Next Article in Journal
Paclitaxel-Coated Versus Sirolimus-Coated Eluting Balloons for Percutaneous Coronary Interventions: Pharmacodynamic Properties, Clinical Evidence, and Future Perspectives
Next Article in Special Issue
Peptidergic G-Protein-Coupled Receptor Signaling Systems in Cancer: Examination of Receptor Structure and Signaling to Foster Innovative Pharmacological Solutions
Previous Article in Journal
Chemical Composition, Antibacterial and Antibiotic-Modifying Activity of Croton grewioides Baill Essential Oil
Previous Article in Special Issue
Pharmacogenetics and the Blood–Brain Barrier: A Whirlwind Tour of Potential Clinical Utility
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Vanadium in Metallodrugs Design and Its Interactive Profile with Protein Targets

by
Otávio Augusto Chaves
1,*,
Francisco Mainardi Martins
2,
Carlos Serpa
1 and
Davi Fernando Back
2,*
1
Department of Chemistry, Coimbra Chemistry Centre-Institute of Molecular Sciences (CQC-IMS), University of Coimbra, Rua Larga, 3004-535 Coimbra, Portugal
2
Laboratory of Inorganic Materials, Department of Chemistry, Center for Natural and Exact Sciences (CCNE), Federal University of Santa Maria (UFSM), Santa Maria 97105-900, RS, Brazil
*
Authors to whom correspondence should be addressed.
Future Pharmacol. 2024, 4(4), 743-774; https://doi.org/10.3390/futurepharmacol4040040
Submission received: 1 July 2024 / Revised: 10 October 2024 / Accepted: 17 October 2024 / Published: 24 October 2024
(This article belongs to the Special Issue Feature Papers in Future Pharmacology 2024)

Abstract

:
Metallodrugs represent a critical area of medicinal chemistry with the potential to address a wide range of diseases. Their design requires a multidisciplinary approach, combining principles of inorganic chemistry, pharmacology, and molecular biology to create effective and safe therapeutic agents. Vanadium, the element of the fifth group of the first transition series (3d metals), has been already detected as a crucial species in the biological action of some enzymes, e.g., nitrogenases and chloroperoxidase; furthermore, vanadium-based compounds have recently been described as physiologically stable with therapeutic behavior, e.g., having anticancer, antidiabetic (insulin-mimicking), antiprotozoal, antibacterial, antiviral, and inhibition of neurodegenerative disease properties. Since the binding of metallodrugs to serum albumin influences the distribution, stability, toxicity (intended and off-target interactions), and overall pharmacological properties, the biophysical characterization between serum albumin and vanadium-based compounds is one of the hot topics in pharmacology. Overall, since vanadium complexes offer new possibilities for the design of novel metallodrugs, this review summarized some up-to-date biological and medicinal aspects, highlighting proteins as the main targets for the inorganic complexes based on this transition metal.

1. Introduction: The Vanadium Element

Vanadium is an element from the fifth group of the first transition series (3d metals) with electronic configuration [Ar]3d34s2, which results in several oxidation states, although 3+, 4+ and 5+ are the most biologically relevant [1,2]. In this sense, the 4+ and 5+ oxidation states are the most explored in drug discovery due to their prodrug character in diverse biological systems [3]. These higher oxidation states are from the strong oxophilicities and Lewis’s acidity, e.g., in the species oxidovanadium (IV/V) ([VIVO]2+ and [VVO]3+, respectively) and cis-dioxidovanadium(V) ([cis-VVO2]) [3,4,5].
The vanadium element receives its name in allusion to Vanadis, a Scandinavian goddess of beauty and fertility, also referring to Freyja [6,7]. Vanadium found in the Earth’s crust, water reserves, and atmosphere has a total concentration of 135 ppm, being the twenty-first most abundant element in the external regions of planet Earth [8,9]. Despite its low concentration in the crust when compared to other elements, vanadium, in different oxidation states, is found in a very wide range of minerals presenting different oxidation states, many with economic interests (Table 1).
In marine environments, vanadium is the second most abundant transition metal, presenting concentrations of 30 to 35 nM in the form of Mn+H2VO4, where M is the metallic counterion, such as sodium. This concentration is lower than those reported with molybdenum, at an average concentration of 100 nM in the form of molybdates (MoO42−) [8,9]. Interestingly, some marine invertebrates, such as tunicates, which belong to the subphylum Tunicata (or Urochordata), also called Ascidiacea (sea squirts or marine tunicates), have an unusual capacity about accumulating vanadium cations in high concentrations. There is evidence that some of these tunicates can concentrate vanadium at up to 350 mM, levels dozens-fold higher than those found in seawater [14,15].
In ascidians, vanadium is assimilated in the pentavalent state in the form of vanadates; it is reduced to the oxidovanadium(IV) species before then being subsequently reduced to the vanadium(III) species in the form of [VIII(H2O)5HSO4]2+/[VIII(H2O)6]3+. These aquacomplexes are accumulated and stored in marine cells called vanadocytes. It is speculated that the function of vanadium in these organisms is to protect them against predators due to the possible toxicity of vanadium in high concentrations [9].
Some reducing agents can offer some type of understanding of how such redox processes can occur, such as glutathione, nicotinamide adenine dinucleotide phosphate (NADP) [16,17], tunichromes [18,19,20] and Vanabin 2 [21,22]. However, among the reducing agents described above, only a reduction from V5+ to V4+ occurs. The reduction to vanadium V3+ may be associated with reductions by organic sulfur species and sulfates [23]. This statement is based on the confirmation of the existence of high concentrations of sulfate ions and a pH close to 2 in the vacuoles of vanadocytes [23].
For humans, potable water that is rich in vanadate and diets that are rich in brown rice, beans, oats, seafood, seaweed, mushrooms, spinach, and some root vegetables, such as radishes, carrots, and beets, can lead to human body concentrations of vanadium of up to 100 mcg/day [24,25]. This daily intake increases the concentration of vanadate anions in the human bloodstream [26], and, due to the highly acidic stomach environment, there is the conversion to VO2+ and VO22+ [8,27,28]. These vanadium species heading to the intestinal environment suffer another chemical transformation to HVO42− and V2O74− (vanadate dimer), even transforming into oligovanadates due to the slightly alkaline pH [29,30,31].
The availability and chemical transformations of vanadium in the biological medium favor its high affinity with biomacromolecules such as serum albumin [32,33], transferrin [34,35], and immunoglobulin [36]. In these cases, selective receptors are sensitized, resulting in endocytosis [32,37], allowing the vanadium species to enter the cell. On the other hand, when endocytosis does not occur, vanadium can enter the cell through phosphate and sulfate channels [38] or monocarboxylate, organic anion, and mitochondrial citrate transport proteins [39,40]. As can be noticed, there are numerous ways for vanadium species to enter the cell; however, all physiological alternatives associated with interaction and intercellular displacement are related to some proteins, e.g., dicarboxylate and tricarboxylate carriers, ATP-binding cassette (ABC) transporters, and divalent metal transporters and mitochondrial citrate transport proteins [40,41,42]. In each case, the strength of the vanadium–protein interaction will be closely related to which ligand ends up donating electron density or charge for its stabilization and subsequent availability. When there is an evaluation of external factors and not just the coordination of the metal center itself, pH and redox environment changes seem to be evident regarding the release of vanadium after binding to transport proteins.

2. Vanadium in Biological Systems

In the 1970s, the biological essentiality of vanadium was confirmed, and its physiological effects began to be better evaluated [43]. At the end of the 1970s, there was a shift in interest in the bioinorganic chemistry of vanadium driven by the discovery of structural and physiological similarities between phosphate and vanadate(V) groups (Figure 1A, such vanadium species strongly inhibit sodium and potassium pumps) [43,44]. In addition, the discovery of the presence of vanadium in two groups of enzymes (bromoperoxidases, belonging to haloperoxidase enzymes, and nitrogenase enzymes) increased scientific interest in understanding the role of vanadium in biological scenarios [43].
Haloperoxidase enzymes, also known as halogenating peroxidases, are enzymes that catalyze the oxidative halogenation reactions of substrates. Basically, these enzymes mediate the oxidation of halides (X, being chlorides, bromides or iodides) by reaction with hydrogen peroxide (H2O2), in a two-electron process, to their respective hypohalides (OX), halogenate species formed in situ. Such species are released when protonated, forming the respective acids (HOX) [45,46]. The reaction involved in the oxidation of halides by H2O2 catalyzed by haloperoxidases and the subsequent halogenation of substrates are shown in Equations (1) and (2), respectively [8,45].
X + H2O2 + H3O+ → HOX + H2O   where X = Cl, Br, or I
R-H + HOX → R-X + H2O               where X = Cl, Br, or I
Haloperoxidases are classified according to two criteria, with one being based on the most electronegative species that suffer oxidation to their respective hypohalide and, consequently, the respective acid. On the other hand, the other classification is based on their prosthetic group and, consequently, their mechanism. Regarding the first criterion, there are chloroperoxidases, which can oxidize chlorides, bromides, and iodides; bromoperoxidases, which are able to oxidize bromides and iodides (and chlorides, to some extent); and, finally, iodoperoxidases, whose capacity is limited to iodine, the halogen with the highest oxidation potential and lowest electronegativity. Interestingly, there are still no reports of fluoroperoxidases due to the low oxidation potential and high electronegativity of fluorine [8,45]. Finally, regarding the second criterion, there are peroxidases that are dependent on the heme groups, i.e., porphyrin rings containing an iron metal center, and others that are dependent on vanadium.
Vanadium-dependent haloperoxidases, commonly abbreviated as VHPOs, are found in marine environments, bacteria, terrestrial fungi, and lichens [8,45]. Structurally, VHPOs present vanadium species in the form of vanadates [45,47,48]. When compared to heme-dependent ones, VHPOs exhibit greater stability with hydrogen peroxide. Curiously, chloroperoxidase obtained from the fungus Curvularia inaequalis is stable for days when exposed to at least 100 mM H2O2 and is resistant to temperatures of 70 °C, as well as the presence of organic cosolvents [45]. Thus, it can be noticed that VHPOs present structural characteristics like the enzymes from thermophilic organisms that resist high temperatures [49]; however, this chloroperoxidase is inactive at high concentrations of chloride ions [50]. The structural projection of the active site of the native chloroperoxidase of the fungus Curvularia inaequalis in a solid state obtained by single-crystal X-ray diffraction (SC-XRD) is shown in Figure 1B.
As can be seen in Figure 1B, in the active site of this native enzyme, the vanadium metal center (V) is coordinated to three non-protein oxygens in the equatorial plane and, apically, to a fourth oxygen in the position above (vanadate species). As the fifth donor atom, imidazole nitrogen from histidine residue (His-496) is coordinated underneath. Thus, in this case, vanadium appears as a trigonal bipyramid [47], a characteristic geometry of pentacoordinate vanadates in biological systems [44]. Overall, excess anionic charges, resulting from oxygen ligands, are stabilized by an extensive network of hydrogen bonds with amino acid residues in their protonated and cationic forms [51], since this enzyme works in an acidic pH range of 4.5 to 7, with 5.5 being the ideal value [49].
Chloroperoxidase from Curvularia inaequalis, as well as other enzymes belonging to this class, has the function of degrading and/or oxidizing the waxy protective layer of leaves and the cell wall of plant hosts to facilitate the microorganism access to nutrients [49]. The direct formation of halogenated metabolites, more specifically 5-chloro-4-hydroxy-3-methoxybenzaldehyde (5-chlorovanillin) and 2-chloro-4-hydroxy-3,5-dimethoxybenzaldehyde (2-chlorosyringaldehyde) (Figure 1C), from the degradation of wood lignin by the action of chloroperoxidase has already been detected and reported in the literature [49,52].
Furthermore, the same biodegradation of lignin by VHPOs leads to the formation of trichloroacetylated compounds which subsequently decay through other environmental processes (at more basic pH values), to chloroform (CHCl3). The releases of CHCl3 in atmosphere might form reactive chlorine species by photolysis, affecting the atmospheric ozone negatively [49,53]. On the other hand, VHPOs from algae and other marine organisms have the functions of regulating the amounts of cellular H2O2 and, with the formation of halogenated compounds, defense against infection by microorganisms [53].
According to both experimental and computational studies [54], Figure 1D depicts the synthesis of the HOX species by haloperoxidase. As evidenced in this figure, the mechanism of hypochlorous acid formation begins with the resting enzyme complex A, a dioxidodihydroxydovanadium(V) species. Studies, including density functional theory (DFT), suggest that hydrogen peroxide coordinates to the metallic center, forming complex B, an unstable hydroperoxidovanadium(V) dioxide(hydroxide) species. This species, in turn, reacts with the halide to form HOX acid (OH+ transfer) [54,55]. During the catalytic cycle, the oxidation state of the metal center constantly remains pentavalent (5+), with no electron transfer from/to the vanadium metal center [54,55]. Previously, through different studies, especially those from crystallographic [47], it was understood that there was the formation of an intermediate species oxidoperoxovanadium(V) (PDB code 1IDU) [47]; however, by DFT calculations, it was shown that this peroxospecies would be a stable secondary product of an alternative proton transfer channel [54,55].
Thus, the final substrate halogenation step has not yet been fully characterized, with it being suggested that the forming HOX species does not diffuse into the solution for subsequent reaction with substrate, but with the protein participating in the substrate halogenation step. Positively charged amino acid residues present below the His-496 residue and negatively charged residues on the other side of the substrate lead to the formation of a dipole moment. Positively charged amino acid residues carry a positive field, generating a gradient between vanadate and a substrate binding site. Thus, the HOX formed in the active site is diffused along the electric field vectors to the site containing the substrate. Furthermore, such vectors are aligned to the C–X bond formation. Overall, due to the presence of electrostatic interactions, dipole moments, and electric fields with favorable vectors, there is a decrease in the energetic values for the halogenation of the substrate [54].
The study, development, and application of compounds containing vanadium metal centers or other oxophilic transition metals that, inspired by VHPOs, can function as biomimetics of the catalytic activity of these enzymes are increasing [56].
Figure 1. (A) Chemical structures for phosphate, vanadate, and pervanadate. In pervanadate, the exact coordination geometry and the number and arrangement of water ligands is uncertain, and other structures may exist in the solution. (B) A 3D structure with a corresponding zoom representation of the active site of native chloroperoxidase from the fungus Curvularia inaequalis (PDB code 1IDQ) [47]. In the zoom representation, the polar amino acid residues are represented in the form of a yellow stick while vanadate (VO4) is represented as a sphere. Elements’ colors: oxygen, nitrogen, and vanadium in red, blue, and gray, respectively. For better interpretation, hydrogen atoms were omitted. (C) Chemical structure for the main products from the biodegradation of lignin by chloroperoxidase. (D) Simplified mechanism of the formation of hypochlorous/hypobromous acid by haloperoxidase-containing vanadium. Elements’ colors: oxygen, nitrogen, vanadium, chlorine, phosphorus, and bromine in red, blue, gray, green, orange, and purple, respectively.
Figure 1. (A) Chemical structures for phosphate, vanadate, and pervanadate. In pervanadate, the exact coordination geometry and the number and arrangement of water ligands is uncertain, and other structures may exist in the solution. (B) A 3D structure with a corresponding zoom representation of the active site of native chloroperoxidase from the fungus Curvularia inaequalis (PDB code 1IDQ) [47]. In the zoom representation, the polar amino acid residues are represented in the form of a yellow stick while vanadate (VO4) is represented as a sphere. Elements’ colors: oxygen, nitrogen, and vanadium in red, blue, and gray, respectively. For better interpretation, hydrogen atoms were omitted. (C) Chemical structure for the main products from the biodegradation of lignin by chloroperoxidase. (D) Simplified mechanism of the formation of hypochlorous/hypobromous acid by haloperoxidase-containing vanadium. Elements’ colors: oxygen, nitrogen, vanadium, chlorine, phosphorus, and bromine in red, blue, gray, green, orange, and purple, respectively.
Futurepharmacol 04 00040 g001

3. Natural Enzymes with Vanadium in Their Catalytic Core

The biomimetic process of the industrial Haber–Bosch synthesis—a method used to synthesize ammonia from hydrogen and nitrogen—involves nitrogenase enzymes in diazotrophic microorganisms. These enzymes are responsible for nitrogen fixation in the soil, i.e., by the transformation of an inert and biologically unusable species of nitrogen, such as dinitrogen (N2) to ammonium (NH4+), which is used in the life cycle of microorganisms [8,57,58]. Among nitrogenases, three types of isozymes are well known: Fe-nitrogenases, dependent only on iron; the Mo-nitrogenases, dependent on iron and molybdenum, and, finally, V-nitrogenases, dependent on both iron and vanadium [57,58]. Although Mo-nitrogenases are the most abundant, several diazotrophic microorganisms have genes for the expression of Fe-nitrogenases or V-nitrogenases, or even both. In this context, factors such as the low environmental concentration of molybdenum or inefficiency in its transport can lead to greater expression of these two other nitrogenases, known as “alternative nitrogenases” [57].
V-nitrogenases are found together with other nitrogenases in the soil bacteria of the genus Azobacter, in the cyanobacteria of the genera Anabaena and Nostoc, and in the lichens of the genus Peltigera [8]. As an example, the reaction involved in nitrogen fixation catalyzed by V-nitrogenases of bacterium Azobacter chroococcum is shown in Equation (3) [58].
N2 + 14H+ + 12e + 40 MgATP → 2NH4+ + 3H2 + 40MgADP + 40Pi
The biotransformation of N2 to NH4+ is a reduction that is conducted through an essential electrochemical process, using protons as a source of hydrogen and employing electrons. To achieve this, the necessary energy comes from the hydrolysis of adenosine triphosphate to its diphosphate form (ATP and ADP, respectively) [57]. Furthermore, the activity of V-nitrogenases in the activation and reductive hydrogenation of carbon dioxide (CO2) to ethylene (C2H4) and other alkenes is of particular interest to the industry [53].
V-nitrogenases have two components: one Fe-protein and one FeV-protein. The Fe-protein consists of α2-dimer containing a [4Fe-4S] cluster and two sites for interaction with ATP. The second component, the FeV-protein, consists of α2β2γ2-heterohexamer with a P-cluster [8Fe-7S] and a cluster of vanadium and iron as the cofactor (FeV-co) [57,58]. From a mechanistic point of view, the electrons necessary for the reduction reactions migrate from the [4Fe-4S] cluster (Fe-protein), where the hydrolysis of ATP to ADP occurs, towards the P-cluster and, finally, to the FeV-co (VFe-protein), where electrons and protons are accumulated for substrate reduction, which occurs in this same cofactor [57,58]. It is important to highlight that the catalytic cofactor FeV-co is a cluster consisting of [V-7Fe-8S-C-(R)-homocitrate], with a bridged carbonate anion coordinated to two iron atoms [57,58]. The structural projections of the FeV-cofactor from the bacterium Azobacter vinelandii in a solid state obtained by SC-XRD, in a non-catalytic process and under catalysis, are shown in Figure 2.
As can be seen from Figure 2, the vanadium metal center appears as hexacoordinated, being coordinated with three sulfur atoms, namely the imidazole nitrogen of histidine residue and two oxygens of a 3-hydroxy-3-carboxy-adipic acid ((R)-homocitrate, HCA) molecule, one of which being alcoholate and the other being carboxylate [58]. From studies with FeV-co reduced by dithionate, it is assumed that, in a resting state, metallic atoms present the oxidation state [VIII, 4FeIII, 3FeII]. It is also suggested that the vanadium center presents a trivalent oxidation state that is unchanged during the catalytic cycle [60]. In the center of this cofactor, there is a carbide linked to six iron atoms [58,61].

4. Bioaccumulation of Vanadium by Terrestrial Organisms

Despite vanadium having an essential role in the biological function of the enzymes haloperoxidases and nitrogenases, some organisms might accumulate vanadium for non-obvious reasons that have not yet been fully comprehended, e.g., Polychaeta marine annelids and marine ascidians, such as Ascidia gemmata. Vanadium is concentrated (bioaccumulation) by mushrooms of the genus Amanita (A.), e.g., A. regalis, A. velatipes, and A. muscaria. In the last species cited, the content of this metal is higher than four hundred times the contents found in mushrooms of the same genus. In 1972, the biomolecule in the form of which vanadium is accumulated by these mushrooms was isolated and identified as amavadin [62]. In 1999, the solid-state structure of amavadin complexed with calcium(II) coordinated to the ligand was elucidated by SC-XRD (Figure 3) [63,64].
The amavadin complex has the formula [Δ-VIV(S,S)-HIDPA)2]2−, which comprehends two units of the pro-ligand (S,S)-N-hydroxyimino-(2,2′)-dipropionic acid in a trianionic form coordinated to a non-oxido vanadium(IV) metal center, i.e., without the presence of a coordinated oxido ligand. It results in the coordination of vanadium(IV) with two nitrogens (1.982(8) Å and 1.999(8) Å) and two oxygens (1.940(7) Å and 1.956(7) Å) of the hydroxyimino functions (C –N(O)–C), as well as four carboxylate oxygens (2.028(7) Å, 2.028(9) Å, 2.042(8) Å and 2.070(8) Å) [62,63,64]. The geometry of the coordination sphere of the vanadium metal center in amavadin leads to a chirality in the metal center. Only the delta (Δ) species were obtained and evaluated crystallographically [64]. From a chemical point of view, this complex is stable in air, having a bluish color in aqueous solution, while, in some organic solvents, e.g., dimethylsulfoxide, N,N-dimethylformamide, and acetone, its metal center is oxidized to vanadium(V), resulting in a reddish solution [63].

5. Vanadium Salts, Complexes, Prodrugs, and Metallodrugs

The discovery of the cytotoxicity of cis-diamindichloretoplatinum(II) (cisplatin) to solid tumors by Rosenberg at the end of 1960s resulted in increasing interest in the field of inorganic medicinal chemistry—the discovery of metallodrugs. Metallodrugs are often prodrugs, in some cases pharmaceutical agents, that contain metal ions [65]. In the form of prodrugs, the metallodrugs might be converted into an active compound by ligand substitution and/or redox reactions in the biological matrix. Mainly due to their inorganic compounds, which might form different geometries with different coordination numbers, their unique mechanism of action has been explored in the treatment of a variety of diseases, disorders, and infections, with anticancer, anti-inflammatory, antibacterial, antifungal, and antiviral elements being subjects of interest [66,67].
Recently, several reports have highlighted the importance of vanadium-based compounds as novel prodrugs and metallodrugs with antiprotozoal, antibacterial, antiviral, and antineoplastic therapeutic actions based on in vitro and in vivo assays [68,69,70,71]. In this sense, high importance is given to the elucidation of the mechanisms that vanadium-based compounds suffer until they reach the targets, e.g., the exchange of ligands and the study of redox chemistry under physiological conditions, as well as their mechanisms of action and interactive profiles with biomacromolecules [43].
Over the past few years, the study of vanadium(III), oxidovanadium(IV), and dioxidovanadium(V) complexes with ligands containing oxygen, nitrogen, and sulfur as electron donor atoms have been explored as antineoplastic elements that have prompted advances and perspectives in regard to clinical applications [43]. Additionally, different authors have been reporting the interactive profile between vanadium salts or vanadium complexes and deoxyribonucleic acid (DNA), as well as, in some cases, extending these biophysical characterizations to the serum albumins [43,72,73].
It is important to highlight that most of the biological reports of vanadium-based compounds focus on the way that they might have mimetic activity in terms of biologically active molecules or ions such as phosphate. Vanadium derivatives, e.g., vanadates, have a high biological similarity to phosphate, mainly due to their similarity in the acidity constant (pKa). The pKa values for the phosphate species H2PO4, HPO42−, and PO43− are 2.1, 7.2, and 12.5, respectively [74], while for the vanadates H2VO4, HVO42−, and VO43−, they are 3.5, 7.8, and 12.7, respectively [75,76]. These values influence the solubility, bioavailability, and pharmacokinetics of the cited ions in the biological environment.
The mimetic activity of vanadium-based compounds can be harmful when the stability of the interaction between enzymes and vanadium is very high [73,77,78,79]. On the other hand, when the biologically active compounds are enzymes with vanadium in their catalytic core, a strong affinity, mainly promoted by hydrogen bond interactions between the inorganic complex and the amino acid residues, is required (driven by the electronic characteristics of the ligands coordinated at the metallic center) [73,80,81,82].
Several vanadium complexes (especially vanadium(IV) complexes, vanadate, oxidovanadium sulfate, and metavanadate) have insulin-mimicking actions [83,84,85,86,87]. Despite these actions having been known for more than a century [88,89], the reasons for them are not quite clear and involve several possibilities, including the phosphorylation of the insulin receptor [19,90,91,92] and activation of the PI3K/AKT pathway [68,93,94,95], which promotes the translocation of glucose transporters (GLUT4) to the cell membrane. This might increase the glucose uptake by muscle and adipose cells via mitogen-activated protein kinases (MAPK), influencing gene expression and cell therapy in increasing expressions of phosphoenolpyruvate carboxykinase (PEPCK) [95,96] and glucose-6-phosphatase (G6Pase) [96,97,98], which are critical for hepatic gluconeogenesis.
A highly relevant factor that must be highlighted is the extreme complexity that vanadium metallodrugs have demonstrated for in vivo assays, e.g., the potential action of reversing peripheral insulin resistance was reported, with it being observed that is possible to activate the MAPK/ERK signaling pathway [99,100,101]. It might generate disorder and differentiation of osteoblasts, which promotes the activity of these cells, resulting in the synthesis of the bone matrix and mineralization.

6. Vanadium Complexes with Antidiabetic Activities

Diabetes mellitus, one of the most common metabolic disorders, is characterized by problems in the metabolism of lipids, proteins, and carbohydrates, leading to a high concentration of sugars in the bloodstream. This disorder results from the autoimmune destruction of β-cells in the pancreas, which leads to little/no production of the hormone insulin (type I) or resistance of the cells to the action of secreted insulin (type II) [102]. Diabetes type II accounts for about 90% to 95% of all diagnosed cases of diabetes [103].
Several vanadium complexes have insulin-mimicking and insulin-like actions [83,84,85,86,87]. An advantage of the application of vanadium salts or complexes is their capacity to decrease glucose levels in hyperglycemic conditions rather than in normoglycemic conditions, making, in this case, hypoglycemic conditions not a concern [104]. In 1899, there was a report by Lyonnet and coworkers [105] about the administration of sodium metavanadate (NaVVO3, Figure 4) in patients with different health problems, including three with diabetes, who reported a slight reduction in glycemic levels [89,106]. In 1985, scientific interest in vanadium compounds for treating diabetes was boosted by an in vivo study reported by Heyliger and coworkers [107]. In this case, the authors orally administered sodium orthovanadate (Na3VVO4, Figure 4) in female Wistar rats in the presence of streptozotocin (STZ, a widely used to induce diabetes). The NaVVO4 administration led to a reduction in the high serum glycemic levels.
Regarding vanadium derivatives, a widely evaluated possibility was oxidovanadium(IV) sulfate, commonly also known as vanadyl sulfate (VIVOSO4, Figure 4). Several tests on the capacity of VIVOSO4 to decrease glucose levels in humans have been carried out. Such vanadium compounds have been shown to decrease glycemic levels as well as increase the sensitivity of insulin receptors. Thus, VIVOSO4 can increase the effectiveness of administered insulin [38,108,109], and its toxicity was reported to be 6- to 10-fold lower than that of vanadate [109], although it is poorly absorbed in the gastrointestinal tract (1–10%) [110]. Unfortunately, VIVOSO4 leads to undesirable gastrointestinal side effects [109,110].
Compared to vanadium salts, vanadium-based complexes, whose vanadium species are coordinated with appropriate organic ligands, may present better efficacy and tissue absorption, as well as a reduction in side effects caused by such metals [109]. Among the complexes with insulin mimetic activity, bis(maltolato)oxidovanadium(IV) (BMOV, Figure 4) [111] stands out—a neutral and water-soluble complex that presents two monoanionic units of the maltol ligand (O,O donor atom set) coordinates to the oxidovanadium(IV) center. Maltol, a food additive used in confectionery and bakeries [112], does not raise major concerns regarding its toxicity in this context [113]. An in vivo study using male Wistar rats with STZ-induced diabetes orally administered BMOV normalized blood glycemic levels without increasing insulin levels. Furthermore, when compared to VIVOSO4 [110] evaluated in a previous study [114], the reduction in glycemia occurred at a lower dosage and time (24 h vs. 1 to 2 weeks) [110,113].
Two inorganic complexes analogs to BMOV with variation in the aliphatic chain of the maltolate ligand: bis(ethylmaltolato)oxidovanadium(IV) (BEOV, Figure 4) and bis(isopropylmaltolato)oxidovanadium(IV) (BIOV, Figure 4), were administered orally and by intraperitoneal injection, respectively. These two analogs had similar insulin mimetic activities compared with BMOV (methyl derivative) [115]; however, BEOV had lower toxicity than BMOV [113]. It is important to highlight that BEOV had a positive assessment in the initial clinical phase (phase I—assess toxicity in healthy humans) and was accepted for the continuity of clinical assays to the intermediate phase (phase II—minimum effective dose assessment). In phase I, pharmacokinetic, bioavailability, and, especially, safety and tolerability parameters were evaluated in 40 volunteers without diabetes who were administered increasing doses of 10–90 mg orally per day. At this stage, no adverse health problems were observed in any of the volunteers, and better uptake and bioavailability were reported than in the positive control (VIVOSO4) [89,115]. In phase II, in 2007, seven patients with diabetes type II received 20 mg of BEOV orally per day, reducing blood glycemic levels. However, in 2009, the trials were suspended due to the reported kidney problems in patients undergoing treatment [38,88].
There are several discussions in the literature about whether it is vanadium in its form (V) or form (IV) (or both) that are responsible for the bioactivity as an insulin mimic. Shechter and coworkers [116] suggest that both forms present bioactivity, with the mechanism varying between them, i.e., tetravalent species are active in the plasma membrane, facilitating glucose uptake and, possibly, inhibiting lipolysis, while pentavalent species act only in the cytosol, promoting increased glucose levels and fat metabolism [116,117]. Furthermore, tetracoordinated vanadium acts as a substrate for enzymes that convert organic phosphates due to the similarities between both species, following the same rule re-ported with vanadates. However, pentacoordinated vanadium acts as a transition-state analog for phosphoester hydrolytic enzymes [104].
The biotransformations and chemical speciation of VVOSO4 and BMOV were studied by X-ray absorption near-edge structure (XANES), a X-ray absorption spectroscopy (XAS) mode, in cell cultures of human hepatoma (HepG1), human lung carcinoma (A549), and mouse adipocytes and preadipocytes (3T3-L1), as well as the corresponding cell culture media. It was evidenced that, in the cell culture media, both compounds undergo changes, resulting in similar mixtures of predominantly five- and six coordinate vanadium(V) species. These formed species were metabolized by HepG2 cells or mature 3T3-L1 adipocytes, resulting in the reduction of pentavalent to tetravalent vanadium as well as oxidation in the reverse direction. In that work, the capacity of the evaluated mammalian cells to generate and maintain VV species when treated with VIV complexes (like BMOV) was highlighted. Furthermore, the subcellular fractionation of A549 cells shows that the reduction in vanadium(V) species occurs mainly in the cell cytoplasm [118].
Despite the insulin mimic capacity of vanadium salts and vanadium-based complexes having been known for more than a century [88,89], the reasons for it are not quite clear and involve some possibilities, including the phosphorylation of the insulin receptor [19,90,91,92] and activation of the PI3K/AKT pathway [68,93,94,95] that promotes the translocation of glucose transporters (GLUT4) to the cell membrane. This might increase the glucose uptake by muscle and adipose cells via MAPK, influencing gene expression and cell therapy in increasing expressions of phosphoenolpyruvate carboxykinase (PEPCK) [95,96] and glucose-6-phosphatase (G6Pase) [96,97,98], which are critical for hepatic gluconeogenesis.

7. Vanadium Complexes with Anti-Inflammatory Action in Neurodegenerative Diseases

Several vanadium complexes that can act as metallodrugs have increasingly demonstrated action on different molecular targets, representing new advances for medicinal chemistry [40]. Several articles in recent years have demonstrated that vanadium complexes play a role in modulating biological processes such as enzyme inhibition [119], modulation of oxidative stress [120,121,122], and regulation of the immune response [123]. These complexes have potential as inhibitors of specific enzymes involved in disease progression, activators, or suppressors of signaling pathways for cellular function, and even as agents that mimic or modulate natural biochemical processes [124].
The therapeutic repertoire of vanadium complexes is dependent on biologically active targets, as, in addition to increasing efficacy and specificity, they also need to evaluate the potential for toxicity and side effects when compared to traditional therapeutic approaches [83,84,85,86,87]. As new vanadium complexes are synthesized and biological tests are evaluated, possible applications in specific therapies are being revealed; however, the optimization of the chemical structure of vanadium complexes (labile groups, organic ligands, and redox potentials) is essential in regard to improving its bioavailability, and possible toxic effects or multiple actions become crucial for efficiency and increased therapeutic results [125,126,127,128,129,130,131,132,133,134,135].
In addition to the best-known activities, such as insulin mimetic [83,84], DNA/HSA interaction [125,135], and anticarcinogenic activity [124], some targets of vanadium-based complexes have been studied with the aim of preventing chronic neuroinflammation, which is directly related to the progression of Alzheimer’s disease (AD) [136,137] as well as with amyloidosis-related diseases. In the last case, Xu and coworkers [136] reported the synthesis and characterization of three oxidovanadium complexes ((NH4)[VO(O2)2(bipy)]·4H2O, bis(ethylmaltolato)oxidovanadium(IV), and (bipyH2)H2[O{VO(O2)(bipy)}2]·5H2O, Figure 5), identifying them as potential inhibitors of aggregation of human islet amyloid polypeptide (hIAPP) with distinct thermodynamic behavior and reductions in the production of oligomers (corresponding IC50 values of 33.0, 14.7, and 8.50 μM). Xu and coworkers [121] reported the replacement of the bromo to nitro group in the corresponding ligands 2-(5-bromo-2-hydroxylbenzylideneamino) with benzoic acid (H2bhbb) and 2-(5-nitro-2-hydroxylbenzylideneamino) benzoic acid (H2nhbb) complexed with oxidovanadium(IV) (VO(bhbb)·H2O and VO(nhbb)·H2O, respectively, Figure 5) that might decrease the inhibitory effect on the aggregation. Interestingly, comparing BEOV with BMOV (Figure 4), the replacement of ethyl with methyl moieties [90,136] increases the inhibitory capacity of aggregation about 3-fold, while the replacement of 2,2′-bipyridine (bipy) with 1,10-phenanthroline (phen, in (NH4)[VO(O2)2(phen)]·2H2O, Figure 5) also increased the inhibitory capacity of aggregation 5-fold, indicating that not only is the nature of the vanadium species important in the development of metallodrugs for amyloidosis-related diseases; the chemical groups in the ligand complexed with the metallic center are significant as well.
In AD, it is necessary to highlight the immunological cells of the central nervous system (CNS), called microglia [138], which are responsible for maintaining cerebral homeostasis and also for responding to injuries or diseases, allowing for assessment and/or responses to disturbances in the brain, such as cellular changes, infections or anomalous concentrations of proteins. One of the most peculiar characteristics of the microglia is the possibility of switching between pro-inflammatory (M1) [139,140] and anti-inflammatory or reparative (M2) [139] phenotypes which, in promoting tissue repair, the resolution of inflammation, or under pathological conditions, can be closely related to neurodegenerative diseases (Alzheimer’s, Parkinson’s) and CNS injuries (stroke) [139,140,141].
Specifically in AD, microglia [142,143,144,145,146] undergo overactivation, releasing pro-inflammatory cytokines, such as interleukin-1 beta (IL-1β) [147], interleukin-6 (IL-6) [148,149] and tumor necrosis factor alpha (TNF-α) [150,151,152,153], increasing neuroinflammation and further damaging neurons. Reports in the literature show that vanadium complexes and new materials derived from vanadium have the capacity to modulate microglia, inhibiting overactivation and decreasing the production and release of harmful pro-inflammatory cytokines. This modulation is most likely associated with signaling pathways or transcription factors that regulate microglial activation states or promote a change towards a more neuroprotective phenotype.
The use of vanadium-based complexes and other metallodrugs can affect a specific point or a series of distinct functions, so, in addition to microglia, the nuclear factor-kappa B (NF-κB) [154,155] pathway is a target. This factor is linked to inhibitory proteins known as inhibitors of kappa B (IκBs), which, after activation by various stimuli (cytokines, pathogens or stress) trigger phosphorylation.
Another group that may be associated with degenerative diseases are the MAPKs [68,154]; however, this is another large family of kinases that involves those regulated by extracellular signal (ERK), N-terminal c-Jun (JNK), and p38 MAPK, which are activated in response to various stimuli such as cytokines, growth factors, and cellular stressors [38,156,157,158].

8. Vanadium Complexes as Inhibitors of Protein Tyrosine Phosphatases (PTPs)

The protein tyrosine phosphatases (PTPs) have a rich and wide range of functions, especially catalytic hydrolysis of the phosphate function of tyrosine residues in proteins [159,160]. This process is vital for metabolism and immunological responses [159,160,161]. Structurally, these phosphatases have catalytic sites containing active sites called PTP loops [162,163,164], which present cysteine amino acid residues involved in the catalytic activity [165]. The amino acid residue cysteine is important because it allows for reversible oxidation–reduction cycles, leading to the formation of sulfenic acid as an intermediate and, subsequently, the return to the thiol form [164,165]. The catalytic mechanism involves the nucleophilic attack of this cysteine residue on the phosphate group of a tyrosine-phosphorylated substrate, resulting in the formation of a covalent enzyme–phosphate intermediate [165,166]. Unwantedly, irreversible oxidation of this sulfenic acid leads to the production of sulfinic and sulfonic acids [167].
Although PTPs demonstrate great specificity for substrates, deregulation (leading to phosphorylation) of their activity, interrupting normal cellular signaling pathways, either due to the mutations in regulatory domains or altered interactions with substrates, contributes to diseases such as cancer, metabolic diseases, and even autoimmune diseases [168]. Regarding these issues, where the control and/or inhibition of PTPs [169,170,171,172] is necessary, vanadium complexes have shown numerous possibilities for future applications. This is due to their structural conditions (vanadium cations in the presence of specific organic molecules) [169,170,171,172,173], the geometry of the inorganic complex (it is believed that the protein stabilizes the vanadium geometry more effectively, generally the trigonal bipyramidal) [168] and redox potential (common oxidation number variation of V3+, V4+, and V5+) to mimic the transition state [173,174,175,176,177] of substrate binding or to selectively disrupt interactions within the catalytic site. Thus, the vanadium complexes act as inhibitors and can compete with the phosphate substrate or bind allosterically [169,177,178] to regulatory domains that control enzymatic activity.
A vanadium complex evaluated for its interactions with PTPs was the potential antidiabetic compound BMOV. Using the crystal-soaking technique of protein tyrosine phosphatase 1B (PTP1B) C215S to trap mutant crystals in BMOV solution, it was possible to evaluate such interactions by SC-XRD. The presence of vanadium in the form of vanadate in the enzyme was observed through this technique, which allowed for the conclusion that this uncoordinated vanadium species is the active form of the BMOV complex. This vanadium species presents itself as a trigonal bipyramid containing a hydroxy group of serine residue in an apical position and participates in a complex network of hydrogen bonds for its stabilization [179].
Some of the most investigated derivatives of vanadium are the simple salts sodium orthovanadate (Na3VVO4, Figure 4) and sodium metavanadate (NaVVO3, Figure 4), which have been found to competitively inhibit tyrosinase phosphatases by mimicking the transition state of phosphate hydrolysis, disrupting oncogenic signaling pathways that depend on anomalous tyrosine phosphorylation and offering, as an example, potential therapeutic benefits in the treatment of cancer [169].
One of the great successes in vanadium chemistry is the structure–activity characteristic that the complexes present, since the coordination with the Lewis bases (ligands) is quite varied, depending on their oxidation state, allowing for coordination to hard, intermediate centers and soft centers, such as bases containing oxygen, nitrogen, and sulfur, respectively [170,173,176,177]. This rich variation in coordination with ligands associated with oxidation–reduction potential allows the complexes to have high inhibitory potency and selectivity for specific phosphatase isoforms [180]. In addition, they also exhibit antioxidant properties, which would allow for the treatment of diseases associated with oxidative stress [169,170,171,172,176,177].

9. Vanadium Complexes as Anticarcinogenic Action

As briefly described in Section 5, metallodrugs and their biological activity compounds, especially anticarcinogens, remind us essentially of drugs and prodrugs derived from platinum, such as cisplatin. Cisplatin was the first drug derived from a platinum(II) used for antineoplastic purposes, receiving approval from the United States Food and Drug Administration (US FDA) in 1978 [181].
Cisplatin is still widely used as a first option for the treatment of carcinomas and sarcomas or as a second option when the tumor is resistant to the cocktail initially used [182]. More specifically, cisplatin is used mainly in the treatment of ovarian carcinomas and testicular teratomas and is also used in various types of head and neck, colon, bladder, esophageal, and gastric neoplasms [183]. Cisplatin and its derivatives, also based on platinum(II), cause collateral damage and resistance processes, even though they are still used today [181,183]. For this reason, research with vanadium complexes has gained attention, as they decrease the side effects and increase the diversification of targets that the inorganic complexes might use.
Vanadium complexes can affect several cells signaling pathways that are frequently deregulated in cancer cells, especially the PI3K/AKT and MAPK pathways [183,184], which are involved in cell proliferation and differentiation. It is also known that vanadium derivatives have the potential to alter the expression of proteins that regulate apoptosis through the activation of caspases such as caspase-3 and caspase-9 [93,185] or through increasing the level of some proteins, such as apoptotic Bax [186,187]. Modulation of the increase in Bax causes an inversion of the concentration of Bcl-2, which is an antiapoptotic protein, favoring cell death.
In addition to specific points of vanadium action, such as those mentioned above, the classical approach of interactions with DNA also appears to be quite effective; however, it is known that vanadium complexes can inhibit the proliferation of tumor cells by interfering in the cell cycle, causing arrest in several specific phases, such as the G1 phase and the G2/M phase [93,124,188]. This means that vanadium in its biologically active form, in the G1 phase, acts directly at the restriction point (R), interfering with the verification and presence of essential nutrients for growth signals and DNA integrity, thus interfering with the p53 protein [189], which can induce the expression of CDK inhibitors (especially p21) [189,190], preventing the repair cycle.
In a similar way, vanadium complexes act in the G2 phase [93,124,188,191] and in the M phase (mitosis), where there is the synthesis of proteins and organelles necessary for cell division, specifically acting on regulatory proteins, including cyclin B [93] and cyclin-dependent kinase 1 (CDK1) [68,69]. In 2019, Lu and coworkers [192] reported a series of octahedral vanadium(III) complexes containing [VIII(Cl)2THF] cores and imine ligands derived from indene moieties and S-alkyl/phenyl anilines ([VIII(Cl)2L′L″THF], Figure 6). These vanadium complexes had antiproliferative activities for the human cancer cells: gastric (MGC803), esophageal (EC109), breast (MCF7), and liver (HepG2). The vanadium(III) complex with the best antiproliferative activity was capable of inducing apoptosis in MGC803 cells in a dose-dependent manner and through caspase-mediated mechanisms. This mechanism occurs due to changes in the expression of the pro- and anti-apoptotic proteins Bax and Bcl-2, respectively [192].
In vitro antineoplastic activity has been reported for different oxidation states of vanadium coordinated with imine ligands derived from S-substituted dithiocarbazates, e.g., Sarhan and coworkers [193] reported the synthesis and characterization of the octahedral oxidovanadium(IV) complex ([VIVO(smdha)(bpy)], Figure 6), where smdha and bpy is dehydroacetic acid Schiff base of S-methyldithiocarbazate and 2,2′-bipyridine, respectively) with a half-maximal inhibitory concentration (IC50) value of 29.30 and 59.16 μM for HepG2 and MCF7, respectively. These IC50 values are better than those corresponding to H2smdha (81.64 and 151.58 μM) and cisplatin (31.60 and 73.30 μM).
Additionally, Banerjee and coworkers [194] synthesized three oxidovanadium(IV) complexes based on 1,10-phenanthroline (phen) and 2,2′-bipyridine (bipy): [VIVOL1(phen)], [VIVOL2(phen)], and [VIVOL1(bipy)] (Figure 6) with IC50 values in the range of 16.60–29.86 μM against cervical cancer cells (HeLa). On the other hand, in a more recent article, Banerjee and coworkers [195] reported a series of mononuclear non-oxido vanadium(IV)complexes ([VIV(L1−4)2], Figure 6) featuring tridentate bi-negative ONS chelating S-alkyl/aryl-substituted dithiocarbazate ligands (H2L1−4) with an improvement in the IC50 values, being in the range of 6.3–17.1 μM. These vanadium-based complexes showed comparable cytotoxic potential with some chemotherapeutic drugs, e.g., cyclophosphamide (IC50 = 21.5 μM) and cisplatin (IC50 = 70 μM), as well as against the HeLa cells [194].
Interestingly, Sahu and coworkers [196] reported IC50 values in the range of 6.73–18.21 μM for one oxidomethoxidovanadium(V) ([VVO(L)(OMe)], Figure 6) and two mixed-ligand oxidovanadium(IV) ([VIVO(L)(phen)] and [VIVO(L)(bipy)]) (Figure 6) against MCF7 cells. The obtained IC50 values for the vanadium-based complexes are better than those obtained for cisplatin (73.55 μM) and for the ligands 1,10-phenanthroline (phen), 2,2′-bipyridine (bipy), and S-benzyl-3-(2-hydroxy-3-ethoxyphenyl)methylenedithiocarbazate (H2L) (25.02–38.67 μM range, Figure 6). The same trend was also identified by Yekke-Ghasemi and coworkers [197] for mono- or dinucleated oxidovanadium(V) complexes with imine ligands also derived from S-substituted dithiocarbazates (Figure 6, IC50 = 0.6692 μM). Overall, these data reinforce the importance of vanadium in the discovery of novel anticarcinogenic candidates.

10. Serum Albumin in Metallodrug Discovery

Serum albumin is the main globular protein found in the blood being synthesized in the liver. Among its several important functions are its buffering pH capacity, neutralization of radicals, and its reducing of oxidative stress, maintaining of oncotic pressure, and binding and transporting of different endogenous and exogenous compounds (including hormones, fatty acids, calcium, magnesium, and drugs) [198,199,200,201,202,203,204,205]. In the drug discovery field, there is one important pre-clinical step about the biophysical characterization of potential drugs with serum albumin. More precisely, bovine serum albumin (BSA) is one of the most studied protein models in this step mainly due to its lowest cost compared with the human serum albumin (HSA) and the high percentage of identity and similarity (76 and 88%, respectively) among them [205,206,207,208]. Thus, the drug-binding affinities of BSA can be comparable with HSA.
The α-helix motifs comprise most of the structural content of the heart-shaped serum albumin. The molecular weight for BSA and HSA is 66,463 and 66,439 Da, respectively, with corresponding total amino acid residues of 583 and 585 [209,210]. The root mean square deviation (RMSD) value for their non-bound structure is around 1.634 dimensionless (Figure 7A), reinforcing their high three-dimensional (3D) structure similarity. For HSA (Figure 7B), residues 1–195, 196–383, and 384–585 form the α-helical domains I, II, and III, respectively, while, for BSA, slight differences can be evidenced (Figure 7C) [211,212]. Compounds with different chemical natures might bind with serum albumin, mainly due to the existence of two subdomains (A and B) for each domain. Overall, long and medium fatty acids can bind into seven possible sites, while four additional sites were reported for short fatty acids. Additionally, there is one binding site to bilirubin and three main sites for drugs located in subdomains IIA, IIIA, and IB, known as sites I, II, and III, respectively (Figure 7) [206,210]. These three subdomains are the main ones explored in drug-displacement assays [213,214,215,216,217].
The interaction between metallodrugs and serum albumin is of significant interest because it can influence the pharmacokinetic, bioavailable, and therapeutic efficacy of these drugs. Generally, metallodrugs bind to serum albumin with high affinity due to the interaction between the metal ions and amino acid residues, impacting their metabolism, distribution, and excretion [219,220,221,222,223]. High affinity might lead to a longer half-life and prolonged circulation time in the bloodstream, as well as facilitate the solubilization of metallodrugs in a complex biological matrix and protect them from premature degradation or inactivation (minimizing potential toxicity) [224]. As an example, cisplatin and its derivatives, carboplatin and oxaliplatin (Figure 8), bind to serum albumin, affecting not only their distribution but also decreasing the renal toxicity and limiting the concentration of free drugs in the bloodstream that might cause damage to the kidney [225,226,227,228,229]. Additionally, in regard to auranofin (Figure 8), a gold-based prodrug used in the treatment of rheumatoid arthritis, it was reported that the binding to albumin affects its anti-inflammatory activity [224,230,231] as well as the interaction between albumin with gallium nitrate (Figure 8, used for treating hypercalcemia and certain cancers) [232] and ruthenium-based compounds (mainly investigated as anticancer, e.g., NAMI-A, NKP1339, RM175, and RAPTA-C, Figure 8) [233], which is crucial for their stability and in reducing potential side effects.

11. Biophysical Characterization on the Interactions Between Albumin and Vanadium-Based Compounds

Biophysical techniques, e.g., ultraviolet-visible (UV-Vis) absorption, isothermal titration calorimetry, steady-state and time-resolved fluorescence, combined with in silico calculations, are widely used to evaluate the interactions between albumin and different metallodrugs or prodrugs, including those which are vanadium-based [125,234,235,236,237]. This characterization offers a crucial role regarding design, efficacy, and safety [202]. As an example, vanadium has antidiabetic properties, and its binding to albumin may influence its availability and efficacy in glucose metabolism regulation, as well as vanadium compounds that can affect enzyme activity, which their interaction with albumin might modulate [126,127,128]. This same interaction might serve as a detoxification pathway for vanadium, reducing its free concentration in the blood and mitigating its potential toxic effects [129].
Albumin has specific binding sites for metals and other small molecules. The vanadium ions vanadate (VO43−) and oxidovanadium (VO2+) interact specifically in the binding sites for metals [130], e.g., Cobbina and coworkers [131] reported that VIVO2+ occupies at least two types of binding sites, specifically a strong vanadium binding site (known as VBS1, without indicating its specific location) and the weak vanadium binding site (known as VBS2, also without indicating its specific location). This evidence was obtained by the combination of circular dichroism (CD), electron paramagnetic resonance (EPR), and UV-Vis absorption data. Interestingly, VBS1 binds with the 1 mol equivalent of VIVO+2 while VBS2 binds with several mol equivalents of VIVO via non-specific interactions involving carboxylate or imidazole side chains from the amino acid residues [131,132]. The presence of 1,2-dimethyl-3-hydroxy-4(1H)-pyridinone (maltol) may enhance the binding capacity of VIVO to albumin, while fatted albumins decreased the binding capacity of VIVO due to the structural change on the biomacromolecule induced by the fatty acids [33].
Unfortunately, a few reports have discussed the structural characterization of the pose of vanadium-based compounds into albumin. Generally, in silico calculations are utilized to suggest it, as depicted in Figure 9. Chaves and coworkers [238] identified subdomain IIA (site I), where the fluorophore Trp-214 residue can be found, as the main region for the interaction between HSA and dioxidovanadium(V) complexes coordinated with pyridoxal and/or resorcinol (complexes C1–C3 in Figure 9), being stabilized mainly by hydrogen bonding and van der Waals forces. However, in a more recent report, Martins and coworkers [125] identified that dioxidovanadium(V) complexes coordinated with hydrazone-type iminic ligands derived from (3-formyl-4-hydroxybenzyl)triphenylphosphonium chloride and aromatic hydrazides (complexes C4–C8 in Figure 9) bind into an external pocket, known as site III, probably due to the high steric volume of the ligands coordinated with vanadium(V) species. Hydrophobic, hydrogen bond, and π-cation interactions were identified as the main forces responsible for the stability of the complex HSA:C4–C8. In this case, potential π-cation interactions were identified due to the high number of positive charged amino acid residues located into subdomain IB.
Interestingly, Fioravanço and coworkers [135] identified that dioxidovanadium(V) complexes coordinated with p-substituted benzohydrazides condensed with salicylaldehyde (complexes C9–C12 in Figure 9) or pyridoxal hydrochloride (complexes C13–C16 in Figure 9) interacts into subdomain IIA (site I) independently of being positive charged or neutral complexes, being stabilized by the same intermolecular forces reported by Chaves and coworkers [238]. Finally, Dias and coworkers [126] also identified site I as the main binding pocket to some vanadium complexes, i.e., [VVO2(maltol)2], [VVO2(dmpp)(OH)(H2O)], and [VVO2(dmpp)2], where maltol and dmpp are 3-hydroxy-2-methyl-4-pyrone and 1,2-dimethyl-3-hydroxy-4(1H)-pyridinone, respectively. In this case, the interaction between the inorganic complexes with the protein surface is mainly driven by hydrogen bonding and/or hydrophobic forces.
Since we summarized the binding sites for vanadium(V) complexes exclusively, there are insights indicating that the main binding site for vanadium-based compounds is more dependent on the nature of the coordinated ligands than on the charge of the metal center. The binding site and type of interactions that lead to the stability of the complex albumin:vanadium-based compounds impact the thermodynamic trend that drives the association.
Generally, the time-resolved fluorescence decays to non-bound albumin are better fitted in a bi-exponential function with the measured corresponding to one shorter (τ1) and other longer lifetimes (τ2), e.g., for HSA 1.76 ± 0.09 and 5.75 ± 0.09 ns, respectively [133]. If the fluorescence lifetimes for albumin did not significantly change in the presence of vanadium-based compounds, it can be stated that the quenching process occurred through a purely static mechanism, indicating that the Stern–Volmer approach is the best mathematical approximation to estimate the binding affinity [134]. It was detected via steady-state and/or time-resolved fluorescence measurements that the binding vanadium-based compounds and albumin occur by ground state association (static fluorescence quenching mechanisms), e.g., for dioxidovanadium(V) complexes with a triphenylphosphonium moiety [125] or the mixture of the ground state (static) and collisional phenomenon (dynamic), e.g., vanadium(V) complexed with salicylaldehyde or pyridoxal hydrochloride [135], indicating that the ligands complexed with vanadium ions might interfere in the main fluorescence quenching mechanisms for albumin. A purely dynamic mechanism for albumin induced by vanadium-based compounds was not reported.
In a mixture of static and dynamic quenching mechanisms, the Stern–Volmer quenching constant (KSV) is not used to determine the binding affinity; therefore, the binding constant (Ka or Kb) obtained mainly by modified Stern–Volmer or double logarithmic approaches [134] can be used to estimate the binding capacity between albumin and vanadium-based compounds. Drug absorption is described as the initial pharmacokinetic step which can be negatively affected by the weak binding capacity of metallodrugs to albumin; however, both the distribution and absorption of metallodrugs to different tissues (targets) is feasible when the binding is moderate, indicating some differences in cases of high binding affinity–feasible absorption, but its distribution to the tissues is limited mainly due to the stability of the complex [202,239]. In this sense, high and moderate binding affinity are detected when KSV, Ka, or Kb values are higher than 106 M−1 and in the range of 104–106 M−1, respectively. The reported data indicated that most of the vanadium-based complexes bind moderately to strongly with albumin [125,126,127,128,129,130,132,135,238], meaning that these potential metallodrugs might be carried out by HSA in the human bloodstream with a favorable binding capacity.
It is important to highlight that, since the binding between albumin and vanadium-based complexes is mainly characterized by in vitro assays without the complexity of a biological medium, it is difficult to extrapolate the biophysical data for in vivo assays. However, some reports [224,240] indicate an increase in the aqueous solubility and residence time, as well as a decrease in the toxicity of inorganic complexes after binding with albumin, leading to a favorable environment on the pharmacological behavior of metallodrugs. Additionally, when metallodrugs bind to albumin, they are often less available to interact with their intended target, potentially reducing efficacy but also potentially reducing off-target effects [223,241,242]. In other words, albumin also can be used in drug delivery systems to improve the pharmacokinetics of vanadium-based compounds, potentially reducing off-target effects.

12. Structural Evaluation of the Interaction Between Model Proteins and Vanadium-Based Compounds

Some proteins have been used as model proteins to evaluate and better understand the interactions between vanadium complexes and proteins by several techniques. Such proteins have structural characteristics that justify their use, such as a small number of binding sites, moderate size, high conformational stability, and high purity grades with commercial availability [82]. Hen egg white lysozyme is a small antimicrobial glycoside hydrolase which has been used as a model protein for X-ray crystallography and is able to produce robust and well-diffracting crystals [243,244]. Bovine pancreatic trypsin is one of the most typical serine proteases. From a crystallographic point of view, both these model proteins produce well-diffracting crystals under different experimental conditions [244]. Some vanadium(IV) complexes, including bis(maltolato)oxidovanadium(IV) (BMOV), had their interactions with hen egg white lysozyme and bovine trypsin evaluated by different techniques, e.g., electrospray ionization mass spectrometry (ESI-MS), electron paramagnetic resonance (EPR), and even X-ray crystallography [245,246,247,248]. Despite the advances of this last technique, there are still a few studies that have reported interactions between oxidovanadium(IV) ions and proteins using this resource [82,245].
Regarding the maltolato-derived complex, ESI-MS(+) spectra were recorded. In the raw spectrum, for each HEWL peak, other signals at higher m/z values are visible, indicating the formation of HEWL−VIVO−maltolate adducts. In the deconvoluted spectrum, interactions and multiple binding of the “initial” VIVO(maltolato)2 specie and [VIVO(maltolato)]+ specie are evidenced at both pH values, as well as the binding of the [VIVO]2+ ions at the more acidic pH. With increasing pH, the adducts with the VIVO(maltolate)2 species also increase. By means of anisotropic EPR, it is patterns that suggest the coordination of oxidovanadium(IV) metal centers to oxygens of aspartic acid/glutamic acid-COO or asparagine/glutamine-CO side chains which contradict the preference of cisplatin (with a softer PtII center) for the histidine residue of the evaluated protein [245].
By means of soaking technique, HEWL single crystals obtained at pH 7.5 were immersed in BMOV solution, obtaining two distinct single crystals, although both have three equivalent binding sites. In both the crystals, there is evidence of different vanadium(IV) species interacting by non-covalent bindings with the protein surface (cis-[VIVO(maltolato)2(H2O)] and [VIVO(maltolato)(H2O)3] in the first crystal and two molecules of [VIVO(maltolato)2(H2O)] in the second) and covalent bindings ([VIVO(H2O)4]2+ in the first crystal and cis-[VIVO(maltolato)2] in the second, both binding to the same asparagine residue). By changing the growth medium of the HEWL single crystal, including reducing the pH to 4.0, which was subsequently immersed in BMOV solution, a different single crystal was obtained. The structure of this one now presents covalent interactions of three species [VIVO(H2O)3]2+ together with two additional vanadium atoms [245].
Recently, by SC-XRD technique, single crystals of the same model protein but obtained at 1.1 M sodium chloride and 0.1 M sodium acetate at pH 4.0 immersed in BMOV solution had their solid-state structure elucidated, evidencing four distinct V binding sites. Between them, an unexpected bis(vanadium)-containing fragment is localized. This species is composed of two vanadium metal centers, both fully coordinated with oxygen species. Also, a modification on the side chain of a lysine residue, corresponding to the formation of an iminic double bond between the keto group of the maltolate moiety and the N-terminal group of this amino acid, has been noted. The two involved maltolate ligands suffered cleavage, both being coordinated to one vanadium center. In addition to the lysine residue, the fragment is coordinated to an aspartate residue from a symmetry related molecule [246]. The presence of this fragment containing two vanadium metal centers is also confirmed by gel electrophoresis under denaturing conditions of solubilized soaked crystals and by mass spectroscopies [246].
Co-crystallization of bovine trypsin in the presence of VIVO/picolinic acid allows crystallographic results showing only a single adduct between vanadium and the model protein, with the ion being elucidated as an oxidovanadium(IV) with a distorted octahedral geometry. The vanadium coordination polyhedra is completed by a bond with a serine residue (alkoxide donor atom) and two picolinate anionic ligands by its N, O donor atoms. Co-crystallization of the same enzyme with vanadium(IV) and phenanthroline (phen) in the presence of imidazole (crystallization conditions in this case) afforded a structure with a single V-enzyme adduct. The metallic center is coordinated to a serine residue by its alkoxide atom, to a bidentate phenanthroline ligand, and to an imidazole. The distorted octahedral geometry is completed by two oxygen donor atoms, which could be interpreted as two oxido or an oxido–hydroxy couple and, consequently, as a dioxidovanadium(V) or an oxidovanadium(IV) species, respectively [247].
The trypsin crystals with vanadium species did not show EPR signals at 77 K, probably indicating a diamagnetic vanadium(V) species. However, fresh solutions of lyophilized trypsin and VIVOSO4 in the presence of phenanthroline/bipyridine (molar ratio 1:1 and 1:2) at pH 7.4 showed an EPR signal, indicating a probable oxidation of the vanadium species in the crystallization process. [247].

13. Future Perspectives

The future of metallodrugs lies in leveraging their unique properties for targeted, effective, and personalized medical treatments. Advances in targeted delivery systems, mechanistic understanding, combination therapies, and personalized medicine will drive the development of next generation metallodrugs. As research continues to uncover new therapeutic applications and optimize existing ones, metallodrugs will play an increasingly important role in modern medicine. Based on the biological importance of vanadium in different organisms, mainly targeting some specific enzymes with a good biophysical characterization of the interaction with serum albumin (in vitro pharmacokinetic profile), the vanadium-based compounds have significant promise regarding advanced medical treatments. These potential drugs leverage the unique properties of metals, such as redox activity, coordination chemistry, and electronic structure, to target diseases in ways that traditional organic molecules cannot.
The design and synthesis of novel vanadium-based compounds with low toxicity and greater biocompatibility are still crucial. Based on this review, the importance of the relationship between the chemical structure of potential metallodrugs and their biological activity to optimize their therapeutic properties was noticed. In addition, the understanding of how vanadium-based compounds interact with biological molecules, e.g., with serum albumin at the molecular level (biophysical characterization), can lead to the design of more effective and selective agents to decrease the toxicity, such as off-target effects, and improve the therapeutic aspects with good residence time in the bloodstream.

Author Contributions

Conceptualization, D.F.B. and O.A.C.; methodology, D.F.B., F.M.M., C.S. and O.A.C.; software, D.F.B. and O.A.C.; validation, D.F.B. and O.A.C.; formal analysis, D.F.B., F.M.M., C.S. and O.A.C.; investigation, D.F.B., F.M.M., C.S. and O.A.C.; resources, D.F.B. and O.A.C.; data curation, O.A.C.; writing—original draft preparation, D.F.B., F.M.M., C.S. and O.A.C.; writing—review and editing, D.F.B., F.M.M., C.S. and O.A.C.; visualization, O.A.C.; supervision, D.F.B. and O.A.C.; project administration, D.F.B. and O.A.C.; funding acquisition, D.F.B. and O.A.C. All authors have read and agreed to the published version of the manuscript.

Funding

This manuscript was funded by the Brazilian Research Councils: Conselho Nacional do Desenvolvimento Científico e Tecnológico (PQ–2022; 308411/2022-6) and Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES-PROEX, 001). This research was also funded by the Portuguese Agency for Scientific Research, Fundação para a Ciência e a Tecnologia (FCT), through the projects UIDB/00313/2020 (https://doi.org/10.54499/UIDB/00313/2020, accessed on 28 June 2024) and UIDP/00313/2020 (https://doi.org/10.54499/UIDP/00313/2020, accessed on 28 June 2024).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

O.A.C. thanks Fundação para a Ciência e a Tecnologia (FCT) for his PhD fellowship 2020.07504.BD (https://doi.org/10.54499/2020.07504.BD, accessed on 28 June 2024).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hernandez, L.; Araujo, M.L.; Madden, W.; Del Carpio, E.; Lubes, V.; Lubes, G. Vanadium complexes with polypyridyl ligands: Speciation, structure and potential medicinal activity. J. Inorg. Biochem. 2022, 229, 111712. [Google Scholar] [CrossRef] [PubMed]
  2. Shaheen, S.M.; Alessi, D.S.; Tack, F.M.G.; Ok, Y.S.; Kim, K.-H.; Gustafsson, J.P.; Sparks, D.L.; Rinklebe, J. Redox chemistry of vanadium in soils and sediments: Interactions with colloidal materials, mobilization, speciation, and relevant environmental implications—A review. Adv. Colloid Interf. Sci. 2019, 265, 1–13. [Google Scholar] [CrossRef]
  3. Sahu, G.; Banerjee, A.; Samanta, R.; Mohanty, M.; Lima, S.; Tiekink, E.R.T.; Dinda, R. Water-soluble dioxidovanadium(V) complexes of aroylhydrazones: DNA/BSA interactions, hydrophobicity, and cell-selective anticancer potential. Inorg. Chem. 2021, 60, 15291–15309. [Google Scholar] [CrossRef]
  4. Bhunia, P.; Gomila, R.M.; Frontera, A.; Ghosh, A. Combine effect of Lewis acidity and electric field of proximal redox innocent metal ions on redox potential of vanadyl Schiff base complexes: Experimental and theoretical study. Dalton Trans. 2023, 52, 3097–3110. [Google Scholar] [CrossRef] [PubMed]
  5. King, A.E.; Nippe, M.; Atanasov, M.; Chantarojsiri, T.; Wray, C.A.; Bill, E.; Neese, F.; Long, J.R.; Chang, C.J. A well-defined terminal vanadium(III) oxo complex. Inorg. Chem. 2014, 53, 11388–11395. [Google Scholar] [CrossRef] [PubMed]
  6. Dieter, R. The future of/for vanadium. Dalton Trans. 2013, 42, 11749–11761. [Google Scholar]
  7. Gambino, D. New trends on vanadium chemistry, biochemistry, and medicinal chemistry. Inorganics 2022, 10, 68. [Google Scholar] [CrossRef]
  8. Rehder, D. The role of vanadium in biology. Metallomics 2015, 7, 730–742. [Google Scholar] [CrossRef]
  9. Del Carpio, E.; Hernández, L.; Ciangherotti, C.; Coa, V.V.; Jiménez, L.; Lubes, V.; Lubes, G. Vanadium: History, chemistry, interactions with α-amino acids and potential therapeutic applications. Coord. Chem. Rev. 2018, 372, 117–140. [Google Scholar] [CrossRef]
  10. Topolska, J.; Puzio, B.; Borkiewicz, O.; Sordyl, J.; Manecki, M. Solubility product of vanadinite Pb5(VO4)3Cl at 25 °C—A comprehensive approach to incongruent dissolution modeling. Minerals 2021, 11, 135. [Google Scholar] [CrossRef]
  11. Faudoa-Gómez, F.G.; Fuentes-Cobas, L.E.; Esparza-Ponce, H.E.; Canche-Tello, J.G.; Reyes-Cortés, I.A.; Fuentes-Montero, M.E.; Eichert, D.M.; Rodríguez-Guerra, Y.; Montero-Cabrera, M.-E. Geological and crystallochemical characterization of the margaritasite–carnotite mineral from the uranium region of Peña Blanca, Chihuahua, Mexico. Minerals 2024, 14, 431. [Google Scholar] [CrossRef]
  12. Rout, C.S.; Kim, B.-H.; Xu, X.; Yang, J.; Jeong, H.Y.; Odkhuu, D.; Park, N.; Cho, J.; Shi, H.S. Synthesis and characterization of patronite form of vanadium sulfide on graphitic layer. J. Am. Chem. Soc. 2013, 135, 8720–8725. [Google Scholar] [CrossRef] [PubMed]
  13. Zanetta, P.-M.; Drexler, M.S.; Barton, I.F.; Zega, T.J. Vanadium electronic configuration determination from L2,3 transition in V-oxide compounds and roscoelite. Microsc. Microanal. 2023, 29, 459–469. [Google Scholar] [CrossRef] [PubMed]
  14. Ueki, T.; Fujie, M.; Romaidi; Satoh, N. Symbiotic bacteria associated with ascidian vanadium accumulation identified by 16S rRNA amplicon sequencing. Mar. Genom. 2019, 43, 33–42. [Google Scholar] [CrossRef] [PubMed]
  15. Ueki, T.; Yamaguchi, N.; Romaidi; Isago, Y.; Tanahashi, H. Vanadium accumulation in ascidians: A system overview. Coord. Chem. Rev. 2015, 301, 300–308. [Google Scholar] [CrossRef]
  16. Yoshinaga, M.; Ueki, T.; Yamaguchi, N.; Kamino, K.; Michibata, H. Glutathione transferases with vanadium-binding activity isolated from the vanadium-rich ascidian Ascidia sydneiensis samea. Biochim. Biophys. Acta 2006, 1760, 495–503. [Google Scholar] [CrossRef]
  17. Michibata, H.; Ueki, T. Advances in research on the accumulation, redox behavior, and function of vanadium in ascidians. BioMol Concepts 2010, 1, 97–107. [Google Scholar] [CrossRef]
  18. Macara, I.G.; McLeod, G.C.; Kustin, K. Tunichromes and metal ion accumulation in tunicate blood cells. Comp. Biochem. Physiol. 1979, B63, 299–302. [Google Scholar] [CrossRef]
  19. Ryan, D.E.; Grant, K.B.; Nakanishi, K. Reactions between tunichrome Mm-1, a tunicate blood pigment, and vanadium ions in acidic and neutral media. Biochemistry 1996, 35, 8640–8650. [Google Scholar] [CrossRef]
  20. Abebe, A.; Kuang, Q.F.; Evans, J.; Robinson, W.E.; Sugumaran, M. Oxidative transformation of a tunichrome model compound provides new insight into the crosslinking and defense reaction of tunichromes. Bioorg. Chem. 2017, 71, 219–229. [Google Scholar] [CrossRef]
  21. Ueki, T.; Satake, M.; Kamino, K.; Michibata, H. Sequence variation of Vanabin2-like vanadium-binding proteins in blood cells of the vanadium-accumulating ascidian Ascidia sydneiensis samea. Biochim. Biophys. Acta 2008, 1780, 1010–1015. [Google Scholar] [CrossRef] [PubMed]
  22. Hamada, T.; Asanuma, M.; Ueki, T.; Hayashi, F.; Kobayashi, N.; Yokoyama, S.; Michibata, H.; Hirota, H. Solution structure of Vanabin2, a vanadium(IV)-binding protein from the vanadium-rich Ascidian Ascidia sydneiensis samea. J. Am. Chem. Soc. 2005, 127, 4216–4222. [Google Scholar] [CrossRef] [PubMed]
  23. Islam, M.K.; Tsuboya, C.; Kusaka, H.; Aizawa, S.; Ueki, T.; Michibata, H.; Kanamori, K. Reduction of vanadium(V) to vanadium(IV) by NADPH, and vanadium(IV) to vanadium(III) by cysteine methyl ester in the presence of biologically relevant ligands. Biochim. Biophys. Acta 2007, 1770, 1212–1218. [Google Scholar] [CrossRef] [PubMed]
  24. Kohlmeier, M. Nutrient Metabolism, 1st ed.; Academic Press: Cambridge, MA, USA, 2003; pp. 762–766. [Google Scholar]
  25. Anke, M. Vanadium—An element both essential and toxic to plants, animals and humans? Anal. Real Acad. Nac. Farm. 2004, 70, 961–999. [Google Scholar]
  26. Levina, A.; McLeod, A.I.; Kremer, L.E.; Aitken, J.B.; Glover, C.J.; Johannessen, B.; Lay, P.A. Reactivity-activity relationships of oral anti-diabetic vanadium complexes in gastrointestinal media: An X-ray absorption spectroscopic study. Metallomics 2014, 6, 1880–1888. [Google Scholar] [CrossRef]
  27. Kiss, T.; Jakusch, T.; Hollender, D.; Dornyei, A.; Enyedy, E.A.; Costa-Pessoa, J.; Sakurai, H.; Sanz-Medel, A. Biospeciation of antidiabetic VO(IV) complexes. Coord. Chem. Rev. 2008, 252, 1153–1162. [Google Scholar] [CrossRef]
  28. Yoshikawa, Y.; Sakurai, H.; Crans, D.C.; Micera, G.; Garribba, E. Structural and redox requirements for the action of anti-diabetic vanadium compounds. Dalton Trans. 2014, 43, 6965–6972. [Google Scholar] [CrossRef]
  29. Crans, D.C.; Bunch, R.L.; Theisen, L.A. Interaction of trace levels of vanadium(IV) and vanadium(V) in biological systems. J. Am. Chem. Soc. 1989, 111, 7597–7607. [Google Scholar] [CrossRef]
  30. Crans, D.C.; Smee, J.J.; Gaidamauskas, E.; Yang, L. The chemistry and biochemistry of vanadium and the biological activities exerted by vanadium compounds. Chem. Rev. 2004, 104, 849–902. [Google Scholar] [CrossRef]
  31. Crans, D.C. Fifteen years of dancing with vanadium. Pure Appl. Chem. 2005, 77, 1497–1527. [Google Scholar] [CrossRef]
  32. Cooper, I.; Ravid, O.; Rand, D.; Atrakchi, D.; Shemesh, C.; Bresler, Y.; Ben-Nissan, G.; Sharon, M.; Fridkin, M.; Shechter, Y. Albumin-EDTA-vanadium is a powerful anti-proliferative agent, following entrance into glioma cells via caveolae-mediated endocytosis. Pharmaceutics 2021, 13, 1557. [Google Scholar] [CrossRef] [PubMed]
  33. Correia, I.; Jakusch, T.; Cobbinna, E.; Mehtab, S.; Tomaz, I.; Nagy, N.V.; Rockenbauer, A.; Pessoa, J.C.; Kiss, T. Evaluation of the binding of oxovanadium(IV) to human serum albumin. Dalton Trans. 2012, 41, 6477–6487. [Google Scholar] [CrossRef] [PubMed]
  34. Azevedo, C.G.; Correia, I.; dos Santos, M.M.C.; Santos, M.F.A.; Santos-Silva, T.; Doutch, J.; Fernandes, L.; Santos, H.M.; Capelo, J.L.; Pessoa, J.C. Binding of vanadium to human serum transferrin—Voltammetric and spectrometric studies. J. Inorg. Biochem. 2018, 180, 211–221. [Google Scholar] [CrossRef] [PubMed]
  35. Levina, A.; Lay, P.A. Vanadium(V/IV)−transferrin binding disrupts the transferrin cycle and reduces vanadium uptake and antiproliferative activity in human lung cancer cells. Inorg. Chem. 2020, 59, 16143–16153. [Google Scholar] [CrossRef]
  36. Sanna, D.; Micera, G.; Garribba, E. Interaction of VO2+ ion and some insulin-enhancing compounds with immunoglobulin G. Inorg. Chem. 2011, 50, 3717–3728. [Google Scholar] [CrossRef]
  37. Schrier, S.L.; Junga, I.; Ma, L. Studies on the effect of vanadate on endocytosis and shape changes in human red blood cells and ghosts. Blood 1986, 68, 1008–1014. [Google Scholar] [CrossRef]
  38. Amaral, L.M.P.F.; Moniz, T.; Silva, A.M.N.; Rangel, M. Vanadium compounds with antidiabetic potential. Int. J. Mol. Sci. 2023, 24, 15675. [Google Scholar] [CrossRef]
  39. Yang, X.; Wang, K.; Lu, J.; Crans, D.C. Membrane transport of vanadium compounds and the interaction with the erythrocyte membrane. Coord. Chem. Rev. 2003, 237, 103–111. [Google Scholar] [CrossRef]
  40. Nechay, B.R.; Saunders, J.P. Inhibition by vanadium of sodium and potassium dependent adenosinetriphosphatase derived from animal and human tissues. J. Environ. Pathol. Toxicol. 1978, 2, 247–262. [Google Scholar]
  41. Aureliano, M.; Ohlin, C.A. Decavanadate in vitro and in vivo effects: Facts and opinions. J. Inorg. Biochem. 2014, 137, 123–130. [Google Scholar] [CrossRef]
  42. Marques, M.P.M.; Gianolio, D.; Ramos, S.; Batista de Carvalho, L.A.E.; Aureliano, M. An EXAFS approach to the study of polyoxometalate-protein interactions: The case of decavanadate-actin. Inorg. Chem. 2017, 56, 10893–10903. [Google Scholar] [CrossRef] [PubMed]
  43. Pessoa, J.C. Thirty years through vanadium chemistry. J. Inorg. Biochem. 2015, 147, 4–24. [Google Scholar] [CrossRef] [PubMed]
  44. Akabayov, S.R.; Akabayov, B. Vanadate in structural biology. Inorg. Chim. Acta 2014, 420, 16–23. [Google Scholar] [CrossRef]
  45. Höfler, G.T.; But, A.; Hollmann, F. Haloperoxidases as catalysts in organic synthesis. Org. Biomol. Chem. 2019, 17, 9267–9274. [Google Scholar] [CrossRef]
  46. Langeslay, R.R.; Kaphan, D.M.; Marshall, C.L.; Stair, P.C.; Sattelberger, A.P.; Delferro, M. Catalytic applications of vanadium: A mechanistic perspective. Chem. Rev. 2019, 119, 2128–2191. [Google Scholar] [CrossRef]
  47. Messerschmidt, A.; Prade, L.; Wever, R. Implications for the catalytic mechanism of the vanadium-containing enzyme chloroperoxidase from the fungus Curvularia inaequalis by X-ray structures of the native and peroxide form. Biol. Chem. 1997, 378, 309–315. [Google Scholar] [CrossRef]
  48. Leblanc, C.; Vilter, H.; Fournier, J.-B.; Delage, L.; Potin, P.; Rebuffet, E.; Michel, G.; Solari, P.L.; Feiters, M.C.; Czjzeka, M. Vanadium haloperoxidases: From the discovery 30 years ago to X-ray crystallographic and V K-edge absorption spectroscopic studies. Coord. Chem. Rev. 2015, 301, 134–146. [Google Scholar] [CrossRef]
  49. Wever, R.; Barnett, P. Vanadium chloroperoxidases: The missing link in the formation of chlorinated compounds and chloroform in the terrestrial environment? Chem. Asian J. 2017, 12, 1997–2007. [Google Scholar] [CrossRef]
  50. Porta, N.; Fejzagić, A.V.; Dumschott, K.; Paschold, B.; Usadel, B.; Pietruszka, J.; Classen, T.; Gohlke, H. Identification and characterization of the haloperoxidase VPO-RR from Rhodoplanes roseus by genome mining and structure-based catalytic site mapping. Catalysts 2022, 12, 1195. [Google Scholar] [CrossRef]
  51. Winter, J.M.; Moore, B.S. Exploring the chemistry and biology of vanadium-dependent haloperoxidases. J. Biol. Chem. 2009, 284, 18577–18581. [Google Scholar] [CrossRef]
  52. Ortiz-Bermúdez, P.; Hirth, K.C.; Srebotnik, E.; Hammel, K.E. Chlorination of lignin by ubiquitous fungi has a likely role in global organochlorine production. Proc. Natl. Acad. Sci. USA 2007, 104, 3895–3900. [Google Scholar] [CrossRef] [PubMed]
  53. Rehder, D. Vanadium in biological systems and medicinal applications. Inorg. Chim. Acta 2023, 549, 121387. [Google Scholar] [CrossRef]
  54. Gérard, E.F.; Mokkawes, T.; Johannissen, L.O.; Warwicker, J.; Spiess, R.R.; Blanford, C.F.; Hay, S.; Heyes, D.J.; de Visser, S.P. How is substrate halogenation triggered by the vanadium haloperoxidase from Curvularia inaequalis? ACS Catal. 2023, 13, 8247–8261. [Google Scholar] [CrossRef] [PubMed]
  55. Mubarak, M.Q.E.; Gérard, E.F.; Blanford, C.F.; Hay, S.; de Visser, S.P. How do vanadium chloroperoxidases generate hypochlorite from hydrogen peroxide and chloride? A computational study. ACS Catal. 2020, 10, 14067–14079. [Google Scholar] [CrossRef]
  56. Zhilong, C. Recent development of biomimetic halogenation inspired by vanadium dependent haloperoxidase. Coord. Chem. Rev. 2022, 457, 214404. [Google Scholar]
  57. Seefeldt, L.C.; Yang, Z.-Y.; Lukoyanov, D.A.; Harris, D.F.; Dean, D.R.; Raugei, S.; Hoffman, B.M. Reduction of substrates by nitrogenases. Chem. Rev. 2020, 120, 5082–5106. [Google Scholar] [CrossRef]
  58. Sippel, D.; Einsle, O. The structure of vanadium nitrogenase reveals an unusual bridging ligand. Nature Chem. Biol. 2017, 13, 956–960. [Google Scholar] [CrossRef]
  59. Sippel, D.; Rohde, M.; Netzer, J.; Trncik, C.; Gies, J.; Grunau, K.; Djurdjevic, I.; Decamps, L.; Andrade, S.L.A.; Einsle, O. A bound reaction intermediate sheds light on the mechanism of nitrogenase. Science 2018, 359, 1484–1489. [Google Scholar] [CrossRef]
  60. Yang, Z.-Y.; Jimenez-Vicente, E.; Kallas, H.; Lukoyanov, D.A.; Yang, H.; Martin Del Campo, J.S.; Dean, D.R.; Hoffman, B.M.; Seefeldt, L.C. The electronic structure of FeV-cofactor in vanadium-dependent nitrogenase. Chem. Sci. 2021, 12, 6913–6922. [Google Scholar] [CrossRef]
  61. Rees, J.A.; Bjornsson, R.; Schlesier, J.; Sippel, D.; Einsle, O.; Debeer, S. The Fe–V cofactor of vanadium nitrogenase contains an interstitial carbon atom. Angew. Chem. Int. Ed. 2015, 54, 13249–13252. [Google Scholar] [CrossRef]
  62. Ugone, V.; Sanna, D.; Sciortino, G.; Crans, D.C.; Garribba, E. ESI-MS study of the interaction of potential oxidovanadium(IV) drugs and amavadin with model proteins. Inorg. Chem. 2020, 59, 9739–9755. [Google Scholar] [CrossRef] [PubMed]
  63. da Silva, J.A.L.; Fraústo da Silva, J.J.R.; Pombeiro, A.J.L. Amavadin, a vanadium natural complex: Its role and applications. Coord. Chem. Rev. 2013, 257, 2388–2400. [Google Scholar] [CrossRef]
  64. Berry, R.E.; Armstrong, E.M.; Beddoes, R.L.; Collison, D.; Ertok, S.N.; Helliwell, M.; Garner, C.D. The structural characterization of amavadin. Angew. Chem. Int. Ed. 1999, 38, 795–797. [Google Scholar] [CrossRef]
  65. Ferraro, G.; Merlino, A. Metallodrugs: Mechanisms of action, molecular targets and biological activity. Int. J. Mol. Sci. 2022, 23, 3504. [Google Scholar] [CrossRef]
  66. Anthony, E.J.; Bolitho, E.M.; Bridgewater, H.E.; Carter, O.W.L.; Donnelly, J.M.; Imberti, C.; Lant, E.C.; Lermyte, F.; Needham, R.J.; Palau, M.; et al. Metallodrugs are unique: Opportunities and challenges of discovery and development. Chem. Sci. 2020, 11, 12888–12917. [Google Scholar] [CrossRef]
  67. Fontana, L.A.; Martins, F.M.; Siqueira, J.D.; Serpa, C.; Chaves, O.A.; Back, D.F. Synthesis of cobalt(III) complexes derived from pyridoxal: Structural cleavage evaluations and in silico calculations for biological targets. Inorganics 2024, 12, 171. [Google Scholar] [CrossRef]
  68. Ferretti, V.A.; León, I. An Overview of vanadium and cell signaling in potential cancer treatments. Inorganics 2022, 10, 47. [Google Scholar] [CrossRef]
  69. Amante, C.; de Sousa-Coelho, A.L.; Aureliano, M. Vanadium and melanoma: A systematic review. Metals 2021, 11, 828. [Google Scholar] [CrossRef]
  70. Selvaraj, S.; Krishnan, U.M. Vanadium-flavonoid complexes: A promising class of molecules for therapeutic applications. J. Med. Chem. 2021, 64, 12435–12452. [Google Scholar] [CrossRef]
  71. Kioseoglou, E.; Petanidis, S.; Gabriel, C.; Salifoglou, A. The chemistry and biology of vanadium compounds in cancer therapeutics. Coord. Chem. Rev. 2015, 301, 87–105. [Google Scholar] [CrossRef]
  72. Arora, J.P.S.; Singh, R.P.; Soam, D.; Sharma, R. Comparison of binding of vanadium(V) with bovine serum albumin and bovine pancreatic trypsin. J. Electroanal. Chem. Interf. Electrochem. 1983, 155, 57–67. [Google Scholar] [CrossRef]
  73. De Sousa-Coelho, A.L.; Fraqueza, G.; Aureliano, M. Repurposing therapeutic drugs complexed to vanadium in cancer. Pharmaceuticals 2024, 17, 12. [Google Scholar] [CrossRef] [PubMed]
  74. Pessoa, J.C.; Etcheverry, S.; Gambino, D. Vanadium compounds in medicine. Coord. Chem. Rev. 2015, 301, 24–48. [Google Scholar] [CrossRef] [PubMed]
  75. Crans, C.D.; Zhang, B.; Gaidamauskas, E.; Keramidas, A.D.; Willsky, G.R.; Roberts, C.R. Is vanadate reduced by thiols under biological conditions? Changing the redox potential of V(V)/V(IV) by complexation in aqueous solution. Inorg. Chem. 2010, 49, 4245–4256. [Google Scholar] [CrossRef] [PubMed]
  76. Cruywagen, J.J.; Heyns, J.B.B.; Westra, A.N. Protonation equilibria of mononuclear vanadate: Thermodynamic evidence for the expansion of the coordination number in VO2+. Inorg. Chem. 1996, 35, 1556–1559. [Google Scholar] [CrossRef] [PubMed]
  77. Mao, L.-L.; Hao, D.-L.; Mao, X.-W.; Xu, Y.-F.; Huang, T.-T.; Wu, B.-N.; Wang, L.-H. Neuroprotective effects of bisperoxovanadium on cerebral ischemia by inflammation inhibition. Neurosci. Lett. 2015, 602, 120–125. [Google Scholar] [CrossRef]
  78. Shaik, A.; Kondaparthy, V.; Begum, A.; Husain, A.; Das Manwal, D. Enzyme PTP-1B inhibition studies by vanadium metal complexes: A kinetic approach. Biol. Trace Elem. Res. 2023, 201, 5037–5052. [Google Scholar] [CrossRef]
  79. Thompson, K.H.; Orvig, C. Coordination chemistry of vanadium in metallopharmaceutical candidate compounds. Coord. Chem. Rev. 2001, 219, 1033–1053. [Google Scholar] [CrossRef]
  80. Mjos, K.D.; Orvig, C. Metallodrugs in medicinal inorganic chemistry. Chem. Rev. 2014, 114, 4540–4563. [Google Scholar] [CrossRef]
  81. Kabir, E.; Noyon, M.R.O.K.; Hossain, M.A. Synthesis, biological and medicinal impacts of metallodrugs: A study. Res. Chem. 2023, 5, 100935. [Google Scholar] [CrossRef]
  82. Pessoa, J.C.; Santos, M.F.A.; Correia, I.; Sanna, D.; Sciortino, G.; Garribba, E. Binding of vanadium ions and complexes to proteins and enzymes in aqueous solution. Coord. Chem. Rev. 2021, 449, 214192. [Google Scholar] [CrossRef]
  83. Treviño, S.; Diaz, A. Vanadium and insulin: Partners in metabolic regulation. J. Inorg. Biochem. 2020, 208, 111094. [Google Scholar] [CrossRef] [PubMed]
  84. Goldwaser, I.; Gefel, D.; Gershonov, E.; Fridkin, M.; Shechter, Y. Insulin-like effects of vanadium: Basic and clinical implications. J. Inorg. Biochem. 2000, 80, 21–25. [Google Scholar] [CrossRef] [PubMed]
  85. Thompson, K.H.; Orvig, C. Design of vanadium compounds as insulin enhancing agents. J. Chem. Soc. Dalton Trans. 2000, 17, 2885–2892. [Google Scholar] [CrossRef]
  86. Scibior, A.; Pietrzyk, L.; Plewa, Z.; Skiba, A. Vanadium: Risks and possible benefits in the light of a comprehensive overview of its pharmacotoxicological mechanisms and multi-applications with a summary of further research trends. J. Trace Elem. Med. Biol. 2020, 61, 126508. [Google Scholar] [CrossRef]
  87. Trevino, S.; Diaz, A.; Sanchez-Lara, E.; Sanchez-Gaytan, B.L.; Perez-Aguilar, J.M.; Gonzalez-Vergara, V. Vanadium in biological action: Chemical, pharmacological aspects, and metabolic implications in diabetes mellitus. Biol. Trace Elem. Res. 2019, 188, 68–98. [Google Scholar] [CrossRef]
  88. Sharfalddin, A.A.; Al-Younis, I.M.; Mohammed, H.A.; Dhahri, M.; Mouffouk, F.; Abu Ali, H.; Anwar, M.J.; Qureshi, K.A.; Hussien, M.A.; Alghrably, M.; et al. Therapeutic properties of vanadium complexes. Inorganics 2022, 10, 244. [Google Scholar] [CrossRef]
  89. Thompson, K.H.; Orvig, C. Vanadium in diabetes: 100 years from Phase 0 to Phase I. J. Inorg. Biochem. 2006, 100, 1925–1935. [Google Scholar]
  90. He, L.; Wang, X.; Zhao, C.; Zhu, D.; Du, W. Inhibition of human amylin fibril formation by insulin-mimetic vanadium complexes. Metallomics 2014, 6, 1087. [Google Scholar] [CrossRef]
  91. Mehdi, M.Z.; Pandey, S.K.; Théberge, J.-F.; Srivastava, A.K. Insulin signal mimicry as a mechanism for the insulin-like effects of vanadium. Cell Biochem. Biophys. 2006, 44, 73–81. [Google Scholar]
  92. Boucher, J.; Kleinridders, A.; Kahn, C.R. Insulin receptor signaling in normal and insulin-resistant states. Cold Spring Harb. Perspect. Biol. 2014, 6, a009191. [Google Scholar] [CrossRef]
  93. Kowalski, S.; Wyrzykowski, D.; Inkielewicz-Stepniak, I. Molecular and cellular mechanisms of cytotoxic activity of vanadium compounds against cancer cells. Molecules 2020, 25, 1757. [Google Scholar] [CrossRef] [PubMed]
  94. Santos, T.M.R.; Tavares, C.A.; Pereira, A.F.; da Cunha, E.F.F.; Ramalho, T.C. Evaluation of autophagy inhibition to combat cancer: (vanadium complex)-protein interactions, parameterization, and validation of a new force field. J. Mol. Model. 2023, 29, 123. [Google Scholar] [PubMed]
  95. Yan He, Y.; Sun, M.M.; Zhang, G.G.; Yang, J.; Chen, K.S.; Xu, W.W.; Li, B. Targeting PI3K/Akt signal transduction for cancer therapy. Signal Transduct. Target Ther. 2021, 6, 425. [Google Scholar]
  96. Marzban, L.; Rahimian, R.; Brownsey, R.W.; McNeill, J.H. Mechanisms by which bis(maltolato)oxovanadium(IV) normalizes phosphoenolpyruvate carboxykinase and glucose-6-phosphatase expression in streptozotocin-diabetic rats in vivo. Endocrinology 2020, 143, 4636–4645. [Google Scholar]
  97. Kiersztan, A.; Modzelewska, A.; Jarzyna, R.; Jagielska, E.; Bryla, J. Inhibition of gluconeogenesis by vanadium and metformin in kidney-cortex tubules isolated from control and diabetic rabbits. Biochem. Pharmacol. 2002, 63, 1371–1382. [Google Scholar] [CrossRef]
  98. Pan, C.-J.; Lei, K.-J.; Annabi, B.; Hemrika, W.; Chou, J.Y. Transmembrane topology of glucose-6-phosphatase. J. Biol. Chem. 1998, 273, 6144–6148. [Google Scholar] [CrossRef]
  99. Facchini, D.M.; Yuen, V.G.; Battell, M.L.; McNeill, J.H.; Grynpas, M.D. The effects of vanadium treatment on bone in diabetic and non-diabetic rats. Bone 2006, 38, 368–377. [Google Scholar]
  100. Laizé, V.; Tiago, D.M.; Aureliano, M.; Cancela, M.L. New insights into mineralogenic effects of vanadate. Cell Mol. Life Sci. 2009, 66, 3831–3836. [Google Scholar] [CrossRef]
  101. König, M.A.; Gautschi, O.P.; Simmen, H.-P.; Filgueira, L.; Cadosch, D. Influence of vanadium 4+ and 5+ ions on the differentiation and activation of human osteoclasts. Int. J. Biomater. 2017, 2017, 9439036. [Google Scholar] [CrossRef]
  102. Lima, L.M.; da Silva, A.K.J.P.F.; Batista, E.K.; Postal, K.; Kostenkova, K.; Fenton, A.; Crans, C.C.; Silva, W.E.; Belian, M.F.; Lira, E.C. The antihyperglycemic and hypolipidemic activities of a sulfur-oxidovanadium (IV) complex. J. Inorg. Biochem. 2023, 24, 112127. [Google Scholar] [CrossRef] [PubMed]
  103. Batista, E.K.; de Lima, L.M.A.; Gomes, D.A.; Crans, D.C.; Silva, W.E.; Belian, M.F.; Lira, E.C. Dexamethasone-induced insulin resistance attenuation by oral sulfur–oxidovanadium(IV) complex treatment in mice. Pharmaceuticals 2024, 17, 760. [Google Scholar] [CrossRef] [PubMed]
  104. Crans, D.C. Antidiabetic, chemical, and physical properties of organic vanadates as presumed transition-state inhibitors for phosphatases. J. Org. Chem. 2015, 80, 11899–11915. [Google Scholar] [CrossRef] [PubMed]
  105. Lyonnet, B.; Martz, X.; Martin, E. L’emploi thérapeutique des dérivés du vanadium. La Presse Médicale 1899, 32, 191–192. [Google Scholar]
  106. Rehder, D. Vanadium. Its role for humans. In Interrelations Between Essential Metal Ions and Human Diseases, 1st ed.; Sigel, A., Sigel, H., Sigel, R.K.O., Eds.; Springer Nature: Dordrecht, The Netherlands, 2013; Volume 1, pp. 139–169. [Google Scholar]
  107. Heyliger, C.E.; Tahiliani, A.G.; McNeil, J.H. Effect of vanadate on elevated blood glucose and depressed cardiac performance of diabetic rats. Science 1985, 227, 1474–1477. [Google Scholar] [CrossRef]
  108. Zhao, Q.; Chen, D.; Liu, P.; Wei, T.; Zhang, F.; Ding, W. Oxidovanadium(IV) sulfate-induced glucose uptake in HepG2 cells through IR/Akt pathway and hydroxyl radicals. J. Inorg. Biochem. 2015, 149, 39–44. [Google Scholar] [CrossRef]
  109. Begum, A.; Vani, K.; Husain, A.; Chinnagalla, T.; Kumar, M.P.; Swapna, S.; Ayodhya, D.; Shaik, A. A comprehensive review of anti-diabetic activity of vanadium-based complexes via PTP-1B inhibition mechanism. Res. Chem. 2023, 6, 101154. [Google Scholar] [CrossRef]
  110. Bhanot, S.; Michoulas, A.; McNeill, J.H. Antihypertensive effects of vanadium compounds in hyperinsulinemic, hypertensive rats. Mol. Cell. Biochem. 1995, 153, 205–209. [Google Scholar] [CrossRef]
  111. McNeill, J.H.; Yuen, V.G.; Hoveyda, H.R.; Orvig, C. Bis(maltolato)oxovanadium(IV) is a potent insulin mimic. J. Med. Chem. 1992, 35, 1489–1491. [Google Scholar] [CrossRef]
  112. Dinu, V.; Kilic, A.; Wang, Q.; Ayed, C.; Fadel, A.; Harding, S.E.; Yakubov, G.E.; Fisk, I.D. Policy, toxicology and physicochemical considerations on the inhalation of high concentrations of food flavour. npj Sci. Food 2020, 4, 15. [Google Scholar] [CrossRef]
  113. Bordbar, A.-K.; Creagh, A.L.; Mohammadi, F.; Haynes, C.A.; Orvig, C. Calorimetric studies of the interaction between the insulin-enhancing drug candidate bis(maltolato)oxovanadium(IV) (BMOV) and human serum apo-transferrin. J. Inorg. Biochem. 2009, 103, 643–647. [Google Scholar] [CrossRef] [PubMed]
  114. Ramanadham, S.; Mongold, J.J.; Brownsey, R.W.; Cros, G.H.; McNeill, J.H. Oral vanadyl sulfate in the treatment of diabetes mellitus in rats. Am. J. Physiol. 1989, 257, H904–H911. [Google Scholar] [CrossRef] [PubMed]
  115. Thompson, K.H.; Liboiron, B.D.; Bellman, Y.S.K.D.D.; Setyawati, I.A.; Patrick, B.O.; Karunaratne, V.; Rawji, G.; Wheeler, J.; Sutton, K.; Cassidy, S.B.C.; et al. Preparation and characterization of vanadyl complexes with bidentate maltol-type ligands; in vivo comparisons of anti-diabetic therapeutic potential. J. Biol. Inorg. Chem. 2003, 8, 66–74. [Google Scholar] [CrossRef] [PubMed]
  116. Shechter, Y.; Goldwaser, I.; Mironchik, M.; Fridkin, M.; Gefel, D. Historic perspective and recent developments on the insulin-like actions of vanadium; toward developing vanadium-based drugs for diabetes. Coord. Chem. Rev. 2003, 237, 3–11. [Google Scholar] [CrossRef]
  117. Srivastava, A.K.; Mehdi, M.Z. Insulino-mimetic and anti-diabetic effects of vanadium compounds. Diabet. Med. 2005, 22, 2–13. [Google Scholar] [CrossRef]
  118. Aviva Levina, A.; McLeod, A.I.; Pulte, A.; Aitken, J.B.; Lay, P.A. Biotransformations of Antidiabetic Vanadium Prodrugs in Mammalian Cells and Cell Culture Media: A XANES Spectroscopic Study. Inorg. Chem. 2015, 54, 6707–6718. [Google Scholar] [CrossRef]
  119. Schmid, A.C.; Byrne, R.D.; Vilar, R.; Woscholski, R. Bisperoxovanadium compounds are potent PTEN inhibitors. FEBS Lett. 2004, 566, 35–38. [Google Scholar] [CrossRef]
  120. Xiong, Z.; Xing, C.; Xu, T.; Yang, Y.; Liu, G.; Hu, G.; Cao, H.; Zhang, C.; Guo, X.; Yang, F. Vanadium induces oxidative stress and mitochondrial quality control disorder in the heart of ducks. Front. Vet. Sci. 2021, 8, 756534. [Google Scholar] [CrossRef]
  121. Xu, J.; Gong, G.; Huang, X.; Du, W. Schiff base oxovanadium complexes resist the assembly behavior of human islet amyloid polypeptide. J. Inorg. Biochem. 2018, 186, 60–69. [Google Scholar] [CrossRef]
  122. Aureliano, M.; De Sousa-Coelho, A.L.; Dolan, C.C.; Roess, D.A.; Crans, D.C. Biological consequences of vanadium effects on formation of reactive oxygen species and lipid peroxidation. Int. J. Mol. Sci. 2023, 24, 5382. [Google Scholar] [CrossRef]
  123. Tsave, O.; Petanidis, S.; Kioseoglou, E.; Yavropoulou, M.P.; Yovos, J.G.; Anestakis, D.; Tsepa, A.; Salifoglou, A. Role of vanadium in cellular and molecular immunology: Association with immune-related inflammation and pharmacotoxicology mechanisms. Oxid. Med. Cell. Longev. 2016, 2016, 4013639. [Google Scholar] [CrossRef] [PubMed]
  124. Scior, T.; Abdallah, H.H.; Mustafa, S.F.Z.; Guevara-García, J.A.; Rehder, D. Are vanadium complexes druggable against the main protease Mpro of SARS-CoV-2?—A computational approach. Inorg. Chem. Comm. 2024, 161, 112014. [Google Scholar]
  125. Martins, F.M.; Iglesias, B.A.; Chaves, O.A.; da Silva, J.L.G.; Leal, D.B.R.; Back, D.F. Vanadium(v) complexes derived from triphenylphosphonium and hydrazides: Cytotoxicity evaluation and interaction with biomolecules. Dalton Trans. 2024, 53, 8315–8327. [Google Scholar] [CrossRef] [PubMed]
  126. Dias, D.M.; Rodruigues, J.P.L.M.; Domingues, N.S.; Bonvin, A.M.J.J.; Castro, M.M.C.A. Unveiling the interaction of vanadium compounds with human serum albumin by using 1H STD NMR and computational docking studies. Eur. J. Inorg. Chem. 2013, 26, 4619–4627. [Google Scholar] [CrossRef]
  127. Kondaparthy, V.; Shaik, A.; Reddy, K.B.; Das Manwal, D. Studies on interaction of vanadium metal complexes with bovine serum albumin—Fluoremetric and UV–visible spectrophotometric studies. Chem. Data Collect. 2019, 20, 100203. [Google Scholar] [CrossRef]
  128. Zhang, Q.; Ma, Y.; Liu, H.; Gu, J.; Sun, X. Comparison of the effects on bovine serum albumin induced by different forms of vanadium. Biol. Trace Elem. Res. 2023, 201, 3088–3098. [Google Scholar] [CrossRef]
  129. Heinemann, G.; Fichtl, B.; Vogt, W. Pharmacokinetics of vanadium in humans after intravenous administration of a vanadium containing albumin solution. Brit. J. Clin. Pharmacol. 2003, 55, 241–245. [Google Scholar] [CrossRef]
  130. Heinemann, G.; Fichtl, B.; Mentler, M.; Vogt, W. Binding of vanadate to human albumin in infusion solutions, to proteins in human fresh frozen plasma, and to transferrin. J. Inorg. Biochem. 2002, 90, 38–42. [Google Scholar] [CrossRef]
  131. Cobbina, E.; Mehtab, S.; Correia, I.; Gonçalves, G.; Tomaz, I.; Cavaco, I.; Jakusch, T.; Enyedi, E.; Kiss, T.; Pessoa, J.C. Binding of oxovanadium(IV) complexes to blood serum albumins. J. Mex. Chem. Soc. 2013, 57, 180–191. [Google Scholar] [CrossRef]
  132. Sanna, D.; Bíró, L.; Buglyó, P.; Micera, G.; Garribba, E. Transport of the anti-diabetic VO2+ complexes formed by pyrone derivatives in the blood serum. J. Inorg. Biochem. 2012, 115, 87–99. [Google Scholar] [CrossRef]
  133. Soares, M.A.G.; Souza-Silva, F.; Alves, C.R.; Vazquez, L.; de Araujo, T.S.; Serpa, C.; Chaves, O.A. Evidence of hyperglycemic levels improving the binding capacity between human serum albumin and the antihypertensive drug hydrochlorothiazide. Sci. Pharm. 2024, 92, 32. [Google Scholar] [CrossRef]
  134. Lakowicz, J.R. Principles of Fluorescence Spectroscopy, 3rd ed.; Springer: Boston, MA, USA, 2006. [Google Scholar]
  135. Fioravanço, L.P.; Porto, J.B.; Martins, F.M.; Siqueira, J.D.; Iglesias, B.A.; Rodrigues, B.M.; Chaves, O.A.; Back, D.F. A Vanadium (V) complexes derived from pyridoxal/salicylaldehyde. Interaction with CT-DNA/HSA, and molecular docking assessments. J. Inorg. Biochem. 2023, 239, 112070. [Google Scholar] [CrossRef] [PubMed]
  136. Xu, J.; Zhang, B.; Gong, G.; Huang, X.; Du, W. Inhibitory effects of oxidovanadium complexes on the aggregation of human islet amyloid polypeptide and its fragments. J. Inorg. Biochem. 2019, 197, 110721. [Google Scholar] [CrossRef] [PubMed]
  137. He, Z.; Han, S.; Zhu, H.; Hu, X.; Li, X.; Hou, C.; Wu, C.; Xie, Q.; Li, N.; Du, X.; et al. The protective effect of vanadium on cognitive impairment and the neuropathology of Alzheimer’s disease in APPSwe/PS1dE9 mice. Front. Mol. Neurosci. 2020, 13, 21. [Google Scholar] [CrossRef]
  138. Gao, C.; Jiang, J.; Tan, Y.; Chen, S. Microglia in neurodegenerative diseases: Mechanism and potential therapeutic targets. Signal Transduct. Target Ther. 2023, 8, 359. [Google Scholar] [CrossRef]
  139. Guo, S.; Wang, H.; Yin, Y. Microglia polarization from M1 to M2 in neurodegenerative diseases. Front. Aging Neurosci. 2022, 14, 815347. [Google Scholar] [CrossRef]
  140. Wendimu, M.Y.; Hooks, S.B. Microglia phenotypes in aging and neurodegenerative diseases. Cells 2022, 11, 2091. [Google Scholar] [CrossRef]
  141. Almeida, Z.L.; Brito, R.M.M. Amyloid disassembly: What can we learn from chaperones? Biomedicines 2022, 10, 3276. [Google Scholar] [CrossRef]
  142. Kim, E.; Cho, S. Microglia and monocyte-derived macrophages in stroke. Neurotherapeutics 2016, 13, 702–718. [Google Scholar] [CrossRef]
  143. Ferrisi, R.; Gado, F.; Ricardi, C.; Polini, B.; Manera, C.; Chiellini, G. The interplay between cannabinoid receptors and microglia in the pathophysiology of Alzheimer’s disease. J. Clin. Med. 2023, 12, 7201. [Google Scholar] [CrossRef]
  144. Almeida, Z.L.; Brito, R.M.M. Structure and aggregation mechanisms in amyloids. Molecules 2020, 25, 1195. [Google Scholar] [CrossRef] [PubMed]
  145. Yenari, M.A.; Kauppinen, T.M.; Swanson, R.A. Microglial activation in stroke: Therapeutic targets. Neurotherapeutics 2010, 7, 378–391. [Google Scholar] [CrossRef] [PubMed]
  146. Hampel, H.; Hardy, J.; Blennow, K.; Chen, C.; Perry, G.; Kim, S.H.; Villemagne, V.L.; Aisen, P.; Vendruscolo, M.; Iwatsubo, T.; et al. The amyloid-β pathway in Alzheimer’s disease. Mol. Psychiatry 2021, 26, 5481–5503. [Google Scholar] [CrossRef] [PubMed]
  147. Bent, R.; Moll, L.; Grabbe, S.; Bros, M. Interleukin-1 beta—A friend or foe in malignancies? Int. J. Mol. Sci. 2018, 19, 2155. [Google Scholar] [CrossRef]
  148. Grebenciucova, E.; Van Haerents, S. Interleukin 6: At the interface of human health and disease. Front. Immunol. 2023, 14, 1255533. [Google Scholar] [CrossRef]
  149. Culic, O.; Erakovic, V.; Parnham, M.J. Anti-inflammatory effects of macrolide antibiotics. Eur. J. Pharmacol. 2001, 429, 209–229. [Google Scholar] [CrossRef]
  150. Brierly, G.; Celentano, A.; Breik, O.; Moslemivayeghan, E.; Patini, R.; McCullough, M.; Yap, T. Tumour necrosis factor alpha (TNF-α) and oral squamous cell carcinoma. Cancers 2023, 15, 1841. [Google Scholar] [CrossRef]
  151. Jang, D.-i.; Lee, A.-H.; Shin, H.-Y.; Song, H.-R.; Park, J.-H.; Kang, T.-B.; Lee, S.-R.; Yang, S.-H. The role of tumor necrosis factor alpha (TNF-α) in autoimmune disease and current TNF-α inhibitors in therapeutics. Int. J. Mol. Sci. 2021, 22, 2719. [Google Scholar] [CrossRef]
  152. Caminero, A.; Comabella, M.; Montalban, X. Tumor necrosis factor alpha (TNF-α), anti-TNF-α and demyelination revisited: An ongoing story. J. Neuroimmunol. 2011, 234, 1–6. [Google Scholar] [CrossRef]
  153. You, K.; Gu, H.; Yuan, Z.; Xu, X. Tumor necrosis factor alpha signaling and organogenesis. Front. Cell Dev. Biol. 2021, 9, 727075. [Google Scholar] [CrossRef]
  154. Jaspers, I.; Samet, J.M.; Erzurum, S.; Reed, W. Vanadium-induced kappa B-dependent transcription depends upon peroxide-induced activation of the p38 mitogen-activated protein kinase. Am. J. Respir. Cell. Mol. Biol. 2000, 23, 95–102. [Google Scholar] [CrossRef] [PubMed]
  155. Huang, C.; Chen, N.; Ma, W.Y.; Dong, Z. Vanadium induces AP-1- and NFkappB-dependent transcription activity. Int. J. Oncol. 1998, 13, 711–715. [Google Scholar] [CrossRef] [PubMed]
  156. Pisano, M.; Arru, C.; Serra, M.; Galleri, G.; Sanna, D.; Garribba, E.; Palmieri, G.; Rozzo, C. Antiproliferative activity of vanadium compounds: Effects on the major malignant melanoma molecular pathways. Metallomics 2019, 11, 1687–1699. [Google Scholar] [CrossRef] [PubMed]
  157. Chen, F.; Demers, L.M.; Vallyathan, V.; Ding, M.; Lu, Y.; Castranova, V.; Shi, X. Vanadate induction of NF-κB involves IκB kinase β and SAPK/ERK kinase 1 in macrophages. J. Biol. Chem. 1999, 274, 20307–20312. [Google Scholar] [CrossRef]
  158. Ghalichi, F.; Saghafi-Asl, M.; Kafil, B.; Faghfouri, A.H.; Jourshari, M.R.; Naserkiadeh, A.A.; Ostadrahimi, A. Insulin receptor substrates regulation and clinical responses following vanadium-enriched yeast supplementation in obese type 2 diabetic patients: A randomized, double-blind, placebo-controlled clinical trial. Biol. Trace Elem. Res. 2023, 201, 5169–5182. [Google Scholar] [CrossRef]
  159. Young, K.A.; Biggins, L.; Sharpe, H.J. Protein tyrosine phosphatases in cell adhesion. Biochem. J. 2021, 478, 1061–1083. [Google Scholar] [CrossRef]
  160. Abdelsalam, S.S.; Korashy, H.M.; Zeidan, A.; Agouni, A. The role of protein tyrosine phosphatase (PTP)-1B in cardiovascular disease and its interplay with insulin resistance. Biomolecules 2019, 9, 286. [Google Scholar] [CrossRef]
  161. Yang, H.; Wang, L.; Shigley, C.; Yang, W. Protein tyrosine phosphatases in skeletal development and diseases. Bone Res. 2022, 10, 10. [Google Scholar] [CrossRef]
  162. Stanford, S.M.; Ahmed, V.; Barrios, A.M.; Bottini, N. Cellular biochemistry methods for investigating protein tyrosine phosphatases. Antioxid. Redox Signal. 2014, 20, 2160–2178. [Google Scholar] [CrossRef]
  163. Crean, R.M.; Biler, M.; van der Kamp, M.W.; Hengge, A.C.; Kamerlin, S.C.L. Loop dynamics and enzyme catalysis in protein tyrosine phosphatases. J. Am. Chem. Soc. 2021, 143, 3830–3845. [Google Scholar] [CrossRef]
  164. Denu, J.M.; Tanner, K.G. Specific and reversible inactivation of protein tyrosine phosphatases by hydrogen peroxide:? Evidence for a sulfenic acid intermediate and implications for redox regulation. Biochemistry 1998, 37, 5633–5642. [Google Scholar] [CrossRef] [PubMed]
  165. Tanner, J.J.; Parsons, Z.D.; Cummings, A.H.; Zhou, H.; Gates, K.S. Redox regulation of protein tyrosine phosphatases: Structural and chemical aspects. Antioxid. Redox Signal. 2011, 15, 77–97. [Google Scholar] [CrossRef] [PubMed]
  166. Meng, T.-C.; Fukada, T.; Tonks, N.K. Reversible oxidation and inactivation of protein tyrosine phosphatases in vivo. Mol. Cell. 2002, 9, 387–399. [Google Scholar] [CrossRef] [PubMed]
  167. Groen, A.; Lemeer, S.; van der Wijk, T.; Overvoorde, J.; Heck, A.J.R.; Ostman, A.; Barford, D.; Slijper, M.; den Hertog, J. Differential Oxidation of protein-tyrosine phosphatases. J. Biol. Chem. 2005, 280, 10298–10304. [Google Scholar] [CrossRef] [PubMed]
  168. Andersen, J.N.; Mortensen, O.H.; Peters, G.H.; Drake, P.G.; Iversen, L.F.; Olsen, O.H. Structural and evolutionary relationships among protein tyrosine phosphatase domains. Mol. Cell. Biol. 2001, 21, 7117–7136. [Google Scholar] [CrossRef]
  169. Irving, E.; Stoker, A.W. Vanadium compounds as PTP inhibitors. Molecules 2017, 22, 2269. [Google Scholar] [CrossRef]
  170. McLauchlan, C.C.; Peters, B.J.; Willsky, G.R.; Crans, D.C. Vanadium-phosphatase complexes: Phosphatase inhibitors favor the trigonal bipyramidal transition state geometries. Coord. Chem. Rev. 2015, 301, 163–199. [Google Scholar] [CrossRef]
  171. Parente, J.E.; Naso, L.G.; Jori, K.; Franca, C.A.; Ferreira, A.M.C.; Williams, P.A.M.; Ferrer, E.G. In vitro experiments and infrared spectroscopy analysis of acid and alkaline phosphatase inhibition by vanadium complexes. New J. Chem. 2019, 43, 17603. [Google Scholar] [CrossRef]
  172. Huyer, G.; Liu, S.; Kelly, J.; Moffat, J.; Payette, P.; Kennedy, B.; Tsaprailis, G.; Gresser, M.J.; Ramachandran, C. Mechanism of inhibition of protein-tyrosine phosphatases by vanadate and pervanadate. J. Biol. Chem. 1997, 272, 843–851. [Google Scholar] [CrossRef]
  173. Morinville, A.; Maysinger, D.; Shaver, A. From Vanadis to Atropos: Vanadium compounds as pharmacological tools in cell death signalling. Trends Pharmacol. Sci. 1998, 19, 452–460. [Google Scholar] [CrossRef]
  174. Shaik, A.; Kondaparthy, V.; Aveli, R.; Vijjulatha, M.; Kanth, S.S.; Manwal, D.D. Interaction of vanadium metal complexes with protein tyrosine phosphatase-1B enzyme along with identification of active site of enzyme by molecular modeling. Inorg. Chem. Comm. 2021, 126, 108499. [Google Scholar] [CrossRef]
  175. Shaik, A.; Kondaparthy, V.; Aveli, R.; Vemulapalli, L.; Manwal, D.D. Vanadium metal complexes’ inhibition studies on enzyme PTP-1B and antidiabetic activity studies on Wistar rats. Appl. Organom. Chem. 2022, 36, e6710. [Google Scholar] [CrossRef]
  176. Krejsa, C.M.; Nadler, S.G.; Esselstyn, J.M.; Kavanagh, T.J.; Ledbetter, J.A.; Schieven, G.L. Role of oxidative stress in the action of vanadium phosphotyrosine phosphatase inhibitors. J. Biol. Chem. 1997, 272, 11541–11549. [Google Scholar] [CrossRef] [PubMed]
  177. Siqueira, J.D.; Menegatti, A.C.O.; Terenzi, H.; Pereira, M.B.; Roman, D.; Rosso, E.F.; Piquini, P.C.; Iglesias, A.B.; Back, D.F. Synthesis, characterization and phosphatase inhibitory activity of dioxidovanadium(V) complexes with Schiff base ligands derived from pyridoxal and resorcinol. Polyhedron 2017, 130, 184–194. [Google Scholar] [CrossRef]
  178. Liu, R.; Mathieu, C.; Berthelet, J.; Zhang, W.; Dupret, J.-M.; Rodrigues Lima, F. Human protein tyrosine phosphatase 1B (PTP1B): From structure to clinical inhibitor perspectives. Int. J. Mol. Sci. 2022, 23, 7027. [Google Scholar] [CrossRef]
  179. Peters, K.G.; Davis, M.G.; Howard, B.W.; Pokross, M.; Rastogi, V.; Diven, C.; Greis, K.D.; Eby-Wilkens, E.; Maier, M.; Evdokimov, A.; et al. Mechanism of insulin sensitization by BMOV (bis maltolato oxo vanadium); unliganded vanadium (VO4) as the active component. J. Inorg. Biochem. 2002, 96, 321–330. [Google Scholar] [CrossRef]
  180. Martins, P.G.A.; Mori, M.; Chiaradia-Delatorre, L.D.; Menegatti, A.C.O.; Mascarello, A.; Botta, B.; Benítez, J.; Gambino, D.; Terenzi, H. Exploring oxidovanadium(IV) complexes as YopH inhibitors: Mechanism of action and modeling studies. Med. Chem. Lett. 2015, 6, 1035–1040. [Google Scholar] [CrossRef]
  181. Kelland, L. The resurgence of platinum-based cancer chemotherapy. Nature Rev. Cancer 2007, 7, 573–584. [Google Scholar] [CrossRef]
  182. Romani, A.M.P. Cisplatin in cancer treatment. Biochem. Pharmacol. 2022, 206, 115323. [Google Scholar] [CrossRef]
  183. Tsvetkova, D.; Ivanova, S. Application of approved cisplatin derivatives in combination therapy against different cancer diseases. Molecules 2022, 27, 2466. [Google Scholar] [CrossRef]
  184. Gallardo-Vera, F.; Tapia-Rodriguez, M.; Diaz, D.; Van der Goes, T.F.; Montano, L.F.; Rendon-Huerta, E.P. Vanadium pentoxide increased PTEN and decreased SHP1 expression in NK-92MI cells, affecting PI3K-AKT-mTOR and Ras-MAPK pathways. J. Immunotox. 2018, 15, 1–11. [Google Scholar] [CrossRef] [PubMed]
  185. Gonçalves, A.P.; Videira, A.; Soares, P.; Máximo, V. Orthovanadate-induced cell death in RET/PTC1-harboring cancer cells involves the activation of caspases and altered signaling through PI3K/Akt/mTOR. Life Sci. 2011, 89, 371–377. [Google Scholar] [CrossRef]
  186. Haifeng, Z.; Yinghou, W.; Dan, L.; Xiuqing, S.; Furong, W. Vanadium rutin complex sensitizes breast cancer cells via modulation of p53/Bax/Bcl2/VEGF correlated with apoptotic events. Acta Pol. Pharm. 2020, 77, 89–98. [Google Scholar] [CrossRef] [PubMed]
  187. Guerrero-Palomo, G.; Rendón-Huerta, E.P.; Montaño, L.F.; Fortoul, T.I. Vanadium compounds and cellular death mechanisms in the A549 cell line: The relevance of the compound valence. J. Appl. Toxicol. 2019, 39, 540–552. [Google Scholar] [CrossRef] [PubMed]
  188. Kowalski, S.; Ha, S.; Wyrzykowski, D.; Zauszkiewicz-Pawlak, A.; Inkielewicz-Stepniak, I. Selective cytotoxicity of vanadium complexes on human pancreatic ductal adenocarcinoma cell line by inducing necroptosis, apoptosis and mitotic catastrophe process. Oncotarget 2017, 8, 60324–60341. [Google Scholar] [CrossRef] [PubMed]
  189. Zhang, Z.; Huang, C.; Li, J.; Shi, X. Vanadate-induced cell growth arrest is p53-dependent through activation of p21 in C141 cells. J. Inorg. Biochem. 2002, 89, 142–148. [Google Scholar] [CrossRef] [PubMed]
  190. Kowalski, S.; Wyrzykowski, D.; Hac, S.; Rychlowski, M.; Radomski, M.W.; Inkielewicz-Stepniak, I. New oxidovanadium(IV) coordination complex containing 2-methylnitrilotriacetate ligands induces cell cycle arrest and autophagy in human pancreatic ductal adenocarcinoma cell lines. Int. J. Mol. Sci. 2019, 20, 261. [Google Scholar] [CrossRef]
  191. Boscaro, V.; Barge, A.; Deagostino, A.; Ghibaudi, E.; Laurenti, E.; Marabello, D.; Diana, E.; Gallicchio, M. Effects of vanadyl complexes with acetylacetonate derivatives on non-tumor and tumor cell lines. Molecules 2021, 26, 5534. [Google Scholar] [CrossRef]
  192. Lu, L.-P.; Suo, F.Z.; Feng, Y.-L.; Song, L.-L.; Li, Y.; Li, Y.-J.; Wang, K.T. Synthesis and biological evaluation of vanadium complexes as novel anti-tumor agents. Europ. J. Med. Chem. 2019, 176, 1–10. [Google Scholar] [CrossRef]
  193. Sarhan, A.M.; Elsayed, S.A.; Mashaly, M.M.; El-Hendawy, A.M. Oxovanadium(IV) and ruthenium(II) carbonyl complexes of ONS-donor ligands derived from dehydroacetic acid and dithiocarbazate: Synthesis, characterization, antioxidant activity, DNA binding and in vitro cytotoxicity. Appl. Organometal. Chem. 2019, 33, e4655. [Google Scholar] [CrossRef]
  194. Banerjee, A.; Mohanty, M.; Lima, S.; Samanta, R.; Garribba, E.; Sasamori, T.; Dinda, R. Synthesis, structure and characterization of new dithiocarbazate-based mixed ligand oxidovanadium(iv) complexes: DNA/HSA interaction, cytotoxic activity and DFT studies. New J. Chem. 2020, 44, 10946. [Google Scholar] [CrossRef]
  195. Banerjee, A.; Patra, S.A.; Sahu, G.; Sciortino, G.; Pisanu, F.; Garribba, E.; Carvalho, M.F.N.N.; Correia, I.; Pessoa, J.C.; Reuter, H.; et al. A series of non-oxido VIV complexes of dibasic ONS donor ligands: Solution stability, chemical transformations, protein interactions, and antiproliferative activity. Inorg. Chem. 2023, 62, 7932–7953. [Google Scholar] [CrossRef] [PubMed]
  196. Sahu, G.; Patra, S.A.; Mohanty, M.; Lima, S.; Das Pattanayak, P.; Kaminsky, W.; Dinda, R. Dithiocarbazate based oxidomethoxidovanadium(V) and mixed-ligand oxidovanadium(IV) complexes: Study of solution behavior, DNA binding, and anticancer activity. J. Inorg. Biochem. 2022, 233, 111844. [Google Scholar] [CrossRef] [PubMed]
  197. Yekke-ghasemia, Z.; Takjoo, R.; Ramezanib, M.; Maguec, J.T. Molecular design and synthesis of new dithiocarbazate complexes; crystal structure, bioactivities and nano studies. RSC Adv. 2018, 8, 41795–41809. [Google Scholar] [CrossRef] [PubMed]
  198. Chaves, O.A.; Acunha, T.V.; Iglesias, B.A.; Jesus, C.S.H.; Serpa, C. Effect of peripheral platinum(II) bipyridyl complexes on the interaction of tetra-cationic porphyrins with human serum albumin. J. Mol. Liq. 2020, 301, 112466. [Google Scholar] [CrossRef]
  199. Soares, M.A.G.; de Aquino, P.A.; Costa, T.; Serpa, C.; Chaves, O.A. Insights into the effect of glucose on the binding between human serum albumin and the nonsteroidal anti-inflammatory drug nimesulide. Int. J. Biol. Macromol. 2024, 265, 131148. [Google Scholar] [CrossRef]
  200. Fasano, M.; Curry, S.; Terreno, E.; Galliano, M.; Fanali, G.; Narciso, P.; Notari, S.; Ascenzi, P. The extraordinary ligand binding properties of human serum albumin. IUBMB Life 2005, 57, 787–796. [Google Scholar] [CrossRef]
  201. Costa-Tuna, A.; Chaves, O.A.; Almeida, Z.L.; Cunha, R.S.; Pina, J.; Serpa, C. Profiling the interaction between human serum albumin and clinically relevant HIV reverse transcriptase inhibitors. Viruses 2024, 16, 491. [Google Scholar] [CrossRef]
  202. Naveenraj, S.; Anandan, S. Binding of serum albumins with bioactive substances—Nanoparticles to drugs. J. Photochem. Photobiol. C 2013, 14, 53–71. [Google Scholar] [CrossRef]
  203. Chaves, O.A.; Iglesias, B.A.; Serpa, C. Biophysical characterization of the interaction between a transport human plasma protein and the 5,10,15,20-tetra(pyridine-4-yl)porphyrin. Molecules 2022, 27, 5341. [Google Scholar] [CrossRef]
  204. Chaves, O.A.; Oliveira, C.H.C.S.; Ferreira, R.C.; Cesarin-Sobrinho, D.; Machado, A.E.H.; Netto-Ferreira, J.C. Synthetic dimethoxyxanthones bind similarly to human serum albumin compared with highly oxygenated xanthones. Chem. Phys. Imp. 2024, 8, 100411. [Google Scholar] [CrossRef]
  205. Chaves, O.A.; Jesus, C.S.H.; Henriques, E.S.; Brito, R.M.M.; Serpa, C. In situ ultra-fast heat deposition does not perturb the structure of serum albumin. Photochem. Photobiol. Sci. 2016, 15, 1524–1535. [Google Scholar] [CrossRef] [PubMed]
  206. Chaves, O.A.; Jesus, C.S.H.; Cruz, P.F.; Sant’Anna, C.M.R.; Brito, R.M.M.; Serpa, C. Evaluation by fluorescence, STD-NMR, docking and semi-empirical calculations of the o-NBA photo-acid interaction with BSA. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2016, 169, 175–181. [Google Scholar] [CrossRef] [PubMed]
  207. Wani, T.A.; Bakheit, A.H.; Abounassif, M.A.; Zargar, S. Study of interactions of an anticancer drug neratinib with bovine serum albumin: Spectroscopic and molecular docking approach. Front. Chem. 2018, 6, 47. [Google Scholar] [CrossRef]
  208. Chaves, O.A.; Loureiro, R.J.S.; Serpa, C.; Cruz, P.F.; Ferreira, A.B.B.; Netto-Ferreira, J.C. Increasing the polarity of β-lapachone does not affect its binding capacity with bovine plasma protein. Int. J. Biol. Macromol. 2024, 263, 130279. [Google Scholar] [CrossRef]
  209. Wardell, M.; Wang, Z.; Ho, J.X.; Robert, J.; Ruker, F.; Ruble, J.; Carter, D.C. The atomic structure of human methemalbumin at 1.9 A. Biochem. Biophys. Res. Commun. 2002, 291, 913–918. [Google Scholar] [CrossRef]
  210. Bujacz, A. Structures of bovine, equine and leporine serum albumin. Acta Crystallogr. 2012, D68, 1278–1289. [Google Scholar] [CrossRef]
  211. Guizado, T.R.C. Analysis of the structure and dynamics of human serum albumin. J. Mol. Modell. 2014, 20, 2450. [Google Scholar] [CrossRef]
  212. Khashkhashi-Moghadam, S.; Ezazi-Toroghi, S.; Kamkar-Vatanparast, M.; Jouyaeian, P.; Mokaberi, P.; Yazdyani, H.; Amiri-Tehranizadeh, Z.; Saberi, M.R.; Chamani, J. Novel perspective into the interaction behavior study of the cyanidin with human serum albumin-holo transferrin complex: Spectroscopic, calorimetric and molecular modeling approaches. J. Mol. Liq. 2022, 356, 119042. [Google Scholar] [CrossRef]
  213. Acunha, T.V.; Chaves, O.A.; Iglesias, B.A. Fluorescent pyrene moiety in fluorinated C6F5-corroles increases the interaction with HSA and CT-DNA. J. Porphyr. Phthalocyanines 2021, 25, 75–94. [Google Scholar] [CrossRef]
  214. Costa-Tuna, A.; Chaves, O.A.; Loureiro, R.J.S.; Pinto, S.; Pina, J.; Serpa, C. Interaction between a water-soluble anionic porphyrin and human serum albumin unexpectedly stimulates the aggregation of the photosensitizer at the surface of the albumin. Int. J. Biol. Macromol. 2024, 225, 128210. [Google Scholar] [CrossRef] [PubMed]
  215. Paz, E.R.S.; Isoppo, V.G.; dos Santos, F.S.; Machado, L.A.; de Freitas, R.P.; Junior, H.C.S.; Chaves, O.A.; Iglesias, B.A.; Rodembusch, F.S.; Júnior, E.N.S. Imidazole-based optical sensors as a platform for bisulfite sensing and BSA/HSA interaction study. An experimental and theoretical investigation. J. Mol. Liq. 2023, 387, 122666. [Google Scholar] [CrossRef]
  216. Chaves, O.A.; Loureiro, R.J.S.; Costa-Tuna, A.; Almeida, Z.L.; Pina, J.; Brito, R.M.M.; Serpa, C. Interaction of two comercial azobenzene food dyes, amaranth and new coccine, with human serum albumin: Biophysical characterization. ACS Food Sci. Technol. 2023, 3, 955–968. [Google Scholar] [CrossRef]
  217. Gelamo, E.L.; Silva, C.H.T.P.; Imasato, H.; Tabak, M. Interaction of bovine (BSA) and human (HSA) serum albumins with ionic surfactants: Spectroscopy and modelling. Biochim. Biophys. Acta 2002, 1594, 84–99. [Google Scholar] [CrossRef] [PubMed]
  218. Hein, K.L.; Kragh-Hansen, U.; Morth, J.P.; Jeppesen, M.D.; Otzen, D.; Moller, J.V.; Nissen, P. Crystallographic analysis reveals a unique lidocaine binding site on human serum albumin. J. Struct. Biol. 2010, 171, 353–36033. [Google Scholar] [CrossRef] [PubMed]
  219. Martins, F.M.; Siqueira, J.D.; Iglesias, B.A.; Chaves, O.A.; Back, D.F. Pyridoxal water-soluble cobalt(II) helicates: Synthesis, structural analysis, and interactions with biomacromolecules. J. Inorg. Biochem. 2022, 233, 111854. [Google Scholar] [CrossRef] [PubMed]
  220. Da Silveira, C.H.; Chaves, O.A.; Marques, A.C.; Rosa, N.M.P.; Costa, L.A.S.; Iglesias, B.A. Synthesis, photophysics, computational approaches, and biomolecule interactive studies of metalloporphyrins containing pyrenyl units: Influence of the metal center. Eur. J. Inorg. Chem. 2022, 12, e202200075. [Google Scholar] [CrossRef]
  221. Sarmento, C.O.; Pinheiro, B.F.A.; Abrahão, J.; Chaves, O.A.; Moreira, M.B.; Nikolaou, S. Interactions of a ruthenium-ketoprofen compound with human serum albumin and DNA: Insights from spectrophotometric titrations and molecular docking calculations. ChemistrySelect 2022, 7, e202104020. [Google Scholar] [CrossRef]
  222. Tisoco, I.; Donatoni, M.C.; Victória, H.F.V.; de Toledo, J.R.; Krambrock, K.; Chaves, O.A.; de Oliveira, K.T.; Iglesias, B.A. Photophysical, photooxidation, and biomolecule-interaction of meso-tetra(thienyl)porphyrins containing peripheral Pt(II) and Pd(II) complexes. Insights for photodynamic therapy applications. Dalton Trans. 2022, 51, 1646–1657. [Google Scholar] [CrossRef]
  223. Siqueira, J.D.; de Pellegrin, S.F.; Fioravanço, L.P.; Fontana, L.A.; Iglesias, B.A.; Chaves, O.A.; Back, D.F. Self-association synthesis with ortho-vanillin to promote mono- and heptanuclear complexes and their evaluation as antioxidant agents. J. Mol. Struct. 2022, 1256, 132480. [Google Scholar] [CrossRef]
  224. Merlino, A. Metallodrug binding to serum albumin: Lessons from biophysical and structural studies. Coord. Chem. Rev. 2023, 480, 215026. [Google Scholar] [CrossRef]
  225. Park, C.R.; Kim, H.Y.; Song, M.G.; Lee, Y.S.; Youn, H.; Chung, J.K.; Cheon, G.J.; Kang, K.W. Efficacy and safety of human serum albumin-cisplatin complex in U87MG xenograft mouse models. Int. J. Mol. Sci. 2020, 21, 7932. [Google Scholar] [CrossRef] [PubMed]
  226. Wang, H.; Fan, L.; Wu, X.; Han, Y. Efficacy evaluation of albumin-bound paclitaxel combined with carboplatin as neoadjuvant chemotherapy for primary epithelial ovarian cancer. BMC Women’s Health 2022, 22, 224. [Google Scholar] [CrossRef] [PubMed]
  227. Yasuda, Y.; Hattori, Y.; Tohnai, R.; Ito, S.; Kawa, Y.; Kono, Y.; Urata, Y.; Nogami, M.; Takenaka, D.; Negoro, S.; et al. The safety and efficacy of carboplatin plus nanoparticle albumin-bound paclitaxel in the treatment of non-small cell lung cancer patients with interstitial lung disease. Ipn. J. Clin. Oncol. 2018, 48, 89–93. [Google Scholar] [CrossRef] [PubMed]
  228. Paul, M.; Ghosh, B.; Biswas, S. Human Serum Albumin-Oxaliplatin (Pt(IV)) prodrug nanoparticles with dual reduction sensitivity as effective nanomedicine for triple-negative breast cancer. Int. J. Biol. Macromol. 2024, 256, 128281. [Google Scholar] [CrossRef]
  229. Mayr, J.; Heffeter, P.; Groza, D.; Galvez, L.; Koellensperger, G.; Roller, A.; Alte, B.; Haider, M.; Berger, W.; Kowol, C.R.; et al. An albumin-based tumor-targeted oxaliplatin prodrug with distinctly improved anticancer activity in vivo. Chem. Sci. 2017, 8, 2241–2250. [Google Scholar] [CrossRef]
  230. Connolly, K.M.; Stecher, V.J.; Pruden, D.J. Effect of auranofin on plasma fibronectin, C reactive protein, and albumin levels in arthritic rats. Ann. Rheum. Dis. 1988, 47, 515–521. [Google Scholar] [CrossRef]
  231. Pratesi, A.; Cirri, D.; Ciofi, L.; Messori, L. Reactions of auranofin and its pseudohalide derivatives with serum albumin investigated through ESI-Q-TOF MS. Inorg. Chem. 2018, 57, 10507–10510. [Google Scholar] [CrossRef]
  232. Warrell, R.P., Jr.; Israel, R.; Frisone, M.; Snyder, T.; Gaynor, J.J.; Bockman, R.S. Gallium nitrate for acute treatment of cancer-related hypercalcemia. A randomized, double-blind comparison to calcitonin. Ann. Intern. Med. 1988, 108, 669–674. [Google Scholar] [CrossRef]
  233. Coverdale, J.P.C.; Laroiya-McCarron, T.; Romero-Canelón, I. Designing Ruthenium Anticancer Drugs: What Have We Learnt from the Key Drug Candidates? Inorganics 2019, 7, 31. [Google Scholar] [CrossRef]
  234. Acunha, T.V.; Rodrigues, B.M.; da Silva, J.A.; Galindo, D.D.M.; Chaves, O.A.; da Rocha, V.N.; Piquini, P.C.; Köhler, M.H.; De Boni, L.; Iglesias, B.A. Unveiling the photophysical, biomolecule binding and photo-oxidative capacity of novel Ru(II)-polypyridyl corroles: A multipronged approach. J. Mol. Liq. 2021, 340, 117223. [Google Scholar] [CrossRef]
  235. Chaves, O.A.; Menezes, L.B.; Iglesias, B.A. Multiple spectroscopic and theoretical investigation of meso-tetra-(4-pyridyl)porphyrin ruthenium(II) complexes in HSA binding studies. Effect of Zn(II) in protein binding. J. Mol. Liq. 2019, 294, 111581. [Google Scholar] [CrossRef]
  236. Santos, F.S.; da Silveira, C.H.; Nunes, F.S.; Ferreira, D.C.; Victória, H.F.V.; Krambrock, K.; Chaves, O.A.; Rodembusch, F.S.; Iglesias, B.A. Photophysical, photodynamical, redox properties and BSA interactions of novel isomeric tetracationic peripheral palladium(ii)-bipyridyl porphyrins. Dalton Trans. 2020, 49, 16278–16295. [Google Scholar] [CrossRef] [PubMed]
  237. Bessega, T.; Chaves, O.; Martins, F.; Acunha, T.; Back, D.; Iglesias, B.; de Oliveira, G. Coordination of Zn(II), Pd(II) and Pt(II) with ligands derived from diformylpyridine and thiosemicarbazide: Synthesis, structural characterization, DNA/BSA binding properties and molecular docking analysis. Inorg. Chim. Acta 2019, 496, 119049. [Google Scholar] [CrossRef]
  238. Chaves, O.A.; de Oliveira, M.C.C.; de Salles, C.M.C.; Martins, F.M.; Iglesias, B.A.; Back, D.F. In vitro tyrosinase, acetylcholinesterase, and HSA evaluation of dioxidovanadium (V) complexes: An experimental and theoretical approach. J. Inorg. Biochem. 2019, 200, 110800. [Google Scholar] [CrossRef]
  239. Franklim, T.N.; Freire-de-Lima, L.; Chaves, O.A.; Larocque-de-Freitas, I.F.; Silva Trindade, J.D.; Netto-Ferreira, J.C.; Freire-de-Lima, C.G.; Decote-Ricardo, D.; Previato, J.O.; Mendonça-Previato, L.; et al. Design, synthesis, trypanocidal activity, and studies on human albumin interaction of novel S-alkyl-1,2,4-triazoles. J. Braz. Chem. Soc. 2019, 30, 1378–1394. [Google Scholar] [CrossRef]
  240. Espósito, B.P.; Najja, R. Interactions of antitumoral platinum-group metallodrugs with albumin. Coord. Chem. Rev. 2002, 232, 137–149. [Google Scholar] [CrossRef]
  241. Cho, H.; Jeon, S.I.; Ahn, C.-H.; Shim, M.K.; Kim, K. Emerging albumin-binding anticancer drugs for tumor-targeted drug delivery: Current understandings and clinical translation. Pharmaceutics 2022, 14, 728. [Google Scholar] [CrossRef]
  242. Santos-Rebelo, A.; Kumar, P.; Pillay, V.; Choonara, Y.E.; Eleutério, C.; Figueira, M.; Viana, A.S.; Ascensão, L.; Molpeceres, J.; Rijo, P.; et al. Development and mechanistic insight into the enhanced cytotoxic potential of parvifloron D albumin nanoparticles in EGFR-overexpressing pancreatic cancer cells. Cancers 2019, 11, 1733. [Google Scholar] [CrossRef]
  243. Santos, M.F.A.; Correia, I.; Oliveira, A.R.; Garribba, E.; Pessoa, J.C.; Santos-Silva, T. Vanadium Complexes as Prospective Therapeutics: Structural Characterization of a VIV Lysozyme Adduct. Eur. J. Inorg. Chem. 2014, 2, 3293–3297. [Google Scholar] [CrossRef]
  244. Santos, M.F.A.; Pessoa, J.C. Interaction of Vanadium Complexes with Proteins: Revisiting the Reported Structures in the Protein Data Bank (PDB). Molecules 2023, 28, 6538. [Google Scholar] [CrossRef] [PubMed]
  245. Ferraro, G.; Paolillo, M.; Sciortino, G.; Garribba, E.; Merlino, A. Multiple and Variable Binding of Pharmacologically Active Bis(maltolato)oxidovanadium(IV) to Lysozyme. Inorg. Chem. 2022, 61, 16458–16467. [Google Scholar] [CrossRef] [PubMed]
  246. Paolillo, M.; Ferraro, G.; Cipollone, I.; Garribba, E.; Monti, M.; Merlino, A. Unexpected in crystallo reactivity of the potential drug bis(maltolato)oxidovanadium(IV) with lysozyme. Inorg. Chem. Front. 2024, 11, 6307. [Google Scholar] [CrossRef]
  247. Santos, M.F.A.; Sciortino, G.; Correia, I.; Fernandes, A.C.P.; Santos-Silva, T.; Pisanu, F.; Garribba, E.; Pessoa, J.C. Binding of VIVO2+, VIVOL, VIVOL2 and VVO2L moieties to proteins: X-ray/theoretical characterization and biological implications. Chem. A Eur. J. 2022, 28, e202200105. [Google Scholar] [CrossRef] [PubMed]
  248. Ferraro, G.; Tito, G.; Sciortino, G.; Garribba, E.; Merlino, A. Stabilization and binding of [V4O12]4− and unprecedented [V20O54(NO3)]n− to lysozyme upon loss of ligands and oxidation of the potential drug VIVO(acetylacetonato)2. Angew. Chem. Int. Ed. 2023, 62, e202310655. [Google Scholar] [CrossRef]
Figure 2. The 3D structure with the corresponding zoom representation of the active site (FeV-cofactor) of the vanadium nitrogenase from the bacterium Azobacter vinelandii in a (A) non-catalytic process (PDB code 5N6Y) [58], (B) under catalysis (PDB code 6FEA) [59], and (C) superposition of the two conditions. In the zoom representation, the key amino acid residues in the catalysis are represented by sticks following the color of the corresponding cartoon, while (R)-homocitrate (HCA) and carbonate (CO32−) are in green and purple, respectively. Elements’ colors: oxygen, nitrogen, vanadium, iron, and sulfur are in red, blue, gray, brown, and yellow, respectively. For better interpretation, hydrogen atoms were omitted.
Figure 2. The 3D structure with the corresponding zoom representation of the active site (FeV-cofactor) of the vanadium nitrogenase from the bacterium Azobacter vinelandii in a (A) non-catalytic process (PDB code 5N6Y) [58], (B) under catalysis (PDB code 6FEA) [59], and (C) superposition of the two conditions. In the zoom representation, the key amino acid residues in the catalysis are represented by sticks following the color of the corresponding cartoon, while (R)-homocitrate (HCA) and carbonate (CO32−) are in green and purple, respectively. Elements’ colors: oxygen, nitrogen, vanadium, iron, and sulfur are in red, blue, gray, brown, and yellow, respectively. For better interpretation, hydrogen atoms were omitted.
Futurepharmacol 04 00040 g002
Figure 3. Structural projection of amavadin in solid state (and the calcium(II) aquacomplex) obtained by SC-XRD (CCDC 102668) [64]. Ellipsoids are calculated with a 50% probability. The hydrogen atoms and water of crystallization were omitted for better visualization.
Figure 3. Structural projection of amavadin in solid state (and the calcium(II) aquacomplex) obtained by SC-XRD (CCDC 102668) [64]. Ellipsoids are calculated with a 50% probability. The hydrogen atoms and water of crystallization were omitted for better visualization.
Futurepharmacol 04 00040 g003
Figure 4. The chemical structures for some of the vanadium salts and vanadium-based complexes assayed as insulin-mimicking. The oxygen atoms that are coordinated with the metallic center are represented in red color.
Figure 4. The chemical structures for some of the vanadium salts and vanadium-based complexes assayed as insulin-mimicking. The oxygen atoms that are coordinated with the metallic center are represented in red color.
Futurepharmacol 04 00040 g004
Figure 5. The chemical structures for some of the vanadium(IV)-based compounds assayed as inhibitors of aggregation of hIAPP. Elements’ colors: oxygen, nitrogen, and bromine in red, blue, and gray, respectively.
Figure 5. The chemical structures for some of the vanadium(IV)-based compounds assayed as inhibitors of aggregation of hIAPP. Elements’ colors: oxygen, nitrogen, and bromine in red, blue, and gray, respectively.
Futurepharmacol 04 00040 g005
Figure 6. The chemical structures for some vanadium-based complexes assayed as anticarcinogenic [192,193,194,195,196,197]. Elements’ colors: oxygen, nitrogen, sulfur, and chlorine in red, blue, yellow, and green, respectively.
Figure 6. The chemical structures for some vanadium-based complexes assayed as anticarcinogenic [192,193,194,195,196,197]. Elements’ colors: oxygen, nitrogen, sulfur, and chlorine in red, blue, yellow, and green, respectively.
Futurepharmacol 04 00040 g006
Figure 7. (A) Superposition of the non-bound 3D structures of BSA (PDB code 4F5S) [210] and HSA (PDB code 3JRY) [218]. The 3D structure highlights each subdomain for (B) BSA and (C) HSA with the corresponding three main binding sites.
Figure 7. (A) Superposition of the non-bound 3D structures of BSA (PDB code 4F5S) [210] and HSA (PDB code 3JRY) [218]. The 3D structure highlights each subdomain for (B) BSA and (C) HSA with the corresponding three main binding sites.
Futurepharmacol 04 00040 g007
Figure 8. The chemical structure for some metallodrugs in clinical use or investigation. Elements’ colors: oxygen, nitrogen, chlorine, sulfur, and phosphorus in red, blue, green, yellow, and orange, respectively.
Figure 8. The chemical structure for some metallodrugs in clinical use or investigation. Elements’ colors: oxygen, nitrogen, chlorine, sulfur, and phosphorus in red, blue, green, yellow, and orange, respectively.
Futurepharmacol 04 00040 g008
Figure 9. In silico data reported for the interaction between some vanadium(V) complexes and HSA into subdomain IIA (site I) or subdomain IB (site III) [125,135,238]. The vanadium, oxygen, phosphorus, fluorine, and nitrogen atoms in gray, red, dark orange, light blue, and dark blue, respectively. For better interpretation, hydrogen atoms were omitted.
Figure 9. In silico data reported for the interaction between some vanadium(V) complexes and HSA into subdomain IIA (site I) or subdomain IB (site III) [125,135,238]. The vanadium, oxygen, phosphorus, fluorine, and nitrogen atoms in gray, red, dark orange, light blue, and dark blue, respectively. For better interpretation, hydrogen atoms were omitted.
Futurepharmacol 04 00040 g009
Table 1. Some minerals containing vanadium in composition and their respective oxidation states.
Table 1. Some minerals containing vanadium in composition and their respective oxidation states.
ComplexMineral NameVanadium Oxidation State
Pb5[(VO4)3]ClVanadinite [10]V
K2(UO2)2[VO4]2·3H2OCarnotite [11]V
VS4Patronite [12]IV
K(V,Al)2(AlSi3O10)(OH)2Roscoelite [13]III
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chaves, O.A.; Martins, F.M.; Serpa, C.; Back, D.F. The Role of Vanadium in Metallodrugs Design and Its Interactive Profile with Protein Targets. Future Pharmacol. 2024, 4, 743-774. https://doi.org/10.3390/futurepharmacol4040040

AMA Style

Chaves OA, Martins FM, Serpa C, Back DF. The Role of Vanadium in Metallodrugs Design and Its Interactive Profile with Protein Targets. Future Pharmacology. 2024; 4(4):743-774. https://doi.org/10.3390/futurepharmacol4040040

Chicago/Turabian Style

Chaves, Otávio Augusto, Francisco Mainardi Martins, Carlos Serpa, and Davi Fernando Back. 2024. "The Role of Vanadium in Metallodrugs Design and Its Interactive Profile with Protein Targets" Future Pharmacology 4, no. 4: 743-774. https://doi.org/10.3390/futurepharmacol4040040

APA Style

Chaves, O. A., Martins, F. M., Serpa, C., & Back, D. F. (2024). The Role of Vanadium in Metallodrugs Design and Its Interactive Profile with Protein Targets. Future Pharmacology, 4(4), 743-774. https://doi.org/10.3390/futurepharmacol4040040

Article Metrics

Back to TopTop