Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2023 Oct 30;62(45):18524–18532. doi: 10.1021/acs.inorgchem.3c02716

Spodium Bonds Involving Methylmercury and Ethylmercury in Proteins: Insights from X-ray Analysis and Computations

Sergi Burguera , Akshay Kumar Sahu ‡,§, Antonio Frontera , Himansu S Biswal ‡,§,*, Antonio Bauza †,*
PMCID: PMC10647129  PMID: 37902775

Abstract

graphic file with name ic3c02716_0006.jpg

In this study, the stability, directionality, and physical nature of Spodium bonds (SpBs, an attractive noncovalent force involving elements from group 12 and Lewis bases) between methylmercury (MeHg) and ethylmercury (EtHg) and amino acids (AAs) have been analyzed from both a structural (X-ray analysis) and theoretical (RI-MP2/def2-TZVP level of theory) point of view. More in detail, an inspection of the Protein Data Bank (PDB) reported evidence of noncovalent contacts between MeHg and EtHg molecules and electron-rich atoms (e.g., O atoms belonging to the protein backbone and S atoms from MET residues or the π-systems of aromatic AAs such as TYR or TRP). These results were rationalized through a computational study using MeHg coordinated to a thiolate group as a theoretical model and several neutral and charged electron-rich molecules (e.g., benzene, formamide, or chloride). The physical nature of the interaction was analyzed from electrostatics and orbital perspectives by performing molecular electrostatic potential (MEP) and natural bonding orbital (NBO) analyses. Lastly, the noncovalent interactions plot (NCIplot) technique was used to provide a qualitative view of the strength of the Hg SpBs and compare them to other ancillary interactions present in these systems as well as to shed light on the extension of the interaction in real space. We believe that the results derived from our study will be useful to those scientists devoted to protein engineering and bioinorganic chemistry as well as to expanding the current knowledge of SpBs among the chemical biology community.

Short abstract

Spodium bonding interactions involving methyl and ethylmercury cations in protein structures were studied at the RI-MP2/def2-TZVP level of theory in a combined Protein Data Bank and computational study.

Introduction

Methylmercury (MeHg), an organometallic compound derived from inorganic mercury (iHg) through a process called methylation,1 is commonly synthesized in aquatic environments by certain bacteria and archaea microorganisms.25 This leads to a bioaccumulation and biomagnification of MeHg through the food chain, resulting in its increased concentration in higher trophic levels, including fish and seafood, which are common dietary sources for humans.6,7 MeHg exhibits particular chemical features that facilitate crossing biological barriers, such as the blood–brain barrier,811 thus exerting a broad spectrum of toxicological effects on the nervous system,12 which ultimately lead to cognitive deficits, developmental delays, and neurological disorders. Additionally, recent studies suggest that MeHg exposure may also contribute to neurodegenerative diseases, such as Alzheimer’s and Parkinson’s.1315 Beyond its neurotoxic effects, MeHg can also impact other organ systems, such as kidney function,16 interfere with the cardiovascular system,17 disrupt the endocrine system,18 and weaken the immune response.19 Its environmental impact cannot be underestimated either, since MeHg can also influence ecosystems by altering the behavior, reproduction, and survival of various organisms, thus impacting the overall biodiversity.2022

From a molecular biology perspective, MeHg acts as a disruptive agent of noncovalent interactions crucial for protein folding, stability, and function.23,24 It can bind to thiolate groups in cysteine residues, leading to conformational changes and the misfolding of proteins. In addition, it can affect DNA structure and function through the binding to DNA bases and phosphate backbone, thus provoking DNA damage, strand breaks, and interference with DNA replication and repair mechanisms.25,26 Hence, advancing in understanding the role of MeHg in biological systems is crucial for understanding its impact, assessing risks, and developing strategies for its mitigation.

In this regard, although the molecular anchoring of MeHg in a protein’s cavity is based on the coordination to a CYS residue, the vicinal amino acids (AAs) also need to alter their disposition and disrupt their native network of noncovalent interactions (NCIs) to accommodate MeHg. The term Spodium bond (SpB)27 was recently proposed to classify the NCIs involving group 12 of elements (Zn, Cd, and Hg) when acting as Lewis acids. The biological implications of this novel noncovalent force have been explored by some of us in the case of Zn,28 thus expanding its structural and functional role in biology. SpBs are part of the “σ-hole chemistry”, which involves NCIs from the p-block2933 and, more recently, from the d-block of elements, such as Wolfium (group 6),34 Matere (group 7),35 Osme (group 8),36 and Regium bonds (group 11).37,38 These imply electrophilic regions located on the Lewis acid molecule (usually characterized by a positive electrostatic potential) that favorably interact with a Lewis base (e.g., a lone pair, a π-system, or an anion).39

In this study, our approach consisted on a combination of a Protein Data Bank (PDB)40 survey and an ab initio theoretical study at the RI-MP2/def2-TZVP level of theory to analyze the NCIs responsible for the stabilization of MeHg and ethylmercury (EtHg) (another toxic Hg methylation derivative) in biological systems. To achieve that, an inspection of the PDB revealed 586 contacts involving MeHg and EtHg coordinated to CYS residues and electron-rich atoms (N, O, or S). Using theoretical models, we evaluated the stability and directionality of several selected Hg SpBs gathered from the PDB search. These results were complemented with an ab initio computational study at the RI-MP2/def2-TZVP level of theory. The noncovalent nature of the interaction was assessed using the quantum theory of atoms in molecules (QTAIM), the natural bonding orbital (NBO), and the noncovalent interactions plot (NCIplot) analyses. We believe the results reported herein (i) will assist in increasing the visibility of SpBs among the bioinorganic chemistry community and (ii) provide new insights into the NCIs responsible for the stabilization of MeHg and EtHg in proteins, which might have great impact in the fields of chemical biology and environmental chemistry.

Methods

The RCSB website was accessed in May 2023 to download PDB files containing mercury (Hg), resulting in 694 files with at least one Hg atom at a resolution of up to 4 Å. These were filtered out for methyl or ethylmercury-containing structures, leaving 34 PDB files for further analysis. Our analysis focused on two main criteria. First, we identified the C–Hg–S moiety by examining the 2.5 Å vicinity around the Hg atom to determine sulfur (S) atom binding. Subsequently, we searched for Hg···A SpBs, where A represented nitrogen (N), oxygen (O), or sulfur (S) atoms. The geometric criterion for Hg···A distance was set between 2.5 and 4.5 Å and the minimum distance was kept at 2.5 Å to avoid the overlap between coordination and Spodium bonding interactions. The entire analysis was performed by using a custom Python code developed by us.

Computation of the SpB Energies in Selected PDB Structures

Once the PDB structures were identified, theoretical models were built containing MeHg/EtHg molecules coordinated to a thiolate group (CH3–Hg–SCH3) and the interacting amino acid (composed by the side chain and the backbone atoms, see the Supporting Information for the Cartesian coordinates of the PDB models used). In a later stage, the H atoms from the PDB models were optimized at the PBE041,42-D343/def2-SVP44 level of theory. These optimized geometries were then taken as a starting point for single-point calculations at the RI-MP2/def2-TZVP level of theory to compute the interaction energies given in Table 1. The interaction energies were corrected by using the Boys and Bernardi basis set superposition error (BSSE) counterpoise technique.45

Table 1. List of PDB Codes Retrieved from the Search (PDB ID) Including the Interacting Partners (SpB Donor and AA), the Resolution of the Structure (in Å), the BSSE-Corrected Energies (ΔEBSSE, in kcal/mol), Geometrical Parameters (Distance dA···Hg, in Å and A···Hg–C Angle, in Degree), and the Value of the Density at the BCP That Characterizes the Hg SpBs (ρ × 102) at the RI-MP2/def2-TZVP Level of Theory.

PDB ID SpB donor AA resolution ΔEBSSE dA···Hg A···Hg–C ρ × 102
1EMS EtHg HIS98 (C) 2.8 –4.5 2.976 91.9 1.79
1IRK EtHg MET1109 (S) 2.1 –2.6 3.264 72.8 1.50
1RHY EtHg SER142 (O) 2.3 –2.1 3.446 112.3 0.57
1X8K EtHg ALA316 (O) 2.8 –2.7 3.951 94.4 0.22
3PYK MeHg GLN137 (O) 1.3 –5.3 3.041 92.2 1.27
5LU8 EtHg TYR220 (C) 1.9 –5.4 3.866 92.0 0.54
TRP232 (C) –10.0 3.846 76.6 0.43
6BZI EtHg THR375 (O) 2.4 –5.2 2.951 73.9 1.54
GLU448 (O) –12.4 3.380 110.2 0.63
6PII EtHg GLN217 (O) 1.9 –5.1 2.868 75.1 1.68
1L9A MeHg TYR68 2.9 –1.9 3.489 65.2 a
2D2N MeHg PHE63 3.2 –5.6 3.084 70.2 1.57
5G5N MeHg CYS219 2.3 –6.3 2.993 92.7 1.35
a

In this complex, no A···Hg BCP was found.

Computation of the Spodium Bond Energies Using Fully Optimized Models (Complexes 117)

The interaction energies of all complexes were computed at the RI-MP246/def2-TZVP44 level of theory (also corrected using the BSSE counterpoise technique), which is adequate for the treatment of NCIs involving neutral and charged electron donors.47 The calculations have been performed using the program TURBOMOLE version 7.0.48 by fully optimizing the geometries without imposing any restraints. Only the Cs symmetry point group was imposed during the optimization process (except for complex 14). The interaction energies were calculated using the supermolecule approximation (ΔESpB = ESpB complexEmonomer1Emonomer2).

The MEP (molecular electrostatic potential) surfaces were computed at the RI-MP2/def2-TZVP level of theory by means of the TURBOMOLE 7.0 program and were analyzed using Multiwfn software49 and visualized using the Gaussview 5.0 program.50 The calculations for the wave function analysis51 have been carried out at the RI-MP2/def2-TZVP level of theory also using Multiwfn software. The NBO52 analyses was performed at the HF/def2-TZVP level of theory by means of the NBO 7.0 program.53 Lastly, the NCIplot54 isosurfaces correspond to both favorable and unfavorable interactions, as differentiated by the sign of the second density Hessian eigenvalue and defined by the isosurface color. The color scheme is a red–yellow–green–blue scale with red for repulsive (ρcut+) and blue for attractive (ρcut) NCI interaction density. Yellow and green surfaces correspond to weak repulsive and attractive interactions, respectively. The VMD55 program was used in the visualization of the results from the QTAIM, NBO, and NCIplot analyses.

Results and Discussion

PDB Analysis

Out of the 34 PDB structures containing methyl or ethylmercury, the Hg atoms of 20 structures were coordinated to the cysteine sulfur atom. These 20 structures yielded 586 Spodium bond contacts that satisfied the imposed geometrical criteria. Further analysis of these contacts was conducted, as shown in Figure 1. In Figure 1a, the Hg···A distance distribution demonstrates that the number of contacts increases with distance, as expected due to the increase in volume around the Hg atom and the corresponding higher probability of the presence of another atom. Additionally, the presence of two peaks near 4.0 and 4.25 Å suggests the existence of SpBs centered around the Hg atom. The X–Hg–A angle distribution, where X represents S or C atoms, is shown in Figure 1b. The distribution reveals a peak centered at around 110°, indicating that the perpendicular area to the X–Hg···A plane experiences less steric crowding. These less crowded areas are known to be favorable for the formation of noncovalent interactions, as denoted in the radial distribution plot depicted in Figure 1c. Concretely, the plot reveals two high-density regions centered around 4.0 Å and 110° as well as 4.25 Å and 130°, representing the distance (Hg···A) and angle (X–Hg···A), respectively.

Figure 1.

Figure 1

(a) Distance distribution of Hg···A contacts, (b) X–Hg–A angle distribution (X = C, S), and (c) radial distribution plot between Hg···A distance and X–Hg···A angle. The density scale is normalized with respect to the maximum count, where red represents maximum counts and blue represents minimum counts.

Selected PDB Examples

With the purpose to analyze in more detail the Hg···A (A = N, O, and S) contacts retrieved from the PDB search, three X-ray structures were chosen for computations at the RI-MP2/def2-TZVP level of theory (see Figure 2). First, structure 1IRK(56) (Figure 2a) exhibits a Hg···S SpB (d = 3.264 Å) involving an EtHg molecule and a vicinal methionine residue (MET1109), likely acting as a bridge between two α helices from the tyrosine kinase domain of the human insulin receptor protein. Second, structure 5LU8(57) (Figure 2b) belonging to the human legumain protein exhibits a “sandwiched” CTYR···Hg···CTRP SpB (dTYR = 3.760 Å and dTRP = 3.846 Å) involving two aromatic residues (TYR220 and TRP232) located on a protein loop and an EtHg molecule. Lastly, in the 3PYK(58) structure (Figure 2c) involving the human carbonic anhydrase II, three simultaneous Hg···O SpBs were undertaken involving VAL135, GLU205, and GLN137 backbone carbonyl groups and a MeHg moiety (dO-VAL = 3.361 Å, dO-GLU = 3.312 Å, and dO-GLN = 3.041 Å).

Figure 2.

Figure 2

Spodium bonds (SpBs) in (a) 1IRK, (b) 5LU8, and (c) 3PYK structures. The interactions are magnified inside the square parts of the figure, also including QTAIM and NBO analyses of each SpB complex. In panel (d), the MEP surfaces of the MeHg and EtHg molecules are shown (energy values in kcal/mol at 0.001 au). Distances are measured as the shortest value between the Hg atom and the interacting AA.

In order to rationalize the nature of these interactions from an electrostatics point of view, we also computed the electrostatic potential surfaces of the alkyl Hg derivatives (Figure 2d), showing an electropositive belt around the Hg atoms with a similar electrostatic potential value (+18.2 and +16.3 kcal/mol for MeHg and EtHg, respectively) in line with that obtained for other linear transition-metal coordination complexes.59,60 It is also important to note that although the common disposition of an SpB implies a σ-hole27,28 (involving an antibonding metal–ligand orbital), the lineal geometry observed in the Hg coordination complexes studied herein precluded the presence of a σ-hole, thus resulting in the electropositive belt observed around the Hg atom (resembling a π-hole).

Furthermore, we also computed the QTAIM and NCIplot analyses of the noncovalent complexes present in these three structures, and the results show a bond critical point (BCP) connecting the lone pair donor atom (S and O) or the π-system from the AA to the Hg atom from the MeHg/EtHg moiety, thus characterizing the SpB. In addition, ancillary CH···CH, lone pair–π (lp–π), and hydrogen bond (HB) interactions were also present (as denoted by their corresponding BCPs and bond paths), also contributing to the binding affinities obtained. Finally, the NCIplot analyses accounted for the weak nature and extension in real space of the interaction, as is deduced from the greenish isosurfaces observed between both counterparts.

In Table 1, the interaction energies, geometrical parameters (including interaction distances and angles), and values of the density at the BCP that connects the Hg atom with the electron-rich moiety are shown for a series of selected X-ray structures gathered from the PDB search. These were selected based on the X-ray resolution, the alkylated Hg moiety involved, and the interaction distance and angle values in order to provide a representative view of the interaction. As noted, the energy values obtained are far from a coordination bond energy (ranging between −10.0 and −1.9 kcal/mol) with Hg···A distances comprised between 2.8 and 4.0 Å.

Energetic Study

To get further insights into the Hg SpBs present in these systems, we designed a computational study at the RI-MP2/def2-TZVP level of theory using a set of electron-rich species and a MeHg molecule coordinated to a thiolate group (Figure 3) as theoretical models. The energetic results are shown in Table 2, and from their inspection, several interesting points arise.

Figure 3.

Figure 3

Schematic representation of the compounds and complexes used in this study.

Table 2. BSSE-Corrected Interaction Energies (ΔEBSSE, in kcal/mol), Geometrical Parameters (Distance DA···Hg, in Å and A···Hg–C Angle, in Degree), and the Value of the Density at the Bond Critical Point (ρ × 102) Involving the SpB and the Ancillary Interactions Present in Complexes 117 at the RI-MP2/def2-TZVP Level of Theory.

complex ΔEBSSE DA···Hg A···Hg–C ρ × 102 (SpB) ρ × 102 (ancillary)
1 –6.5 3.191 108.9 1.26 0.85 (HB)/0.77 (CH–π)
2 –4.4 2.967 95.4 1.52 0.80 (HB)
3 –7.9 2.821 93.7 1.96 1.40 (HB)
4 –2.0 3.422 109.5 0.77 a
5 –4.2 3.191 98.1 0.98 0.78 (lp–π)
6 –8.0 2.747 97.9 2.79 0.81 (HB)
7 –5.2 2.884 102.8 2.20 a
8 –0.9 3.420 107.1 0.51 a
9 –4.8 2.841 97.2 2.05 0.54 (HB)/0.54 (HB)
10 –9.6 3.277b 94.9b 1.93 1.80 (HB)/0.71 (CH–π)
11 –7.1 2.797 98.1 2.60 0.94 (HB)/0.51 (HB)
12 –7.2 2.769 87.1 2.15 1.13 (HB)
13 –5.8 3.230 98.9 1.88 0.75 (HB)/0.75 (HB)
14 –11.4 3.007 96.9 1.77 1.06 (HB)/0.61 (CH–π)
15 –9.5 2.680 92.2 2.23 1.02 (HB)
16 –23.1 2.674 98.2 4.65 a
17 –19.6 2.858 97.6 3.84 a
a

In this complex, no BCP was found.

b

Distance and angle measured from the ring centroid.

First, in all cases, attractive interaction energies were obtained (between −23.1 and −0.9 kcal/mol), spread between weak (complex 8) and moderately strong values (complex 16). Also, the equilibrium distances obtained ranged between 2 and 3.5 Å, in line with the selected structures from the PDB search. As expected, those complexes involving Br and Cl (16 and 17) obtained the largest interaction energy values (−23.1 and −19.6 kcal/mol, respectively) in the study. In addition, complex 15 involving BF4 as the electron donor moiety achieved a less favorable interaction energy value due to its lower basicity compared to the monatomic anionic species.

Among the neutral complexes (114), complex 14 involving an indole ring obtained the most favorable SpB energy (−11.4 kcal/mol) owing to the simultaneous establishment of an SpB and an HB with the MeHg molecule (see the QTAIM and NCIplot Analyses section). On the other hand, complex 8 involving the weakest Lewis base (OC) achieved the poorest interaction energy value of the set (−0.9 kcal/mol). Among the π-system donors used (benzene and phenol), complex 10 involving the latter obtained a more favorable interaction energy value (−9.6 kcal/mol) due to (i) its higher π-basicity and (ii) the formation of an ancillary S···HO HB vs a S···CH HB in complex 1 (see Figure S1 for their respective QTAIM and NCIplot analyses).

Also, for the N-donating species (imidazole, HCN, NH3, and pyridine; complexes 5 to 7 and 11), complexes 6 and 11 involving imidazole and pyridine achieved the most favorable interaction energy values (−8.0 and −7.1 kcal/mol, respectively), although being weaker Lewis bases than NH3. This is due to the formation of ancillary HBs involving NH and CH groups from the imidazole and pyridine rings, respectively, and the thiolate group coordinated to MeHg (see Figure S1 for more details). On the other hand, complex 5 involving HCN as electron donor species obtained the lowest energy of this set (−4.2 kcal/mol), as expected.

Finally, among the O donor molecules (formaldehyde, formamide, dimethyl ether, and pyridine-N-oxide), complexes 3 and 12 involving formamide and pyridine-N-oxide obtained the largest interaction energy values (−7.9 and −7.2 kcal/mol, respectively). This was an unexpected result in the case of complex 3; however, a strong HB involving the NH2 group of formamide and the S atom from the Hg moiety was undertaken, thus noticeably contributing to the total stabilization of this supramolecular complex. On the other hand, in complex 12 involving pyridine-N-oxide, this HB involved a CH group, being weaker than that in complex 3 (see the QTAIM and NCIplot Analyses section).

QTAIM and NCIplot Analyses

In Figure 4, the combined QTAIM and NCIplot analyses for some representative complexes are shown (the rest are included in Figure S1), and in all cases, a BCP (red sphere) and a bond path connecting the electron donor and Hg atoms were observed, which characterized the Spodium bonding interactions studied herein. Also, ancillary HBs (in complexes 3, 12, 14 and 15) and lp–π (complex 5) interactions were observed. For instance, in complexes 3, 12, and 14, the HBs involved the NH and CH groups from the formamide and indole moieties and the lone pairs of the S atom coordinated to the Hg center. On the other hand, in the case of complex 12, the lp–π interaction implied a lone pair from the S atom and the π-system of the HCN molecule. Lastly, in complex 15, an ancillary HB interaction was described by the presence of a BCP connecting a F atom from the BF4 moiety and a CH group from the Hg coordination complex. Interestingly, the value of the density at the BCP that characterizes the SpB interaction exhibits a larger magnitude than that for the ancillary HB and lp–π interactions, thus highlighting the directing role of the Hg SpBs as the predominant noncovalent force in the supramolecular complexes studied herein (see Table 2 for the complete list of ρ × 102 values).

Figure 4.

Figure 4

NCIplot analysis and QTAIM distribution of intermolecular bond critical points (BCP in red spheres) and bond paths in complexes 3, 5, 7, 12, 14, and 15. The value of the density at the BCPs characterizing the SpB interaction is also indicated in red. Ancillary interactions with their respective BCP density values are also included. NCIplot surfaces involve only intermolecular contacts between the Sp coordination complex and the electron donor molecule. NCIplot color range −0.04 au ≤ (signλ2)ρ ≤ + 0.04 au. Isosurface value RGD = 0.5 and ρ cutoff 0.04 au.

In Table 3, the values of the Laplacian at the BCP that characterize the Hg SpB (∇2ρ × 100) are shown, resulting in positive values in all cases, as is common in closed shell calculations. Furthermore, the values of the potential (V × 100) and kinetic (G × 100) energy densities lie within the same range in all cases, thus confirming the noncovalent nature of the A···Hg interaction (|Vr|/Gr) ≈ 1.

Table 3. Values of the Laplacian of ρ (∇2ρ × 102, in au), the Potential (V × 102, in au) and Kinetic (G × 102, in au) Energy Densities, and the Total Energy Density (H × 102, in au) Gathered at the BCP That Characterizes the Hg SpB.

complex 2ρ × 102 V × 102 G × 102 H × 102
1 3.67 –0.79 0.86 0.06
2 5.80 –1.17 1.31 0.14
3 7.86 –1.70 1.83 0.13
4 2.48 –0.42 0.52 0.10
5 3.58 –0.64 0.77 0.13
6 9.31 –2.39 2.36 –0.03
7 6.94 –1.73 1.74 0.00
8 2.14 –0.28 0.41 0.13
9 7.66 –1.74 1.83 0.09
10 4.69 –0.79 0.98 0.19
11 8.34 –2.14 2.11 –0.03
12 8.70 –1.94 2.06 0.12
13 4.85 –1.23 1.22 –0.01
14 5.12 –1.26 1.27 0.01
15 10.73 –2.22 2.45 0.23
16 14.11 –4.61 4.07 –0.54
17 10.01 –3.22 2.86 –0.36

Regarding the NCIplot analyses, in all of the cases, a greenish isosurface was obtained between the electron donor molecule and the Hg moiety, thus indicating the presence of weak interactions. Also worth noting is the fact that the portion of the NCIplot isosurface devoted to the Hg SpB is bluish instead of greenish in the case of complexes 3, 7, and 15, thus indicating that the SpB interaction is noticeably stronger than the ancillary HBs, in line with the results obtained from QTAIM analyses. Finally, in the case of complexes 5 and 14, all surfaces exhibited a similar color, thus indicating a similar contribution of the ancillary HB and lp–π interactions and the Hg SpBs (see Figure S2 for the plots regarding the reduced density gradient (RDG) vs the sign(λ2)ρ).

NBO Analysis

To further investigate the participation of orbital contributions in the stabilization of the noncovalent complexes studied, we carried out NBO calculations focusing on the second-order perturbation analysis, which is useful to evaluate donor–acceptor interactions (see Table 4). Among the neutral complexes (1 to 14), the Hg SpBs were characterized by the interaction between either a lone pair (LP) from an O, S, N, or F atom or a bonding (BD) C–C/N–C orbital belonging to the electron-rich species and an antibonding (BD*) Hg–C orbital from the MeHg moiety. The magnitude of these orbital interactions ranges from 0.19 kcal/mol in complex 8 involving OC as an electron donor molecule to 4.88 kcal/mol in complex 6 involving an imidazole ring. On the other hand, in the case of the anionic complexes (1517), the orbital contribution is larger in the case of the monatomic anions Cl and Br, with values of 18.30 and 16.90 kcal/mol, in line with their respective interaction energies. In addition, in complexes 1 to 3, 6, and 12 to 15, ancillary lp–π, HB, and CH–π interactions were also observed, involving (i) an LP from a S atom and an antibonding (BD*) C–C orbital, (ii) an LP from a S atom and an antibonding (BD*) C–H and N–H orbital, (iii) an LP from a F atom and an antibonding (BD*) C–H orbital, and (iv) a bonding (BD) C–C orbital and an antibonding (BD*) C–H orbital. In all of these cases, the magnitude of the orbital interaction is lower than that observed for the Hg SpB, in line with the results derived from the QTAIM and NCIplot Analyses section (Figure 5).

Table 4. Donor and Acceptor NBOs with an Indication of the Second-Order Interaction Energy E(2) in Complexes 117a.

complex type of interaction donor acceptor E(2)
1 SpB BD C–C BD* Hg–C 1.45
lp–π LP S BD* C–C 0.67
2 SpB LP O BD* Hg–C 0.74
HB LP S BD* C–H 0.63
3 SpB LP O BD* Hg–C 2.28
HB LP S BD* N–H 3.76
4 SpB LP C BD* Hg–C 0.89
5 SpB BD N–C BD* Hg–C 0.43
6 SpB LP N BD* Hg–C 4.88
HB LP S BD* C–H 0.58
7 SpB LP N BD* Hg–C 3.47
8 SpB LP O BD* Hg–C 0.19
9 SpB LP O BD* Hg–C 0.77
10 SpB BD C–C BD* Hg–C 0.86
11 SpB LP N BD* Hg–C 4.17
12 SpB LP O BD* Hg–C 1.37
HB LP S BD* C–H 2.28
13 SpB LP S BD* Hg–C 3.57
HB LP S BD* C–H 0.62
14 SpB BD C–C BD* Hg–C 1.41
HB LP S BD* N–H 1.11
CH–π BD C–C BD* C–H 0.24
15 SpB LP F BD* Hg–C 2.57
HB LP F BD* C–H 0.75
16 SpB LP Cl BD* Hg–C 18.30
17 SpB LP Br BD* Hg–C 16.90
a

LP, BD, and BD* stand for lone pair, bonding orbital, and antibonding orbital, respectively. Energy values are in kcal/mol.

Figure 5.

Figure 5

Graphical representation of the donor–acceptor orbitals involved in the formation of complexes 3, 5, 7, 12, 14, and 15. LP, BD, and BD* stand for lone pair, bonding orbital, and antibonding orbital, respectively. Energy values are in kcal/mol.

Conclusions

In conclusion, we demonstrated the presence of Hg SpBs involving MeHg and EtHg in proteins using a combination of X-ray analysis and theoretical calculations at the RI-MP2/def2-TZVP level of theory. The PDB survey revealed a preference of the SpB at 4.0 and 110° as well as 4.25 and 130°, representing the distance (Hg···A) and angle (X–Hg···A), respectively. This is in alignment with the expected less hindered region in the X–Hg···A plane. Besides, a variety of electron donor molecules was used to analyze the physical nature and extension in real space of the Hg SpB interaction (including O, S, and π-systems from aromatic residues) in a computational study. These computations were complemented with QTAIM and NCIplot analyses, which are utilized to further understand the weak nature of the interaction from a charge–density perspective as well as with NBO analyses, which highlighted the main orbital contributions responsible for the stabilization of the SpBs studied herein. We expect that the results derived from our study will be useful to those scientists devoted to protein engineering and bioinorganic chemistry as well as to expand the current knowledge of the SpBs among the chemical biology community.

Acknowledgments

S.B., A.B., and A.F. thank the MICIU/AEI (Project PID2020-115637GB-I00 FEDER funds) for financial support. The authors thank the CTI (UIB) for computational facilities. A.K.S. and H.S.B. acknowledge financial support from the Department of Atomic Energy, Department of Science and Technology (Project File No. CRG/2022/001096), Government of India.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.3c02716.

  • Figure S1 including additional QTAIM and NCIplot analyses and Figure S2 showing the NCI plots of reduced gradient vs sign(λ2)ρ and Cartesian coordinates of complexes 117 and of the PDB models used (PDF)

Author Contributions

S.B. and A.K.S. contributed equally to this work. Most of the computational studies were conducted by S.B. and A.F. A.K.S. and H.S.B. conducted the database analyses, and A.B. wrote the article and directed the study.

The authors declare no competing financial interest.

Supplementary Material

ic3c02716_si_001.pdf (373.2KB, pdf)

References

  1. Lehnherr I.; St Louis V. L.; Hintelmann H.; Kirk J. L. Methylation of inorganic mercury in polar marine waters. Nat. Geosci. 2011, 4, 298–302. 10.1038/ngeo1134. [DOI] [Google Scholar]
  2. Raposo J. C.; Ozamiz G.; Etxebarria N.; Tueros I.; Muñoz A.; Muela I.; Arana I.; Barcina I. Mercury biomethylation assessment in the estuary of Bilbao (North of Spain). Environ. Pollut. 2008, 156, 482–488. 10.1016/j.envpol.2008.01.017. [DOI] [PubMed] [Google Scholar]
  3. Hamelin S.; Planas D.; Amyot M. Mercury methylation and demethylation by periphyton biofilms and their host in a fluvial wetland of the St. Lawrence River (QC, Canada). Sci. Total Environ. 2015, 512–513, 464–471. 10.1016/j.scitotenv.2015.01.040. [DOI] [PubMed] [Google Scholar]
  4. Cao D.; Chen W.; Xiang Y.; Qianfen M.; Hang L.; PengYu F.; Hong S.; Cheng Z.; Yongmin W.; Dingyong W. The efficiencies of inorganic mercury bio-methylation by aerobic bacteria under different oxygen concentrations. Ecotoxicol. Environ. Saf. 2021, 207, 111538 10.1016/j.ecoenv.2020.111538. [DOI] [PubMed] [Google Scholar]
  5. Parks J. M.; Johs A.; Podar M.; Bridou R.; Hurt R. A. Jr.; Smith S. D.; Tomanicek S. J.; Qian Y.; Brown S. D.; Brandt C. C.; Palumbo A. V.; Smith J. C.; Wall J. D.; Elias D. A.; Liang L. Y. The genetic basis for bacterial mercury methylation. Science 2013, 339, 1332–1335. 10.1126/science.1230667. [DOI] [PubMed] [Google Scholar]
  6. Mergler D.; Anderson H. A.; Chan L. H. M.; Mahaffey K. R.; Murray M.; Sakamoto M.; Stern A. H. Methylmercury exposure and health effects in humans: A worldwide concern. Ambio 2007, 36, 3–11. 10.1579/0044-7447(2007)36[3:MEAHEI]2.0.CO;2. [DOI] [PubMed] [Google Scholar]
  7. Van Oostdam J.; Donaldson S. G.; Feeley M.; Arnold D.; Ayotte P.; Bondy G.; Chan L.; Dewaily E.; Furgal C. M.; Kuhnlein H.; Loring E.; Muckle G.; Myles E.; Receveur O.; Tracy B.; Gill U.; Kalhok S. Human health implications of environmental contaminants in Arctic Canada: A review. Sci. Total Environ. 2005, 351–352, 165–246. 10.1016/j.scitotenv.2005.03.034. [DOI] [PubMed] [Google Scholar]
  8. Kerper L. E.; Ballatori N.; Clarkson T. W. Methylmercury transport across the blood-brain barrier by an amino acid carrier. Am. J. Physiol. – Regul. Integr. Comp. Physiol. 1992, 262, R761–R765. 10.1152/ajpregu.1992.262.5.R761. [DOI] [PubMed] [Google Scholar]
  9. Aschner M.; Aschner J. L. Mercury neurotoxicity: mechanisms of blood-brain barrier transport. Neurosci. Biobehav. Rev. 1990, 14, 169–176. 10.1016/S0149-7634(05)80217-9. [DOI] [PubMed] [Google Scholar]
  10. Lohren H.; Bornhorst J.; Fitkau R.; Pohl G.; Galla H.-J.; Schwerdtle T. Effects on and transfer across the blood-brain barrier in vitro—Comparison of organic and inorganic mercury species. BMC Pharmacol. Toxicol. 2016, 17, 63. 10.1186/s40360-016-0106-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Kern J. K.; Geier D. A.; Homme K. G.; Geier M. R. Examining the evidence that ethylmercury crosses the blood-brain barrier. Environ. Toxicol. Pharmacol. 2020, 74, 103312 10.1016/j.etap.2019.103312. [DOI] [PubMed] [Google Scholar]
  12. Azevedo B. F.; Barros; Furieri L.; Maciel Peçanha F.; Wiggers G. A.; Frizera Vassallo P.; Ronacher Simoes M.; Fiorim J.; Rossi de Batista P.; Fioresi M.; Rossoni L.; Stefanon I.; Alonso M. J.; Salaices M.; Vassallo D. V. Toxic Effects of Mercury on the Cardiovascular and Central Nervous Systems. J. Biomed. Biotechnol. 2012, 2012, 949048 10.1155/2012/949048. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Cariccio V. L.; Samà A.; Bramanti P.; Mazzon E. Mercury Involvement in Neuronal Damage and in Neurodegenerative Diseases. Biol. Trace Elem. Res. 2019, 187, 341–356. 10.1007/s12011-018-1380-4. [DOI] [PubMed] [Google Scholar]
  14. Pamphlett R.; Bishop D. P. Mercury is present in neurons and oligodendrocytes in regions of the brain affected by Parkinson’s disease and co-localises with Lewy bodies. PLoS One 2022, 17, e0262464 10.1371/journal.pone.0262464. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. da Silva Raposo R.; Pinto V.; Pinto D.; Moreira R.; Pereira Dias R.; Fontes Ribeiro C. A.; Barreto Oriá R.; Oliveira Malva J. Methylmercury Impact on Adult Neurogenesis: Is the Worst Yet to Come From Recent Brazilian Environmental Disasters?. Front. Aging Neurosci. 2020, 12, 591601 10.3389/fnagi.2020.591601. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Mela M.; Randi M. A. F.; Ventura D. F.; Carvalho C. E. V.; Pelletier E.; Oliveira Ribeiro C. A. Effects of dietary methylmercury on liver and kidney histology in the neotropical fish Hoplias malabaricus. Ecotoxicol. Environ. Saf. 2007, 68, 426–435. 10.1016/j.ecoenv.2006.11.013. [DOI] [PubMed] [Google Scholar]
  17. Genchi G.; Sinicropi M. S.; Carocci A.; Lauria G.; Catalano A. Mercury Exposure and Heart Diseases. Int. J. Environ. Res. Public Health 2017, 14, 74. 10.3390/ijerph14010074. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Tan S. W.; Meiller J. C.; Mahaffey K. R. The endocrine effects of mercury in humans and wildlife. Crit. Rev. Toxicol. 2009, 39, 228–269. 10.1080/10408440802233259. [DOI] [PubMed] [Google Scholar]
  19. Lubick N. IMMUNITY: Mercury Alters Immune System Response in Artisanal Gold Miners. Environ. Health Perspect. 2010, 118, A243. 10.1289/ehp.118-a243. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Al-Sulaiti M. M.; Soubra L.; Al-Ghouti M. A. The Causes and Effects of Mercury and Methylmercury Contamination in the Marine Environment: A Review. Curr. Pollut. Rep. 2022, 8, 249–272. 10.1007/s40726-022-00226-7. [DOI] [Google Scholar]
  21. Gworek B.; Dmuchowski W.; Baczewska-Dąbrowska A. H. Mercury in the terrestrial environment: a review. Environ. Sci. Eur. 2020, 32, 128. 10.1186/s12302-020-00401-x. [DOI] [Google Scholar]
  22. Nogara P. A.; Oliveira C. S.; Schmitz G. L.; Piquini P. C.; Farina M.; Aschner M.; Rocha J. B. T. Methylmercury’s chemistry: From the environment to the mammalian brain. Biochim. Biophys. Acta 2019, 1863, 129284 10.1016/j.bbagen.2019.01.006. [DOI] [PubMed] [Google Scholar]
  23. Stratton A.; Ericksen M.; Harris T. V.; Symmonds N.; Silverstein T. P. Mercury(II) binds to both of chymotrypsin’s histidines, causing inhibition followed by irreversible denaturation/aggregation. Protein Sci. 2017, 26, 292–305. 10.1002/pro.3082. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Ajsuvakova O. P.; Tinkov A. A.; Aschner M.; Rocha J. B. T.; Michalke B.; Skalnaya M. G.; Skalny A. V.; Butnariu M.; Dadar M.; Sarac I.; Aaseth J.; Bjørklundk G. Sulfhydryl groups as targets of mercury toxicity. Coord. Chem. Rev. 2020, 417, 213343 10.1016/j.ccr.2020.213343. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Roque C. R.; Sampaio L. R.; Ito M. N.; Pinto D. V.; Caminha J. S. R.; Nunes P. I. G.; Raposo R. S.; Santos F. A.; Windmöller C. C.; Crespo-Lopez M. E.; Alvarez-Leite J. I.; Oriá R. B.; Pinheiro R. F. Methylmercury chronic exposure affects the expression of DNA single-strand break repair genes, induces oxidative stress, and chromosomal abnormalities in young dyslipidemic APOE knockout mice. Toxicology 2021, 464, 152992 10.1016/j.tox.2021.152992. [DOI] [PubMed] [Google Scholar]
  26. Cediel-Ulloa A.; Yu X.; Hinojosa M.; Johansson Y.; Forsby A.; Broberg K.; Rüegg J. Methylmercury-induced DNA methylation-From epidemiological observations to experimental evidence. Front. Genet. 2022, 13, 993387 10.3389/fgene.2022.993387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Bauzá A.; Alkorta I.; Elguero J.; Mooibroek T. J.; Frontera A. Spodium Bonds: Noncovalent Interactions Involving Group 12 Elements. Angew. Chem., Int. Ed. 2020, 59, 17482–17487. 10.1002/anie.202007814. [DOI] [PubMed] [Google Scholar]
  28. Biswal H. S.; Sahu A. K.; Frontera A.; Bauzá A. Spodium Bonds in Biological Systems: Expanding the Role of Zn in Protein Structure and Function. J. Chem. Inf. Model. 2021, 61, 3945–3954. 10.1021/acs.jcim.1c00594. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Bauzá A.; Frontera A. Aerogen Bonding Interaction: A New Supramolecular Force?. Angew. Chem., Int. Ed. 2015, 54, 7340–7343. 10.1002/anie.201502571. [DOI] [PubMed] [Google Scholar]
  30. Cavallo G.; Metrangolo P.; Milani R.; Pilati T.; Priimagi A.; Resnati G.; Terraneo G. The Halogen Bond. Chem. Rev. 2016, 116, 2478–2601. 10.1021/acs.chemrev.5b00484. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Wang W.; Ji B.; Zhang Y. Chalcogen Bond: A Sister Noncovalent Bond to Halogen Bond. J. Phys. Chem. A 2009, 113, 8132–8135. 10.1021/jp904128b. [DOI] [PubMed] [Google Scholar]
  32. Scheiner S. The Pnicogen Bond: Its Relation to Hydrogen, Halogen, and Other Noncovalent Bonds. Acc. Chem. Res. 2013, 46, 280–288. 10.1021/ar3001316. [DOI] [PubMed] [Google Scholar]
  33. Bauzá A.; Mooibroek T. J.; Frontera A. Tetrel-Bonding Interaction: Rediscovered Supramolecular Force?. Angew. Chem., Int. Ed. 2013, 52, 12317–12321. 10.1002/anie.201306501. [DOI] [PubMed] [Google Scholar]
  34. Bauzá A.; Frontera A. Noncovalent Interactions Involving Group 6 in Biological Systems: The Case of Molybdopterin and Tungstopterin Cofactors. Chem. – Eur. J. 2022, 28, e202201660 10.1002/chem.202201660. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Daolio A.; Pizzi A.; Terraneo G.; Frontera A.; Resnati G. Anion···Anion Interactions Involving σ-Holes of Perrhenate, Pertechnetate and Permanganate Anions. ChemPhysChem 2021, 22, 2281–2285. 10.1002/cphc.202100681. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Daolio A.; Pizzi A.; Calabrese M.; Terraneo G.; Bordignon S.; Frontera A.; Resnati G. Molecular Electrostatic Potential and Noncovalent Interactions in Derivatives of Group 8 Elements. Angew. Chem., Int. Ed. 2021, 60, 20723–20727. 10.1002/anie.202107978. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Stenlid J. H.; Johansson A. J.; Brinck T. σ-Holes and σ-lumps direct the Lewis basic and acidic interactions of noble metal nanoparticles: introducing regium bonds. Phys. Chem. Chem. Phys. 2018, 20, 2676–2692. 10.1039/C7CP06259A. [DOI] [PubMed] [Google Scholar]
  38. Frontera A.; Bauzá A. Regium−π bonds: An Unexplored Link between Noble Metal Nanoparticles and Aromatic Surfaces. Chem. – Eur. J. 2018, 24, 7228–7234. 10.1002/chem.201800820. [DOI] [PubMed] [Google Scholar]
  39. Politzer P.; Murray J. S.; Clark T.; Resnati G. The σ-hole revisited. Phys. Chem. Chem. Phys. 2017, 19, 32166–32178. 10.1039/C7CP06793C. [DOI] [PubMed] [Google Scholar]
  40. Berman H. M.; Westbrook J.; Feng Z.; Gilliland G.; Bhat T. N.; Weissig H.; Shindyalov I. N.; Bourne P. E. Protein Data Bank. Nucleic Acids Res. 2000, 28, 235–242. 10.1093/nar/28.1.235. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Weigend F.; Häser M. RI-MP2: first derivatives and global consistency. Theor. Chem. Acc. 1997, 97, 331–340. 10.1007/s002140050269. [DOI] [Google Scholar]
  42. Schäfer A.; Horn H.; Ahlrichs R. Fully optimized contracted Gaussian-basis sets for atoms Li to Kr. J. Chem. Phys. 1992, 97, 2571–2577. 10.1063/1.463096. [DOI] [Google Scholar]
  43. Balasubramani S. G.; Chen G. P.; Coriani S.; et al. TURBOMOLE: Modular program suite for ab initio quantum-chemical and condensed-matter simulations. J. Chem. Phys. 2020, 152, 184107 10.1063/5.0004635. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Adamo C.; Barone V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6169. 10.1063/1.478522. [DOI] [Google Scholar]
  45. Boys S. F.; Bernardi F. The Calculation of Small Molecular Interactions by the Differences of Separate Total Energies. Some Procedures with Reduced Errors. Mol. Phys. 1970, 19, 553. 10.1080/00268977000101561. [DOI] [Google Scholar]
  46. Ernzerhof M.; Scuseria G. E. Assessment of the Perdew–Burke–Ernzerhof exchange-correlation functional. J. Chem. Phys. 1999, 110, 5029–5036. 10.1063/1.478401. [DOI] [PubMed] [Google Scholar]
  47. Bauzá A.; Alkorta I.; Frontera A.; Elguero J. On the Reliability of Pure and Hybrid DFT Methods for the Evaluation of Halogen, Chalcogen, and Pnicogen Bonds Involving Anionic and Neutral Electron Donors. J. Chem. Theory Comput. 2013, 9, 5201–5210. 10.1021/ct400818v. [DOI] [PubMed] [Google Scholar]
  48. Grimme S.; Antony J.; Ehrlich S.; Krieg H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104 10.1063/1.3382344. [DOI] [PubMed] [Google Scholar]
  49. Lu T.; Chen F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. 10.1002/jcc.22885. [DOI] [PubMed] [Google Scholar]
  50. Dennington R.; Keith T. A.; Millam J. M.. GaussView, Version 5; Semichem Inc.: Shawnee Mission, KS, 2016.
  51. Bader R. F. W. A quantum theory of molecular structure and its applications. Chem. Rev. 1991, 91, 893–928. 10.1021/cr00005a013. [DOI] [Google Scholar]
  52. Weinhold F.; Landis C. R.. Valency and Bonding: A Natural Bond Orbital Donor-Acceptor Perspective; Cambridge University Press: Cambridge, U.K., 2005. [Google Scholar]
  53. Glendening E. D.; Badenhoop J. K.; Reed A. E.; Carpenter J. E.; Bohmann J. A.; Morales C. M.; Karafiloglou P.; Landis C. R.; Weinhold F.. NBO 7.0; Theoretical Chemistry Institute, University of Wisconsin: Madison, 2018.
  54. Contreras-García J.; Johnson E. R.; Keinan S.; Chaudret R.; Piquemal J.-P.; Beratan D. N.; Yang W. NCIPLOT: A Program for Plotting Noncovalent Interaction Regions. J. Chem. Theory Comput. 2011, 7, 625–632. 10.1021/ct100641a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Humphrey W.; Dalke A.; Schulten K. ″VMD - Visual Molecular Dynamics. J. Mol. Graphics 1996, 14, 33–38. 10.1016/0263-7855(96)00018-5. [DOI] [PubMed] [Google Scholar]
  56. Hubbard S. R.; Wei L.; Hendrickson W. A. Crystal structure of the tyrosine kinase domain of the human insulin receptor. Nature 1994, 372, 746. 10.1038/372746a0. [DOI] [PubMed] [Google Scholar]
  57. Zhang Z.; Obianyo O.; Dall E.; Du Y.; Fu H.; Liu X.; Kang S. S.; Song M.; Yu S.-P.; Cabrele C.; Schubert M.; Li X.; Wang J.-Z.; Brandstetter H.; Ye K. Inhibition of delta-secretase improves cognitive functions in mouse models of Alzheimer’s disease. Nat. Commun. 2017, 8, 14740 10.1038/ncomms14740. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Monnard F. W.; Heinisch T.; Nogueira E. S.; Schirmer T.; Ward T. R. Human Carbonic Anhydrase II as a host for piano-stool complexes bearing a sulfonamide anchor. Chem. Commun. 2011, 47, 8238–8240. 10.1039/c1cc10345h. [DOI] [PubMed] [Google Scholar]
  59. de las Nieves Piña M.; Frontera A.; Bauzá A. Regium−π Bonds Are Involved in Protein–Gold Binding. J. Phys. Chem. Lett. 2020, 11, 8259–8263. 10.1021/acs.jpclett.0c02295. [DOI] [PubMed] [Google Scholar]
  60. Burguera S.; Frontera A.; Bauzá A. Regium−π Bonds Involving Nucleobases: Theoretical Study and Biological Implications. Inorg. Chem. 2023, 62, 6740–6750. 10.1021/acs.inorgchem.3c00369. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ic3c02716_si_001.pdf (373.2KB, pdf)

Articles from Inorganic Chemistry are provided here courtesy of American Chemical Society

RESOURCES