
Isoflavones
Daidzein 52Ce(IV)1[77]
Genistein 53Cu(II), Fe(III)2[76,159]
Biochanin A 54Cu(II), Fe(III)2[159]
Anthocyanidins
Cyanidin 55 and its glycosidesCs(I), Mg(II), Ca(II), Mn(II),
Fe(II), Co(II), Ni(II), Cu(II), Sr(II),
Zn(II), Cd(II), Sn(II), Ba(II),
Hg(II), Pb(II), B(III), Al(III),
V(III), Cr(III), Fe(III), Ga(III), As(III), Bi(III), Ge(IV), VO3, MoO42−, WO42−
27[13,160,161,162,163]
Delphinidin 56 and its glycosidesMg(II), Zn(II), Sn(II), Al(III),
Cr(III), Fe(III), Ga(III)
7[13,163,164]
Petunidin 57 and its glycosidesMg(II), Sn(II), Al(III), Cr(III), Fe(III), Ga(III) 6[164,165]
Xanthonoids
Mangiferin 58Fe(II), Cu(II), Zn(II), Fe(III),
Se(IV), Ge(IV)
6[121,166,167]
Stilbenes
Resveratrol 59Fe(II), Cu(II), Zn(II), Al(III),
Fe(III)
5[80,130,168,169]
Curcuminoids
Curcumin 60Mg(II), Ca(II), Mn(II), Fe(II),
Co(II), Ni(II), Cu(II), Zn(II),
Se(II), Pd(II), Cd(II), Sn(II),
Hg(II), Pb(II), Al(III), Cr(III), Fe(III), Ga(III), Y(III), Ru(III), In(III), Re(III), Sm(III), Eu(III), Dy(III), Au(III), VO(IV), Nb(V)
28[130,170,171,172]
Lignans
Secoisolariciresinol diglucoside 61Ag(I), Ca(II), Fe(II), Ni(II), Cu(II), Pb(II)6[173]
Flavonolignans
Silibinin(silybin) 62Ni(II), Cu(II), Zn(II), Fe(III), Ga(III), VO(IV)6[156,174,175,176]
Lignins
Ligno-cellulosic substrateMn(II), Cu(II), Fe(III)3[177]
Tannins
Condensed tanninsFe(II), Cu(II), Zn(II), Al(III)4[178,179]
Oenothein B 63Al(III)1[180]
Ellagic acid 64Mg(II), Ca(II), Mn(II), Fe(II),
Co(II), Cu(II), Fe(III)
7[181]
Tannic acid 65Mg(II), Mn(II), Fe(II), Co(II),
Ni(II), Cu(II), Zn(II), Mo(II),
Cd(II), Al(III), V(III), Cr(III),
Fe(III), Ru(III), Rh(III), Ce(III), Eu(III), Gd(III), Tb(III), Ti(IV), Zr(IV)
21[153,182,183,184,185]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fedenko, V.S.; Landi, M.; Shemet, S.A. Metallophenolomics: A Novel Integrated Approach to Study Complexation of Plant Phenolics with Metal/Metalloid Ions. Int. J. Mol. Sci. 2022, 23, 11370. https://doi.org/10.3390/ijms231911370

AMA Style

Fedenko VS, Landi M, Shemet SA. Metallophenolomics: A Novel Integrated Approach to Study Complexation of Plant Phenolics with Metal/Metalloid Ions. International Journal of Molecular Sciences. 2022; 23(19):11370. https://doi.org/10.3390/ijms231911370

Chicago/Turabian Style

Fedenko, Volodymyr S., Marco Landi, and Sergiy A. Shemet. 2022. "Metallophenolomics: A Novel Integrated Approach to Study Complexation of Plant Phenolics with Metal/Metalloid Ions" International Journal of Molecular Sciences 23, no. 19: 11370. https://doi.org/10.3390/ijms231911370

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop