Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 4474 2023-03-07 08:31:03 |
2 format Meta information modification 4474 2023-03-08 02:51:11 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Dudev, T.M. Theoretical Evaluations for Designing/Engineering Metalloproteins. Encyclopedia. Available online: https://encyclopedia.pub/entry/41928 (accessed on 19 May 2024).
Dudev TM. Theoretical Evaluations for Designing/Engineering Metalloproteins. Encyclopedia. Available at: https://encyclopedia.pub/entry/41928. Accessed May 19, 2024.
Dudev, Todor Minkov. "Theoretical Evaluations for Designing/Engineering Metalloproteins" Encyclopedia, https://encyclopedia.pub/entry/41928 (accessed May 19, 2024).
Dudev, T.M. (2023, March 07). Theoretical Evaluations for Designing/Engineering Metalloproteins. In Encyclopedia. https://encyclopedia.pub/entry/41928
Dudev, Todor Minkov. "Theoretical Evaluations for Designing/Engineering Metalloproteins." Encyclopedia. Web. 07 March, 2023.
Theoretical Evaluations for Designing/Engineering Metalloproteins
Edit

Almost half of all known proteins contain metal co-factors. Crucial for the flawless performance of a metalloprotein is the selection with high fidelity of the cognate metal cation from the surrounding biological fluids. Therefore, elucidating the factors controlling the metal binding and selectivity in metalloproteins is of particular significance. The knowledge thus acquired not only contributes to better understanding of the intimate mechanism of these events but, also, significantly enriches the researcher’s toolbox that could be used in designing/engineering novel metalloprotein structures with pre-programmed properties. A powerful tool in aid of deciphering the physical principles behind the processes of metal recognition and selectivity is theoretical modeling of metal-containing biological structures.

metalloproteins metal affinity and selectivity dft calculations

1. Intrinsic Physicochemical Properties of the Metal Cations Determine to a Great Extent the Metal Selectivity of the Binding Site

1.1. Competition between Ca2+ and Sr2+ in Calcium Receptors/Signaling Proteins

Strontium (Sr2+), a member of the alkaline earth metal group of the periodic table, is employed in medicine as a diagnostic or therapeutic agent. Its radioactive isotopes 85Sr and 89Sr—with a half-life of ~65 and ~51 days, respectively—are used to follow the Ca2+ kinetics and treat bone cancer sufferers [1][2]. Another strontium formulation, strontium ranelate, containing stable, non-radioactive strontium isotopes, has been shown to exert beneficial effects in treating people with osteoporosis, mostly postmenopausal women [1][3][4][5][6]. It has been shown that strontium exerts a dual beneficiary effect: it reduces the bone degradation and promotes bone anabolism [1][3][4][5][6].
Strontium’s clinical applications stem from its ability to closely imitate biogenic calcium ions (Ca2+) and follow Ca2+-specific pathways involved in cell signaling and bone formation. Both Ca2+ and Sr2+ are spherical, doubly charged “hard” cations that strongly prefer “hard” oxygen-containing ligands (side chains of Asp/Glu, Ser, Asn/Gln and backbone peptide groups). Their ionic radii are similar: 1.0/1.06 Å for Ca2+ and 1.18/1.21 Å for Sr2+ in hexacoordinated/heptacoordinated complexes, respectively [7]. The respective hydration free energies are also quite close: −359.7 kcal/mol for Ca2+ and −329.8 kcal/mol for Sr2+ [8]. In the human body, the two metals behave similarly, both exhibiting distinct bone-seeking properties [1]. They possess flexible coordination/hydration spheres comprising 6 to 9 ligands [9][10]. As a Ca2+-mimetic species, Sr2+ radio-isotopes preferentially accumulate at sites of increased osteogenesis, thus focusing the radiation exposure on the cancerous regions. By mimicking Ca2+, non-radioactive Sr2+ has been postulated to bind and activate the calcium-sensing receptor (CaSR), a representative of the G-protein coupled receptor (GPCR) family [1][3][6][11][12]. CaSR activation triggers a cascade of signaling pathways promoting apoptosis of bone tissue-degrading cells (osteoclasts) and differentiation of bone-synthesizing cells (osteoblasts). In addition to CaSr, Sr2+ can also compete with Ca2+ in binding to proteins such as parvalbumin, alkaline phosphatase, calbindin, Ca2+-sensitive ATPase, and calmodulin [13][14].
The intimate mechanism of Sr2+ activation of CaSR is, however, not fully understood. Several important questions remain: How Ca2+/Sr2+-selective are the metal binding sites of the activated CaSR? How efficiently could the “alien” Sr2+ compete with the native Ca2+ for binding to the receptor? What are the key determinants of the metal affinity/selectivity of CaSR in the activated state? To address these questions, a QM/PCM modeling study has been undertaken [15] and the Gibbs free energies for the Ca2+ → Sr2+ exchange in different dielectric media have been evaluated, as shown in Equation (1):
[Sr2+-aq] + [Ca2+-CaSR] → [Sr2+-CaSR] + [Ca2+-aq]
In Equation (1), [Ca2+/Sr2+-CaSR] and [Ca2+/Sr2+-aq] represent the metal cation bound to receptor ligands inside the binding pockets and unbound outside the binding cavity (in bulk solvent), respectively. A positive free energy for Equation (1) implies a Ca2+-selective site, whereas a negative value suggests a Sr2+-selective one.
The Ca2+/Sr2+-loaded binding pockets of CaSR (Sites 1–4 [6]) have been modeled and their thermodynamic characteristics evaluated [15]. Site 1 is situated in a loop region where backbone peptide groups of Ile81, Ser84, Leu87 and Leu88 orbit the metal cation. The metal center in Site 2 is directly coordinated by the side chain of Thr100 and indirectly (via a water molecule) by the side chain of Asn102. The calcium ion in Site 3 is bound in an outer-shell mode (via water molecules) to Ser302 and Ser303, whereas the side chain of Asp234 and backbone carbonyl groups of Glu231 and Gly557 coordinate the metal cation in Site 4 in an inner-shell fashion. The role of bound Ca2+ ions in activating CaSR has been found to be mostly structural: they stabilize the active state by strengthening the homodimer interactions between membrane-proximal domains [6].
The modeling study reveals that the metal binding sites—although comprising a different number and type of protein ligands, overall structure and charge state—are all selective for Ca2+ over Sr2+. Thus, strontium is predicted to be unable to dislodge the cognate calcium from the respective metal centers. The four binding sites, regardless of their structural differences, exhibit almost equal metal selectivity.
Data analysis suggests that several factors—such as the number and type of protein ligands, charge state of the binding pocket and its solvent exposure—do not seem to play any significant role in governing the competition between Ca2+ and Sr2+ in CaSR. Rather, these are the intrinsic physicochemical properties of the two competing metal species that to a great extent orchestrate the process: Ca2+ has higher charge density than Sr2+ (0.40 vs. 0.27 e/Å3, respectively) and is a better Lewis acid than its bulkier counterpart. As a result, Ca2+ interacts more favorably with the protein ligands than Sr2+, yielding higher absolute value interaction energies. Similar conclusions have been drawn for the competition between Ca2+ and Sr2+ in parvalbumin—a representative of the EF-hand family of proteins involved in calcium signaling [14]. CD and EF heptacoordinated binding sites, comprising Asp, Glu, Ser side chains, water and backbone peptide ligands, have been predicted to be Ca2+/Sr2+ selective. The conclusions have been confirmed by experimental measurements [14].

1.2. Fe2+ vs. Mg2+, Mn2+ and Zn2+ in Non-Heme Iron Proteins

Iron, a redox-active element with oxidation state alternating between +2 and +3 (and sometimes +4), plays a key role in a number of essential biological processes such as respiration, cell division, nitrogen fixation, oxygen transport, nucleotide synthesis, oxidant protection, gene regulation, and protein structure stabilization [16][17]. In mononuclear non-heme iron proteins, the metal cation is usually coordinated to His and Asp/Glu side chains [18]. Typical Fe2+ binding site configuration is His2(Asp/Glu)1, designated as “2-His-1-carboxylate facial triad motif” [19], which has been found in a large group of iron dioxygenases, hydrolases, and synthases [19][20][21][22][23][24]. Other combinations between His and acidic residues exist as well: His1(Asp/Glu)2, His2(Asp/Glu)2 and His3(Asp/Glu)1 [18]. The coordination number of Fe2+ varies between 5 and 6 with water or substrate molecules supplementing the coordination sphere. Inside the cell, Fe2+ faces a competition from other biogenic metal species (i.e., Mg2+, Mn2+, and Zn2+) for binding the protein. Although these metal species are characterized with the same charge and similar ionic radii (RFe2+ = 0.78 Å, RMg2+ = 0.72 Å, RZn2+ = 0.74 Å, and RMn2+ = 0.83 Å for hexacoordinated cations [7]), they possess different ligand affinities (due mostly to varying charge accepting abilities), as reflected in the Irving−Williams series [25]:
Mg2+ < Mn2+ < Fe2+ < Co2+ < Ni2+ < Cu2+ > Zn2+
Magnesium and manganese ions, positioned at the far left-hand side of the series, have weaker ligand affinities than Fe2+. Zinc cations, on the other side, are much stronger binders than their Fe2+ counterparts and form, as a rule, more stable complexes. Several outstanding questions arise:
  • How does the Fe2+ binding site sequesters the “right” (native) cation from the cellular fluids and protect itself from attacks by other biogenic cations such as Mg2+, Mn2+, and Zn2+?
  • What kind of selectivity strategies do iron binding sites employ toward metal cations having different ligand affinities and cytosolic concentrations?
  • What are the key factors governing the metal selectivity in Fe2+ proteins?
These questions have been addressed in a modeling study employing a combined DFT/PCM approach [18].
Results obtained reveal the following trends: (i) Mg2+ and Mn2+ are not able to dislodge Fe2+ from the respective binding sites, as evidenced by the positive free energies of metal substitution in both the gas phase and condensed media. This finding comes as no surprise in view of the weaker ligand affinities of Mg2+ and Mn2+ cations relative to those of the Fe2+ cation. However, Mn2+—being closer in physicochemical properties to Fe2+ than Mg2+ to Fe2+—is a more potent iron contender than Mg2+ (less positive free energies for the Fe2+ → Mn2+ exchange than for the Fe2+ → Mg2+ substitution). (ii) The Fe2+ binding sites, however, are ill protected against attacks by the rival Zn2+ cations, which form more stable complexes and are able to displace Fe2+ from the respective metal centers (negative ΔG values for the Fe2+ → Zn2+). (iii) Solvation does not appear to be a key determinant of the metal selectivity in these systems, as it weakly affects the free energies of metal substitution and does not alter the trends observed in the gas phase.
The theoretical results imply that Mg2+ cannot successfully compete with Fe2+ in these binding sites (relatively high positive free energies evaluated for the Fe2+ → Mg2+ substitution). The major determinant of the high Fe2+/Mg2+ selectivity in these systems is the chemical nature of the contending metal species which confers on Fe2+ higher ligand affinity than on Mg2+.
Divalent manganese and iron cations are neighbors in the Irving−Williams series, exhibiting similar ligand affinities, ion radii (see above), coordination preferences (penta- or hexacoordinated first-shell ligand complexes), and cytosolic concentrations (in the micromolar range [26]), thus appearing to be comparably strong contenders for protein binding sites. The calculations show, not surprisingly, that iron centers, although still preferably binding Fe2+ (the latter being a better complexation agent than Mn2+), are weakly selective for Fe2+ over Mn2+ and are vulnerable to Mn2+ attacks. This is evidenced by positive, but low in absolute value, free energies (just a few kcal/mol) of the Fe2+ → Mn2+ exchange in both the gas phase and protein environment. The poor Fe2+/Mn2+ selectivity—supposedly resulting in the easily surmountable thermodynamic barrier for the Fe2+ → Mn2+ substitution—might, however, be advantageous for the cell metabolism and/or cell survival. Under conditions of Fe2+ deprivation, the iron protein may sequester Mn2+ cations from the surrounding fluids which, due to the close resemblance between the two metal species, might secure uninterrupted cell metabolism [27][28].
The zinc cation, characterized by greater ligand affinity than Fe2+, can outcompete the iron cation and displace it from its binding sites regardless of their composition, structure and solvent exposure (negative free energies for the Fe2+ → Zn2+ substitution in the entire series of complexes. The results are in line with a number of in vitro experiments showing that, indeed, Zn2+ binds to the host protein with much greater affinity than Fe2+ [29][30][31][32][33]. Note that, although the protein preferentially coordinates to Zn2+ in vitro, it binds and is activated by Fe2+ in vivo [29][30][31][32][33][34]. Inside the living cell, since the protein alone is not able to repel the attacks by the rival Zn2+, it is the cell machinery which, by tightly controlling the metal homeostasis and maintaining the free Zn2+ concentration at very low levels (in the picomolar to femtomolar range [26]), turns the balance in favor of Fe2+ whose free cytosolic concentration is in the micromolar range [26].
The theoretical study suggests that the inherent physicochemical properties of the contending metal species, reflected in the Irving−Williams series, are the major factor governing the metal selectivity in the non-heme iron centers. In addition, the free cytosolic concentration of the metal competitors, which correlates inversely with the Irving−Williams series, also affects the process of metal competition in vivo.
Notably, by using theoretical calculations, Kumar and Satpati have also found that the metal affinity of the wild-type and mutant CRISPR-associated protein 1 (with the first-shell E190, H254 and D268 residues lining the metal binding site) is in the order Ca2+ < Mg2+ < Mn2+ which is in agreement with the Irving−Williams series [35].

1.3. Competition between Cr3+ and Fe3+, Fe2+, Mg2+ and Zn2+ in Chromodulin

Chromodulin (low molecular weight chromium-binding substance, LMWCr) is a 1.5 kDa oligopeptide that, in Cr3+-loaded form, plays an essential role in the metabolism of carbohydrates and lipids by interfering with the insulin signaling pathways [36]. It has been implicated in reducing the insulin resistance in type 2 diabetic patients [37][38]. Although chromodulin’s primary and 3D structure have not yet been unraveled, it is known that chromodulin contains only four amino acid types in the ratio of Glu:Gly:Cys:Asp = 4:2:2:2 [39]. An indispensable integral part of the oligopeptide in its active (holo-) form are four chromium cations in the oxidation state of 3+, located in two metal binding sites containing three and one Cr3+ ions (“3 + 1” mode of binding). Structural investigations on holo-chromodulin are not abundant and, to date, only limited information is available about the basic characteristics of the Cr3+ binding sites. The seminal experimental study of Jacquamet et al. sheds light on the following aspects of the metal-occupied binding centers [40]: (i) chromium does not alter its oxidation state upon binding and forms Cr3+ complex with the host chromodulin; (ii) the four Cr3+ cations are clustered into two separated centers containing 3 and 1 metals; (iii) metal cations are six-coordinated surrounded by oxygen-containing ligands arranged in a nearly octahedral fashion; (iv) cysteine side chains, oligomer end groups as well as water molecules appear not to be likely ligands for the metal, nor have sulfur bridges involving cysteines been identified; (v) Cr3+ cations in the trinuclear center are organized in the form of a (not ideal) isosceles triangle with the shorter side intermetallic distance of ~2.79 Å and longer sides lengths of ~3.79 Å; (vi) the metal ion pair forming the shorter side of the triangle is “glued” by hydroxo (but not oxo) bridges, whereas Asp/Glu carboxylate bridges connect these Cr3+ cations with the third, more distant Cr3+.
Note that the paradigm of Cr3+ binding to chromodulin is especially intriguing from both experimental and theoretical points of view, since LMWCr appears to be the only molecule of biochemical importance whose native metal cofactor is Cr3+. This prompts several questions: Why chromium? What are the advantages of binding Cr3+ over other cellular biogenic metal species (e.g., Fe3+, Fe2+, Mg2+, Zn2+)? What factors influence the metal cation competition in chromodulin? These questions have been addressed recently by modeling the holo-chromodulin binding sites (following closely the findings from the experimental structural studies mentioned above) and evaluating the free energies of metal competition between the native Cr3+ and other biologically relevant metal species such as Fe3+, Fe2+, Mg2+ and Zn2+ [41]. A combination of density functional theory (DFT) calculations and polarizable continuum method (PCM) computations has been employed.
Guidelines derived from the experiment (see above) have been used to model the structures of the mono- and trinuclear Cr3+ metal centers. The basic characteristics of the optimized metal complexes are in agreement with the available experimental data: (i) the Cr3+ cations are six-coordinated with oxygen-containing ligands (acetates, backbone amide groups and hydroxyls) orbiting the metals in an octahedral fashion; (ii) neither water nor sulfur-containing ligands coordinate the metal cations; (iii) in the trinuclear sites, the metal cations form an isosceles triangle with two Cr3+ from the shorter side being connected by hydroxo bridges, whereas the third Cr3+ is linked to them by acetate bridges and a hydroxo-bridge. Note that the calculated geometrical parameters of the triangle are in good agreement with those evaluated experimentally [40]: RCr–Cr shorter side (Calc) = 2.79–2.80 Å and RCr–Cr shorter side (Exp) = ~2.79 Å; RCr–Cr longer side (Calc) = 3.58–3.65 Å and RCr–Cr longer side (Exp) = ~3.79 Å.
The experiment does not provide information about the nature of the non-bridging oxygen-containing ligands. Therefore, several metal centers have been constructed comprising various combinations of non-bridging acetates and backbone amides while, at the same time, maintaining the triangle structure with the respective OH and acetate bridges. 
The respective mononuclear binding sites have been modeled in agreement with the composition of their partner trinuclear structures: since the number of carboxylic residues in the host oligopeptide is 6 (2Asp and 4Glu, see above), the number of acetates in the mononuclear constructs has been adjusted to that in the trinuclear center so that the total number of carboxylates in the two binding sites (trinuclear and mononuclear) sums up to 6.
The data presented reveals that the trivalent Fe3+ cannot outcompete Cr3+ in either mononuclear or trinuclear metal centers, as demonstrated by positive ΔGs ranging from 2 to 7 kcal/mol for the former and between 14 and 20 kcal/mol for the latter. Evidently, the trinuclear chromium center is more resistant to Fe3+ attack than its mononuclear counterpart (higher positive ΔGs for the trinuclear structure compared to those for the mononuclear binding site). These findings are not unexpected in view of the higher affinity of Cr3+ to oxygen-containing ligands stabilizing the chromic complexes to a greater extent than the respective ferric structures. As seen from the data collected (bottom two rows), the energies of formation of chromic-single ligand complexes (ligand = species building metal binding sites in chromodulin, i.e., acetate, backbone amide and hydroxyl) are lower (more favorable) than their Fe3+ counterparts. Note, however, that the trend reverses for sulfur- (CH3S) and nitrogen-containing (imidazole) amino-acid residues, which preferably bind Fe3+ over Cr3+. Importantly, such ligands, which would promote Fe3+ over Cr3+ selectivity, do not participate in metal binding in chromodulin, as they are either absent from the amino acid sequence (His) or, even though part of the oligopeptide buildup (Cys), are, apparently, far from the metal binding center [42].
Since the Lewis acidity/complexation power of divalent metals is lower than those of the trivalent cations, it is expected that M2+ metals would be weaker competitors to Cr3+. Indeed, Cr3+ binding sites are very well protected against attacks from divalent biogenic metals: ΔGs of the Cr3+ → M2+ (M = Fe, Mg and Zn) substitution vary between 127 and 165 kcal/mol in mononuclear complexes, and between 251 and 330 kcal/mol in the trinuclear constructs. The inherent physicochemical characteristics of the rival metal species emerge as the key factor controlling the metal competition in LMWCr.

2. Metal Coordination Number Is an Important Determinant of the Metal Selectivity

Sodium (Na+) is an indispensable allosteric regulator in a number of signal-transducing proteins, such as neurotransmitter transporters and G-protein coupled receptors (GPCRs). These have been recognized as drug targets for psychiatric disorders [43][44] and addictive behavior [45]. In the holo-protein, sodium cation(s) are usually penta- or hexacoordinated and predominantly bind to oxygen-containing ligands such as Asp/Glu and Ser/Thr side chains, backbone peptide groups or water [46]. The competition between Na+ and Li+ (non-biogenic metal cation known for its beneficial therapeutic effect on patients with mental disorders) in model sodium-binding sites have been studied by a combined DFT/PCM approach [46], and key determinants controlling the selectivity for Na+ over Li+ in sodium proteins have been elucidated.
Two types of sodium binding sites have been modeled: flexible ones that allow for ligand rearrangement upon Na+ → Li+ exchange; and rigid binding pockets which preserve the original ligand arrangements in the “mother” sodium complex during metal substitution. The results reveal that the coordination number of the competing metals is an important factor in the selectivity process: Li+, when allowed to adopt its preferred tetrahedral ligand arrangement (decreasing its coordination number from 6 to 4), outcompetes the six-coordinated Na+ in the entire dielectric range. On the other hand, in rigid binding sites, competitiveness of the lithium cation decreases (positive free energies in protein environment; blue numbers in parentheses) as it is forced to adopt the unfavorable octahedral ligand surrounding of the native sodium: thus, ligand repulsion between the six bulky ligands around the small Li+ cation attenuates its efficiency of binding. Note that the rigid binding sites preserve the original, relatively large, binding cavity optimized to fit the size of the bulkier Na+ but not the smaller Li+, which additionally decreases the strength of the interactions between Li+ and ligands lining the pore.

3. Adjacent Metal Cation May Reverse the Metal Selectivity in Binuclear/Trinuclear Metal Binding Sites

Nickel-containing enzymes are of vital importance for a number of plants and primitive organisms, such as archaea, bacteria, fungi and low-trophic level marine eucaryota [47][48], where they fulfill various tasks ranging from energy generation to detoxification, oxidative stress protection and virulence [49]. To date, nine nickel-dependent biocatalysts, subdivided into non-redox (urease, glyoxylase I, acireductone dioxygenase and lactate racemase) and redox enzymes (CO-dehydrogenase, acetyl-CoA synthase, [NiFe]-hydrogenase, methyl-SCoM reductase, and Ni-superoxide dismutase), have been identified and characterized [50]. The structure and composition of their metal centers are quite diverse, varying from mononuclear nickel binding sites to homo-binuclear and hetero-binuclear constructs [48][49][50][51]. Cysteine is the predominant amino acid residue in redox enzymes, whereas histidines and aspartates/glutamates are the ligands of choice in the non-redox metaloproteins. Nickel’s valence state in the latter is 2+, while it alternates between 1+, 2+ and 3+ in the former.
Nickel enzymes can be deactivated/inhibited by other metal cations such as Zn2+ (in Escherichia coli glyoxalase I) [52] and Ag+ (in urease) [53][54][55]. The inhibition by Ag+ of urease (a pathogen in humans) is of high significance as it may have strong implications for pharmacology and medicine. The mechanism of silver antibacterial action in urease, however, is still not completely understood. Although the prevailing hypothesis postulates that Ag+ binds to some sulfur containing amino acid ligands (not nickel-bound) at the periphery of the active site which disrupts the enzyme structure [53], the substitution of Ni2+ cations from the metal center by Ag+ cannot be excluded [54][55]. Therefore, it is of special interest to determine to what extent the Ni2+ binding sites in urease (and other nickel enzymatic binding sites) are predisposed to Ni2+ → Ag+ substitution. Of note, information on the factors controlling the competition between the cognate Ni2+ and abiogenic (“alien”) Ag+ in biological systems is critically lacking.
To fill in the gap, the rivalry between Ni2+ and Ag+ in nickel enzymes has been studied by a combined DFT/PCM approach and key determinants of the process have been revealed [56].
Glyoxalase I utilizes intracellular thiols to convert cytotoxic ketoaldehydes, such as methylglyoxal, into nontoxic D-hydroxy acids [50]. The enzyme contains a mononuclear nickel center where two histidine and two glutamate amino acid residues along with two water molecules surround the metal in an octahedral fashion. Upon Ni2+ → Ag+ substitution, the binding site undergoes quite drastic structural changes resulting in a distorted tetrahedral silver complex with an acetate (model for the Glu- side chain) and a water ligand transferred to the second coordination layer of the metal. The free energy calculations suggest that the nickel active center is well protected against Ag+ attack and that the “alien” Ag+ cannot dislodge the native Ni2+ from its binding site as evidenced by highly positive ΔGεs spanning the entire dielectric region. It is of note that the gas-phase exchange reaction, which is entirely dominated by electronic effects (being in favor of the divalent Ni2+), is characterized by quite high positive ΔG1s. These are attenuated in the condensed phase, where solvation effects favor to a greater extent the monovalent Ag+, yielding smaller (although still positive) free energies of metal exchange. This is in line with findings from a similar investigation [56] (not shown here), which demonstrates that increasing the number of charged/polar ligands, surrounding the metal (2 methylimidazoles and 2 acetates representing His and Glu- amino acid side chains, respectively) and donating more charge to divalent Ni2+ than to its monovalent contender, increases the competitiveness of Ni2+ over Ag+.
Acireductone dioxygenase is a mononuclear nickel enzyme involved in the methionine salvage pathway. In the process, methylthioadenosin is transformed into acireductone which, consequently, is converted to formate, carbon monoxide, and methylthiobutyric acid [50]. The nickel binding site comprises three histidines, one glutamate and two water molecules octahedrally arranged around the Ni2+ cation. Upon Ni2+ → Ag+ exchange, the complex isomerizes to a four-coordinated structure with a water molecule and methylimidazole ligand relegated to the metal second coordination sphere. The positive ΔGεs evaluated for the metal substitution reaction suggest that Ag+ cannot outcompete the native Ni2+ in this system (due mainly to the presence of strong charge-donating ligands in the binding cavity) and that the acireductone dioxygenase binding site is reliably shielded from “alien” monocationic attack.
Urease, a nickel-dependent enzyme, catalyzes the hydrolytic decomposition of urea producing ammonia and carbamic acid which, subsequently, decomposes into another molecule of ammonia and carbonic acid [49]. The enzyme has been recognized as a virulence factor in several pathogenic (mostly antibiotic-resistant) bacteria, which makes it a plausible target for antibacterial therapy [53][54][55]. Urease comprises a homo-binuclear active center where the protein donates four histidines (two to each metal), an aspartate (in binding site 2) and a bridging carbamylated lysine to the metal cations.
Incorporating Ag+ in the binding sites, as expected, alters their structure (the silver cation prefers smaller coordination numbers and longer metal-ligand bond distances), although, as shown, in different fashion. Binding site 1 preserves its pentacoordinated structure but alters its shape from almost regular squire pyramidal construct (with the native Ni2+) to a distorted squire pyramidal one (with the “alien” Ag+). Moreover, the Ag1-ligand coordinative bonds are considerably elongated in comparison with the respective Ni1-ligand counterparts: the mean of Ag1-ligand bond distance is 2.488 Å, whereas that of the Ni1-ligand bond distance is 2.068 Å. The binding site 2 structure changes more dramatically upon Ni2+ → Ag+ exchange: it transforms from a hexacoordinated (nearly octahedral) complex to a semi-squire planar complex with two water molecules transferred to the second coordination layer of Ag+.
Thermodynamic evaluations reveal that the first Ag+ → Ni2+ exchange in buried binding pockets (ε = 4) in either site is favorable, and characterized with negative ΔG4s (−4.2 kcal mol−1 for binding site 1, and −2.7 kcal mol−1 for binding site 2. left-hand side). Substituting the second Ni2+ cation with Ag+, however, is thermodynamically unfavorable, as Gibbs free energies for the entire dielectric region are positive. Notably, the calculations suggest that only one mole of metal cations is exchanged during the process.
Why is the Ag+ → Ni2+ substitution in binuclear urease more favorable than that in the mononuclear glyoxalase I and acireductone dioxygenase (see above)? Why is the competitiveness of Ni2+ over Ag+ in bimetallic center compromised? This is mainly due to the presence of a second neighboring metal atom in the active center which, with its positive charge, attenuates the strength of the charge transfer through the bridging carboxylate to the other metal. Indeed, instead of coordinating to the “pure” strong charge-donating anionic carbamylated lysine (represented by CH3CH2NHCOO), the Ni12+ cation, in fact, binds to the metal-bound cationic [CH3CH2NHCOO-Ni22+]+ ligand characterized with a poorer charge-donating power that significantly attenuates the strength of the interaction between the Ni12+ and [CH3CH2NHCOO-Ni22+]+. Note that, in [CH3CH2NHCOO-Ni22+]+, the Ni22+ cation withdraws electron charge density toward itself, thus reducing the amount of charge that the metal-bound carbamylated lysine can donate to the adjacent Ni12+. As a result of the decreased charge-donating strength of the metal-bound carbamylated lysine, the Ni12+–ligand interactions are attenuated to a greater extent than those of the Ag1+–ligand interactions which, as expected, results in reduced Ni12+/Ag1+ competitiveness. The same considerations hold for the Ni22+/Ag2+ exchange as well.

References

  1. Nielsen, S.P. The biological role of strontium. Bone 2004, 35, 583–588.
  2. Kraeber-Bodéré, F.; Campion, L.; Rousseau, C.; Bourdin, S.; Chatal, J.-F.; Resche, I. Treatment of bone metastases of prostate cancer with strontium-89 chloride: Efficacy in relation to the degree of bone involvement. Eur. J. Nucl. Med. 2000, 27, 1487–1493.
  3. Cabrera, W.E.; Schrooten, I.; De Broe, M.E.; D’Haese, P.C. Strontium and bone. J. Bone Miner. Res. 1999, 14, 661–668.
  4. Marie, P.J.; Ammann, P.; Boivin, G.; Rey, C. Mechanisms of Action and Therapeutic Potential of Strontium in Bone. Calcif. Tissue Int. 2001, 69, 121–129.
  5. Cianferotti, L.; Gomes, A.R.; Fabbri, S.; Tanini, A.; Brandi, M.L. The calcium-sensing receptor in bone metabolism: From bench to bedside and back. Osteoporos. Int. 2015, 26, 2055–2071.
  6. Geng, Y.; Mosyak, L.; Kurinov, I.; Zuo, H.; Sturchler, E.; Cheng, T.C.; Subramanyam, P.; Brown, A.P.; Brennan, S.C.; Mun, H.-C.; et al. Structural mechanism of ligand activation in human calcium-sensing receptor. eLife 2016, 5, e13662.
  7. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Cryst. 1976, A32, 751–766.
  8. Marcus, Y. Thermodynamics of solvation of ions. Part 5—Gibbs free energy of hydration at 298.15 K. J. Chem. Soc. Faraday Trans. 1991, 87, 2995–2999.
  9. Dudev, M.; Wang, J.; Dudev, T.; Lim, C. Factors Governing the Metal Coordination Number in Metal Complexes from Cambridge Structural Database Analyses. J. Phys. Chem. B 2006, 110, 1889–1895.
  10. Hofer, T.S.; Randolf, B.R.; Rode, B.M. Sr(II) in Water: A Labile Hydrate with a Highly Mobile Structure. J. Phys. Chem. B 2006, 110, 20409–20417.
  11. Brown, E.M. Is the calcium receptor a molecular target for the actions of strontium on bone? Osteoporos. Int. 2003, 14, S25–S34.
  12. Diepenhorst, N.A.; Leach, K.; Keller, A.N.; Rueda, P.; Cook, A.E.; Pierce, T.L.; Nowell, C.; Pastoureau, P.; Sabatini, M.; Summers, R.J.; et al. Divergent effects of strontium and calcium-sensing receptor positive allosteric modulators (calcimimetics) on human osteoclast activity. Br. J. Pharmacol. 2018, 175, 4095–4108.
  13. Uversky, V.N. Strontium binding to proteins. In Encyclopedia of Metalloproteins; Uversky, V.N., Kretsinger, R.H., Permyakov, E.A., Eds.; Springer Science: New York, NY, USA, 2013; pp. 2133–2138.
  14. Vologzhannikova, A.A.; Shevelyova, M.P.; Kazakov, A.S.; Sokolov, A.S.; Borisova, N.I.; Permyakov, E.A.; Kircheva, N.; Nikolova, V.; Dudev, T.; Permyakov, S.E. Strontium Binding to α-Parvalbumin, a Canonical Calcium-Binding Protein of the “EF-Hand” Family. Biomolecules 2021, 11, 1158.
  15. Cheshmedzhieva, D.; Ilieva, S.; Permyakov, E.A.; Permyakov, S.E.; Dudev, T. Ca2+/Sr2+ Selectivity in Calcium-Sensing Receptor (CaSR): Implications for Strontium’s Anti-Osteoporosis Effect. Biomolecules 2021, 11, 1576.
  16. Elizabeth, C.T.; Uversky, V.N. Encyclopedia of Metalloproteins; Uversky, V.N., Kretsinger, R.H., Permyakov, E.A., Eds.; Springer Science: New York, NY, USA, 2013; p. 1032.
  17. Crain, A.V.; Duschene, K.S.; Peters, J.W.; Broderick, J.B. Iron-Sulfur Cluster Proteins, Fe/SS-adenosylmethionine Enzymes and Hydrogenases. In Encyclopedia of Metalloproteins; Uversky, V.N., Kretsinger, R.H., Permyakov, E.A., Eds.; Springer Science: New York, NY, USA, 2013; pp. 1034–1044.
  18. Dudev, T.; Nikolova, V. Determinants of Fe2+over M2+ (M = Mg, Mn, Zn) Selectivity in Non-Heme Iron Proteins. Inorg. Chem. 2016, 55, 12644–12650.
  19. Costas, M.; Mehn, M.P.; Jensen, M.P.; Que, J. Dioxygen Activation at Mononuclear Non-heme Iron Active Sites: Enzymes, Models and Intermediates. Chem. Rev. 2004, 104, 939–986.
  20. Hausinger, R.P. Fell/alpha-ketoglutarate-dependent Hydroxylases and Related Enzymes. Crit. Rev. Biochem. Mol. Biol. 2004, 39, 21–68.
  21. Solomon, E.I.; Decker, A.; Lehnert, N. Non-heme iron enzymes: Contrasts to heme catalysis. Proc. Natl. Acad. Sci. USA 2003, 100, 3589–3594.
  22. Bruijnincx, P.C.A.; van Koten, G.; Klein Gebbink, R.J.M. Mononuclear Non-heme Iron Enzymes with the 2-His-1 caboxylate Facial Triad: Recent Developments in Enzymology and Modeling Studies. Chem. Soc. Rev. 2008, 37, 2716–2744.
  23. Kovaleva, E.G.; Lipscomb, J.D. Versatility of Biological Nonheme Fe(II) Centers in Oxyden Activation Reactions. Nat. Chem. Biol. 2008, 4, 186–193.
  24. Mbughuni, M.M.; Lipscomb, J.D. Iron Proteins, Mononuclear (non-heme) Iron Oxygenases. In Encyclopedia of Metalloproteins; Uversky, V.N., Kretsinger, R.H., Permyakov, E.A., Eds.; Springer Science: New York, NY, USA, 2013; pp. 1006–1015.
  25. Irving, H.; Williams, R.J.P. Order of Stability of Metal Complexes. Nature 1948, 162, 746–747.
  26. Dudev, T.; Lim, C. Competition among Metal Ions for Protein Binding Sites: Determinants of Metal Ion Selectivity in Proteins. Chem. Rev. 2013, 114, 538–556.
  27. Anjem, A.; Varghese, S.; Imlay, J.A. Manganese import is a key element of the OxyR response to hydrogen peroxide in Escherichia coli. Mol. Microbiol. 2009, 72, 844–858.
  28. Campos-Bermudez, V.A.; Leite, N.R.; Krog, R.; Costa-Filho, A.J.; Soncini, F.C.; Oliva, G.; Vila, A.J. Biochemical and Structural Char-acterization of Salmonella Typhimurium Glyoxalase II: New Insights into Metal Ion Selectivity. Biochemistry 2007, 46, 11069–11079.
  29. Gattis, S.G.; Hernick, M.; Fierke, C.A. Active Site Metal Ion in UDP-3-O-((R)-3-Hydroxymyristoyl)-N-acetylglucosamine Deacetylase (LpxC) Switches between Fe(II) and Zn(II) Depending on Cellular Conditions. J. Biol. Chem. 2010, 285, 33788–33796.
  30. Hernick, M.; Gattis, S.G.; Penner-Hahn, J.E.; Fierke, C.A. Activation of E. coli UDP-3-O-(R-3-hydroxymyristoyl)-N-acetylglucosamine Deacetylase by Fe2+ Yields a More Efficient Enzyme with Altered Ligand Affinity. Biochemistry 2010, 49, 2246–2255.
  31. Kim, B.; Pithadia, A.S.; Fierke, C.A. Kinetics and thermodynamics of metal-binding to histone deacetylase 8. Protein Sci. 2015, 24, 354–365.
  32. Dowling, D.; Gattis, S.G.; Fierke, C.A.; Christianson, D.W. Structures of Metal-Substituted Human Histone Deacetylase 8 Provide Mechanistic Inferences on Biological Function. Biochemistry 2010, 49, 5048–5056.
  33. Becker, A.; Schlichting, I.; Kabsch, W.; Groche, D.; Schultz, S.; Wagner, A.F.V. Iron center, substrate recognition and mechanism of peptide deformylase. Nat. Genet. 1998, 5, 1053–1058.
  34. Zhu, J.; Dizin, E.; Hu, X.; Wavreille, A.-S.; Park, J.; Pei, D. S-Ribosylhomocysteinase (LuxS) Is a Mononuclear Iron Protein. Biochemistry 2003, 42, 4717–4726.
  35. Kumar, A.; Satpati, P. Divalent-Metal-Ion Selectivity of the CRISPR-Cas System-Associated Cas1 Protein: Insights from Classical Molecular Dynamics Simulations and Electronic Structure Calculations. J. Phys. Chem. B 2021, 125, 11943–11954.
  36. Lai, M.-H. Antioxidant Effects and Insulin Resistance Improvement of Chromium Combined with Vitamin C and E Supplementation for Type 2 Diabetes Mellitus. J. Clin. Biochem. Nutr. 2008, 43, 191–198.
  37. Anderson, R.A.; Cheng, N.; Bryden, N.A.; Polansky, M.M.; Chi, J.; Feng, J. Elevated intakes of supplemental chromium improve glucose and insulin variables in individuals with type 2 diabetes. Diabetes 1997, 46, 1786–1791.
  38. Martin, J.; Wang, Z.Q.; Zhang, X.H.; Wachtel, D.; Volaufova, J.; Matthews, D.E.; Cefalu, W.T. Chromium Picolinate Supplementation Attenuates Body Weight Gain and Increases Insulin Sensitivity in Subjects with Type 2 Diabetes. Diabetes Care 2006, 29, 1826–1832.
  39. Vincent, J.B. The Nutritional Biochemistry of Chromium (III), 1st ed.; Elsevier: Amsterdam, The Netherlands, 2007.
  40. Jacquamet, L.; Sun, Y.; Hatfield, J.; Gu, W.; Cramer, S.P.; Crowder, M.W.; Lorigan, G.A.; Vincent, J.B.; Latour, J.-M. Characterization of Chromodulin by X-ray Absorption and Electron Paramagnetic Resonance Spectroscopies and Magnetic Susceptibility Measurements. J. Am. Chem. Soc. 2003, 125, 774–780.
  41. Kircheva, N.; Toshev, N.; Dudev, T. Holo-chromodulin: Competition between the native Cr3+ and other biogenic cations (Fe3+, Fe2+, Mg2+, and Zn2+) for the binding sites. Metallomics 2022, 14, mfac082.
  42. Kooshki, F.; Tutunchi, H.; Vajdi, M.; Karimi, A.; Niazkar, H.R.; Shoorei, H.; Gargari, B.P. A Comprehensive insight into the effect of chromium supplementation on oxidative stress indices in diabetes mellitus: A systematic review. Clin. Exp. Pharmacol. Physiol. 2021, 48, 291–309.
  43. Malinauskaite, L.; Quick, M.; Reinhard, L.; Lyons, J.A.; Yano, H.; Javitch, J.; Nissen, P. A mechanism for intracellular release of Na+ by neurotransmitter/sodium symporters. Nat. Struct. Mol. Biol. 2014, 21, 1006–1012.
  44. Neubig, R.R.; Siderovski, D. Regulators of G-Protein signalling as new central nervous system drug targets. Nat. Rev. Drug Discov. 2002, 1, 187–197.
  45. Amara, S.G.; Sonders, M.S. Neurotransmitter transporters as molecular targets for addictive drugs. Drug Alcohol Depend. 1998, 51, 87–96.
  46. Dudev, T.; Mazmanian, K.; Lim, C. Competition between Li+ and Na+ in sodium transporters and receptors: Which Na+-binding sites are “therapeutic” Li+ targets? Chem. Sci. 2018, 9, 4093–4103.
  47. Mulrooney, S.B.; Hausinger, R. Nickel uptake and utilization by microorganisms. FEMS Microbiol. Rev. 2003, 27, 239–261.
  48. Ragsdale, S.W. Nickel-Based Enzyme Systems. J. Biol. Chem. 2009, 284, 18571–18575.
  49. Boer, J.L.; Mulrooney, S.B.; Hausinger, R.P. Nickel-dependent metalloenzymes. Arch. Biochem. Biophys. 2014, 544, 142–152.
  50. Alfano, M.; Cavazza, C. Structure, function, and biosynthesis of nickel-dependent enzymes. Protein Sci. 2020, 29, 1071–1089.
  51. Maroney, M.J.; Ciurli, S. Nonredox Nickel Enzymes. Chem. Rev. 2014, 114, 4206–4228.
  52. Clugston, S.L.; Barnard, J.F.J.; Kinach, R.; Miedema, D.; Ruman, R.; Daub, E.; Honek, J.F. Overproduction and Characterization of a Dimeric Non-Zinc Glyoxalase I from Escherichia coli: Evidence for Optimal Activation by Nickel Ions. Biochemistry 1998, 37, 8754–8763.
  53. Mazzei, L.; Cianci, M.; Vara, A.G.; Ciurli, S. The structure of urease inactivated by Ag(i): A new paradigm for enzyme inhibition by heavy metals. Dalton Trans. 2018, 47, 8240–8247.
  54. Ambrose, J.F.; Kistiakowsky, G.B.; Kridl, A.G. Inhibition of Urease by Silver Ions. J. Am. Chem. Soc. 1951, 73, 1232–1236.
  55. Krajewska, B. Mono- (Ag, Hg) and di- (Cu, Hg) valent metal ions effects on the activity of jack bean urease. Probing the modes of metal binding to the enzyme. J. Enzym. Inhib. Med. Chem. 2008, 23, 535–542.
  56. Dobrev, S.; Kircheva, N.; Nikolova, V.; Angelova, S.; Dudev, T. Competition between Ag+ and Ni2+ in nickel enzymes: Implications for the Ag+ antibacterial activity. Comput. Biol. Chem. 2022, 101, 107785.
More
Information
Subjects: Biophysics
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 207
Revisions: 2 times (View History)
Update Date: 08 Mar 2023
1000/1000