Abstract
Arsenic is highly toxic and a significant threat to human health, but certain bacteria have developed defense mechanisms initiated by AsIII binding to AsIII-sensing proteins of the ArsR family. The transcriptional regulator AfArsR responds to AsIII and SbIII by coordinating the metalloids with three cysteines, located in a short sequence of the same monomer chain. Here, we characterize the binding of AsIII and HgII to a model peptide encompassing this fragment of the protein via solution equilibrium and spectroscopic/spectrometric techniques (pH potentiometry, UV, CD, NMR, PAC, EXAFS, and ESI-MS) combined with DFT calculations and MD simulations. Coordination of AsIII changes the peptide structure from a random-coil to a well-defined structure of the complex. A trigonal pyramidal AsS3 binding site is formed with almost exactly the same structure as observed in the crystal structure of the native protein, implying that the peptide possesses all of the features required to mimic the AsIII recognition and response selectivity of AfArsR. Contrary to this, binding of HgII to the peptide does not lead to a well-defined structure of the peptide, and the atoms near the metal binding site are displaced and reoriented in the HgII model. Our model study suggests that structural organization of the metal site by the inducer ion is a key element in the mechanism of the metalloid-selective recognition of this protein.
Introduction
Organic and inorganic arsenic compounds in any of their regular oxidation states (+3, + 5, or −3) display toxicity to all lifeforms.1−6 Arsenic compounds have been proposed to interfere with numerous biochemical processes of living cells by, e.g., inhibiting enzymes (pyruvate dehydrogenase (PDH), thioredoxin (Trx), DNA repair enzymes, etc.),7 substituting phosphates,8 and playing roles in the generation of reactive oxygen species, leading to oxidative damage of biomolecules.4−6 It is widely accepted that the molecular mechanism of action of trivalent arsenicals involves the interaction of AsIII with the thiol groups of proteins and enzymes or their cofactors.7,9 The inhibition of the PDH and α-ketoglutarate dehydrogenase (KGDH) enzyme complexes was proposed to be related to the binding of arsenous acid and monomethylarsonous acid to the reduced form of lipoic acid.10 Direct interaction of trivalent arsenicals with the C3H or C4 motifs of zinc finger proteins (e.g., Fpg (formamidopyrimidine DNA glycosylase), XPA (xeroderma pigmentosum group A), PARP-1 (poly(ADP-ribose) polymerase-1)) was suggested to inhibit DNA repair processes.11,12 The binding of various arsenicals to the CGPC peptide segment of Trx was also demonstrated and characterized by different mass spectrometry techniques.13
Inorganic arsenic(III) (iAsIII), i.e., arsenous acid (or arsenite) in aqueous solutions, may form one, two, or three As–S bonds via condensation reactions. Kitchin et al. performed systematic studies on the interaction of arsenous acid with cysteine-containing peptides bearing 0–4 cysteine residues14 and with peptides displaying two cysteines separated by 0–17 intervening amino acids.15 These experiments provided basic information both about the binding affinity of arsenous acid to mono-, di-, and trithiol binding sites and to separated thiols modeling AsIII coordination to more than one ligand, as well as about the association/dissociation kinetics of these binding schemes.16 The relatively high iAsIII-binding affinity of di- and trithiols indicates their greater biochemical relevance, as compared to monothiol coordination.16 There is only a slight extra stabilization promoted by the coordination of a third thiol sulfur donor as compared to dithiolate AsIII complexes. However, tricoordinated sites may be more common in AsIII-binding proteins. Three cysteine-coordinated metalloid centers were identified, e.g., in zinc finger proteins,12,17 metallothioneins,18 the metallochaperone ArsD,19 the AsIII/SbIII translocating ATPase ArsA,20 the key enzyme in arsenic biomethylation S-adenosylmethionine methyltransferase,21 or the metalloid-responsive bacterial metalloregulator ArsR.22−26
Specific proteins classified in the large ArsR protein family27 play a key role in controlling arsenic (or antimony) resistance in bacterial cells through transcription regulation.27−29 The binding affinities of Escherichia coli R773 ArsR,22 AfArsR23 (from Acidithiobacillus ferrooxidans), CgArsR24 (from Corynebacterium glutamicum), and CviArsR (from Chromobacterium violaceum)30 for AsIII (or SbIII) were found to be in the Kd = 10–150 μM range for AsIII (and Kd = 2–10 μM for SbIII). While the selective responses of these ArsR proteins have rarely been tested in in vitro studies of the bare proteins, just as the metalloid selectivity of AfArsR against CdII in promoting the dissociation of the protein from the regulated DNA,23 biosensing assays showed the efficiency of AsIII and SbIII to selectively promote the transcription of reporter genes.31−36 However, the molecular details of the metalloid recognition mechanism of ArsR proteins are still largely unresolved.
Designed three-stranded coiled-coil oligopeptidic constructs offering three Cys thiols for metal(loid) binding, as model systems, were investigated by Pecoraro et al.37−40 AsIII stabilized the formation of the α-helical three-stranded structures, by forming three As–SCys bonds, under conditions where the two-stranded coiled-coil conformation is favored in the absence of metal ions.37 The X-ray structure of a three-stranded Cys-containing Coil Ser (L9C)3 showed that the three Cys residues are locked into their interior conformation by AsIII, which is located at the same side of the three-atom sulfur plane as all three Cys Cβ carbons,38 mimicking the conformation of the ArsR metal centers.25 It is proposed that AsIII binding in this all-endo trigonal pyramidal geometry is stabilized by the optimized Cys rotamers, forming a cavity that is ideal for the small AsIII with its short-length As–S bonds.38 Importantly, there is space below the As center for the As lone pair; thus, steric hindrance by bulky hydrophobic residues from the second coordination sphere is minimized.39,40 Such a small cavity is nonideal for PbII, a soft metal ion also having a lone electron pair but a different size and charge.40 It was also proposed that the hydrophobic layers above and below the 3-Cys planes influence the shape and stability of trigonally coordinated HgII structures, manifested in a difference in the pKa values of the third coordinating thiols.39 AsIII binding by two tripodal pseudopeptides also demonstrated the preferred formation of the all-endo conformers, mimicking well the structure of ArsR AsIII-centers.41 DFT studies indicated not only the importance of the proper cavity size, accommodating the As lone pair, but also the role of electrostatic interactions and H-bonding in the second coordination sphere of AsIII in stabilizing the energetically favored structures.
Here, we present spectroscopic and theoretical studies with the oligopeptide Ac-NCCHGTRDCA-NH2 (L) (Figure 1), encompassing the effector binding fragment of the ArsR metalloid regulator from A. ferrooxidans. The AfArsR protein is particularly interesting for potential biotechnological applications since it is the only ArsR family metalloregulator with a known crystal structure where the three cysteine residues forming an AsS3 site are located within a short sequence at a single monomer chain,25 in contrast to, e.g., CgArsR25 and several other dimeric metalloid regulators where the coordinating donor ligands are recruited from two monomers. AsIII and HgII interaction with the model peptide is explored aiming to characterize AsIII binding to the metalloid site in solution and to elucidate if differences between AsIII and HgII binding may indicate how the selective response for AsIII is achieved.
Figure 1.

Schematic structure of the studied ligand, Ac-NCCHGTRDCA-NH2 (L).
Results and Discussion
Titrations of the Ligand by AsIII and HgII – UV Absorption and CD
The interaction of arsenous acid (AsIII) and the peptide (Figure 1) was studied at different pH values by recording the UV-absorption spectra of the ligand at gradually increasing AsIII:L concentration ratios (Figures 2 and S1). The gradually increasing absorption between 215 and 330 nm is associated with S– → AsIII charge transfer bands, similar to the transitions observed for the binding of AsIII by simple mono-,42 bis-,42−44 or tristhiol41-type compounds. While these bands usually appear as smooth or more definite shoulders, in the spectra recorded for the present system, there is a clearcut local absorption maximum slightly below 290 nm. The observed λmax value (= 287 nm) and the intensity of this band (ε287 nm = 1.91 × 103 M–1 cm–1) unambiguously reflect three AsIII-bound thiolates41,42 and are clearly distinct from the less intense transitions characteristic for species with two thiolates.42−44 The trends of the spectroscopic features (Figures 2 and S1) suggest the formation of the same AsIII–peptide complexes between pH 2.0 and 7.5 and binding of one AsIII per peptide.
Figure 2.
UV spectra recorded at different c(AsIII):c(L) ratios at pH 7.5. The thick lines are highlighted spectra for 0.0 (blue), 1.0 (red), and 5.0 equiv (black) of AsIII per ligand. The inset shows the trend in change in absorbances at 245 and 285 nm using data from two independent measurements (symbols) and the fit (line) of data by the model AsIII + L ⇌ AsL (c(L) = 50.0 μM).
Evaluation of the UV-monitored AsIII titrations of L allowed determining the AsIII binding affinity of the peptide. The fitting of data obtained at four different pH values (pH 2.0, 4.5, 6.0, 7.5; Figures 2 and S1), executed by the PSEQUAD software,45 reflects a slightly increasing stability with pH increase and an AsIII binding affinity of the peptide under the neutral condition: logKpH7.5 = 6.35(3) falling slightly above the upper end of the range reported for multiple Cys-containing oligopeptides.14,15 A comparative analysis of stabilities and the details of calculations are presented in the Supporting Information.
Titration of the peptide by HgII also results in an initial increase of absorbances between 215 and 330 nm, but the trend in the evolution of spectra observed up to 1:1 HgII:L concentration ratios are clearly different at pH 7.5 (Figure 3) and pH 2.0 (Figure S2). This is also demonstrated by the pH-dependent series of spectra recorded for the c(HgII):c(L) 1:1 system (Figure S3). A broad shoulder, centered around 275–280 nm, develops upon adding 1 equiv of HgII per ligand at pH 7.5, closely resembling the S– → HgII charge transfer bands observed for the binding of HgII to various peptides with a {HgS3} coordination mode.46,47 The same tristhiolate donor group environment of the metal ion is further supported by the observed molar absorbances (ε280 nm = 7.77 × 103 M–1 cm–1; ε240 nm = 1.60 × 104 M–1 cm–1). These values are in accordance with those reported for HgII–peptide complexes with the three Cys units in different (CXCXC or CXCXXC) patterns but are also close to the molar absorbances of the {HgS3} species formed with NTA-based tripodal compounds48,49 bearing Cys or d-Pen arms or with three-stranded coiled coils.37,50 Under excess of HgII over L, the broad shoulder collapses, but in parallel, intensities in the low-wavelength regime remarkably increase up to a 1.5:1 HgII:L concentration ratio, potentially reflecting species with a 3:2 HgII:L composition, as found also for other short 3-Cys-containing peptides.46,47 Larger HgII-excess resulted in unreliable absorbance readings, possibly due to aggregation processes or precipitate formation. The detected transformation of spectra between 1.0 and 1.5 equiv of HgII per ligand, leading to a weak absorption above ∼240 nm but an absorbance increase below ca. 225 nm, is in accord with the conversion of the {HgS3} species into a structure with HgII ions coordinated in a {HgS2} environment.46−54
Figure 3.
UV spectra recorded at different c(HgII): c(L) ratios at pH 7.5. The thick lines are highlighted spectra for 0.0 (blue), 1.0 (red), and 1.5 equiv (black) of HgII per ligand. The inset shows the trend in the change of absorbances at 215 and 280 nm (c(L) = 40.0 μM).
At pH 2.0, the shape of the spectra and the absence of notable absorption above 240 nm up to 1.0 equiv of HgII per L clearly suggest that only two of the Cys-thiolates are bound to HgII (Figure S2). This is further corroborated by a pH-dependent series, obtained at a 1:1 HgII:L ratio, indicating that a deprotonation process takes place with a pKa ∼ 6.43, as fitted to the sigmoidal plots of absorbance traces, driving the {HgS2}-type species into a {HgS3} complex dominating at pH 7.5 (Figure S3). Previous studies with various ligands, offering three Cys thiol groups for HgII coordination, all showed that a transformation from bisthiolate to tristhiolate binding modes occurs in the slightly acidic/neutral pH-regime; however, the observed pKa values span a rather broad range (from ∼4.346 to 8.637). Several parameters, including the position of the donor groups,46,47 coordination geometry around the metal ion center,39 the presence or absence of nearby bulky and/or electron-withdrawing/-donating substituents,48,49 and the net charge at the metal center, influence these pKa values. Some of the studied cyclic peptides seem to provide the highest stabilization of the tristhiolate-bound HgII center, ultimately resulting in low pKa values, while {HgS3} coordination within the interior of three-stranded coiled coils with Cys residues at the so-called “d positions” is the least favored.37,39,55 The {HgS2}-type complexes of short linear peptides transform into {HgS3} species with pKa values of ca. 5.0,46,47 which notably differ from the data observed in the studied system (pKa ∼6.4), suggesting that the positions (2Cys, 3Cys, 9Cys) of the three Cys residues in the metal binding fragment of AfArsR are not ideal for stabilizing the tristhiolate coordination environment over the {HgS2} structure.
The HgII binding affinity of L was also determined by combining data of a peptide displacement reaction with iodide ions (followed by UV spectroscopy, Figure S4) and pH-potentiometric titrations (Figure S5). As expected, the apparent stability obtained for the HgII-L species (logKpH7.5 = 39.0) reflects a huge difference in favor of HgII binding compared to AsIII. These studies are described in the Supporting Information.
Interaction of AsIII and HgII with the AfArsR model peptide at pH 7.5 and 2.0 was also followed by circular dichroism (CD) spectroscopy (Figures 4 and S6). The trends in the change of CD intensities at selected wavelength values, accompanying the titration of L by either AsIII or HgII, perfectly correlate with the observed trends in the UV-monitored titrations. The recorded ellipticities follow a saturation-like tendency and level off slightly above 1.0 equiv of AsIII per ligand at both pH values, and the observed isodichroic points at ∼232 and 257 nm demonstrate a simple equilibrium between the free ligand and a single complex species.
Figure 4.
CD spectra recorded at different c(AsIII):c(L) (A) and c(HgII): c(L) (B) ratios at pH 7.5. Presented data are combined from experiments performed under different conditions: data below 220 nm (left axes) were obtained in a 0.1 mm cylindrical cuvette with c(L) = 650 μM, while data above 220 nm (right axes) were measured in a 1.0 cm conventional cell with c(L) = 65.0 μM. (Full-range spectra measured in the cylindrical cuvette are presented in the Supporting Information, Figure S6.) The thick lines are highlighted spectra for 0.0 (blue), 1.0 (red), and 2.5 or 1.5 equiv (black) of AsIII (A) or HgII (B) per ligand. The insets show the trend in the change of CD intensities at 225 and 290 nm. Panel (C) compares the CD spectra obtained without metal ions and with 1.0 equiv of AsIII and HgII at pH 2.0 (solid lines) and 7.5 (dashed lines) with c(L) = 650 μM, path length = 0.1 mm.
A clear breakpoint near 1.0 equiv of HgII reflects a stronger binding to the ligand, as compared to AsIII, as well as a shift in the speciation between 1:1 and 1.5:1 HgII:L concentration ratios from monomeric toward multinuclear structures. The series of spectra display information about both the changes in the secondary structure of the peptide via AsIII or HgII binding and the coordination mode of the ligand. The free peptide displays an intense negative Cotton effect (∼198 nm) with a lower energy shoulder, attributed to π–π* and n–π* transitions of the amide bond,56,57 reflecting a disordered structure.58−60 This random-coil signature completely diminishes as a consequence of binding of AsIII to L (Figures 4A,C and S6A,B), indicating the organization of the peptide skeleton into a more structured form. Moreover, the same CD signal and thus the same fold of the peptide are observed at both pH 2.0 and 7.5 for the 1:1 AsIII:L complex. The CD signature of the AsIII-bound ligand, i.e., a combined CD minimum–maximum–minimum at 198, 216, and 236 nm, respectively, is very similar to the spectrum of the CdII complex of the (Glu–Cys)4–Gly phytochelatin peptide.61 Based on this CD pattern, the binding of AsIII to L presumably enforces different turns, such as α-helix and 310 helix-like turns or β-turns in the peptide backbone.61−63
The CD spectra recorded with 1.0 equiv of HgII at pH 2.0 and 7.5 (Figure 4C) demonstrate that the peptide adopts a remarkably different structure compared to its AsIII-bound state, indicating that HgII fails to fold the peptide into the structure of the AsIII:L complex. Surprisingly, at pH 7.5, the CD spectrum of the 1:1 HgII:L complex exhibits rather similar features as the (disordered) free peptide, implying that the complex remains, at least partially, disordered. These conclusions are based on the assumption that the CD signal from the peptide backbone chromophore dominates over charge transfer bands involving HgII and AsIII in the short wavelength range of the spectra, vide infra.
The right side of panels A and B in Figure 4 (from the vertical dashed lines) highlights the near-UV region of the recorded spectra by presenting experiments performed in a long path length cell (1.0 cm). The positive CD bands appearing at ≥240–250 nm are related to charge transfer (S– → AsIII or S– → HgII LMCT) transitions being chirally perturbed by the peptide. Charge transfer-related CD bands are relatively rarely reported in metal ion–peptide systems,37,47 probably because of their typically weaker intensity as compared to the peptide backbone-related transitions. Nevertheless, as demonstrated by the present example, such bands can be rather prominent in samples of short-length peptides if they are centered in the wavelength regime that do not overlap with the ligand transitions but may also modulate the observed signals at higher energies, i.e., below ∼210 nm. Furthermore, in contrast to the well-defined signs (positive or negative) of the CD signals related to the different secondary structural elements of the peptide backbone, literature data suggests that the LMCT-induced transitions may result in negative or positive signals, depending on the type of ligand, the number/configuration of chirality centers, and their relative positions to the thiolate–metal ion bonds.34,41,47,49,64 The evolution of the low-energy, CT-related CD bands in the AsIII-L system at any studied pH (Figures 4A and S6A,B) and with HgII at pH 7.5 (Figures 4B and S6D) provides strong support for the formation of the previously proposed tristhiolate binding mode of L. Indeed, the absence of such features in the HgII-L sample at pH 2.0 (Figure S6C) is in accord with the higher energy of the LMCT bands in {HgS2}-type species, shifting also the related CD signals into the wavelength range of the amide-bond transitions.
Considering the short length of the AfArsR model peptide, we cannot exclude that the CD features below 240–250 nm are also affected by LMCT signatures. In general, the relative intensity of the peptide-related transitions and LMCT contributions to the CD of low-molecular-weight metal complexes is an unresolved and open question. With an aim to elucidate the influence of LMCT-related contributions to the observed CD pattern at higher energies, especially, near the negative CD signal (∼199 nm) characteristic for the unbound peptide, we compared the UV-absorption spectra as well as the CD spectra of a number of MII/MIII–peptide systems in the high-energy wavelength regime (MII: HgII, CdII, ZnII, and PbII; MIII: AsIII, SbIII—in the form of its tartrate complex to hinder the hydrolysis of the metalloid65) (Figure 5). All of these (semi)metal ions have well-known strong affinities to multithiol ligands and are expected to be bound efficiently by the studied peptide displaying three cysteine residues.51,65−67 The peptide was expected to fully displace the SbIII-bound tartrate, according to previous studies describing the complete exchange of the SbIII-coordinated tartrate by a presumably less efficient monothiol ligand, glutathione.68 Metal ion titration series demonstrated efficient binding of all of the studied ions to the peptide under the applied conditions. The UV data shown in Figure 5A indicate that the effect of any of the metal ions on the absorbance of the free ligand around 200 nm, especially that of AsIII, is not particularly strong.
Figure 5.
Comparison of UV spectra (A) and CD spectra (B) recorded in different MII-L and MIII-L samples (MII: HgII, CdII, ZnII or PbII; MIII: AsIII or SbIII) at pH 7.5 (phosphate buffer, c = 20 mM). (A) UV spectra were recorded using c(MII):c(L) 1:1 and c(MIII):c(L) 2:1 concentration ratios. Semimetals were used in excess to ensure complete transformation of the ligand to the semimetal-bound species. The inset presents difference spectra obtained by subtracting the spectrum of the unbound ligand, recorded at the same concentration, from the spectra of the metal ion-containing samples (c(L) = 10.0 μM, path length = 1.0 cm). (B) CD spectra were recorded at equimolar ratios of all of the components in all systems. The inset shows the metal ion-induced change of CD intensities at 200 nm, as compared to the intensity of the CD spectrum recorded for the free ligand (c(L) = 650 μM, path length = 0.1 mm).
In contrast, the observed magnitudes of the CD spectra around 200 nm are rather different for the different ions (Figure 5B). All of the divalent ions, even PbII bearing a lone electron pair, induces a significantly smaller change in the intensity of the CD signal centered at ∼199 nm, as compared to the effect of AsIII or SbIII. Although the influence of the coordinating (semi)metal ions on the CD of such a short peptide is clearly rather complex, the observed effects, solely at higher energies, i.e., near 200 nm, may still imply that the two semimetals shape the peptide in a more defined conformation, as compared to the divalent metal ions, whereas the contributions of LMCT to the CD spectra at longer wavelengths (>220 nm) are substantial. The question of peptide backbone-related transitions to the recorded CD spectra was also explored by semiempirical approaches and TD-DFT calculations leading to similar, but not conclusive, results. The applicability of these methods for low-molecular-weight metal ion–peptide complexes needs further investigation; nevertheless, we have added our attempts to the Appendix of the Supporting Information for interested readers.
The distinctly different conformations that the peptide adopts in its AsIII- and HgII-bound states, based on the CD data, may be related to the different coordination geometry preferences of AsIII and HgII. The lone electron pair of AsIII restricts the metalloid into a trigonal pyramidal geometry, whereas HgII is more likely to accommodate a trigonal planar or a distorted trigonal environment. Indeed, it is interesting that the peptide model of the metalloid site of AfArsR, even without the surrounding protein fold, shapes into clearly distinguishable ensembles of conformations upon binding the different ions, and we speculate that this may have a relevance in the protein’s metalloid-selective function.
ESI-MS and 1H NMR Characterization of the Metal(loid) Binding of L
ESI-MS spectra provide further support for the speciation models, proposed by UV and CD spectroscopies and pH potentiometry (Supporting Information) for the AsIII-L and/or HgII-L systems, along with information about possible binding isomers of the {HgS2}-type species present under acidic conditions. These data with representative figures (Figures S7–S9) are presented in the Supporting Information.
Full assignment of the 1H NMR resonances of the free peptide was done, via 1D and 2D NMR experiments (Experimental Procedures in the Supporting Information), at pH 7.5 and 2.0 (Tables S4 and S5 and Figure S10) at 282 K, resulting in significantly better resolved spectra, as compared to those recorded at 298 K, especially at pH 7.5. Spectra obtained for samples with a 1:1 AsIII:L concentration ratio at either pH 7.5 or 2.0 (Figure 6) demonstrate that the presence of AsIII induces significant changes in the chemical shifts and the splitting pattern of the peptide both in the aliphatic (ca. δ < 4.7 ppm) and in the amide region (ca. 6.8 < δ < 9.5 ppm). Signals of nearly all amino acids are notably affected, but the observed new resonances are only slightly broadened and can be assigned to the same dominant species at both pH values (Tables S6 and S7). The chemical shifts of the different protons change very little with pH with the exception of those that are notably affected by the change of the protonation state of nearby noncoordinating residues. This indicates that the binding of all three Cys-thiolates to AsIII affects the structure of the whole ligand. The sharp lines imply that either AsIII binding “locks” the peptide into a well-defined structure or that several structures are present in a fast exchange on the NMR time scale. Our previous 1H NMR results, obtained in samples of iAsIII and bis- or tristhiol ligands, all reflected slow exchange between different AsIII complexes and/or the free ligand when such species were in equilibrium.41,43,44 In contrast, 1H NMR experiments following the addition of the monothiol ligands cysteine and glutathione to AsIII showed a relatively fast (Cys) and intermediate (GSH) exchange rate between the free and bound forms.69 This difference demonstrates that the “origin” of the multiple As–S bonds, whether they form with the same molecule (even if the chelate size is large) or with separate ligands, affects, besides stability, the kinetics of iAsIII complexation.
Figure 6.

Parts of the 1H NMR spectra of L in the absence and presence of 1.0 equiv of AsIII, recorded in H2O/D2O 90/10 V/V % at pH 7.5 (bottom) and 2.0 (top) and assignment of the Cys residue-related resonances (T = 282 K, c(L) = 1.0 mM). Note that the free ligand sample at pH 7.5 contained the reducing agent TCEP (signals of P-CH2 and CH2-COO– appear at ∼2.55 and ∼2.30 ppm, respectively) to maintain the reduced (thiol) form of the peptide (c(TCEP) = 1.0 mM). Representative resonances of minor isomers present in AsIII-L are marked by * symbols (see also Tables S6 and S7). The sharp peak between 3.3 and 3.4 ppm is due to a small amount of methanol.
It is not surprising that the most severely affected resonances are those of the CysβH2 protons that are all shifted downfield by several tens of ppm by AsIII binding. Besides, the NH (amide) signals of the Cys residues also appear at notably higher chemical shifts compared to the unbound peptide. A careful inspection reveals that small additional resonances for almost all groups of protons, besides those attributed to the complexed peptide, also emerge in the spectra of the AsIII-L samples (see Figures 6 and S11). Most of these are also fundamentally shifted, as compared to the signals of the free ligand, implying that the ligand is indeed present in (at least) two complexed isomeric forms, and one of these isomers is significantly more populated than the other(s).
1H–1H ROESY and NOESY spectra demonstrate several medium- and long-range NOE interactions (Figure S12) that could not be observed in the unbound peptide, displaying a random-coil structure. The contacts indicate a turn-like orientation of the central part of the peptide in the presence of AsIII, which is a consequence of the tristhiolate-type coordination of the AsIII center.
The presence of HgII also has a strong impact on the 1H NMR spectra of L (Figure S13); however, the signals are severely broadened, providing very little direct information about the binding modes of the peptide. Both at pH 7.5 and at pH 2.0, all three CysβH2 resonances seem to be shifted downfield (Figure S13). At pH 7.5, it is in line with the proposed tristhiolate binding mode of the peptide. Indeed, at this pH, a small fraction of the ligand is found in a {HgS2}-type species (see Figure S5B), and an exchange between the {HgS2} (isomers) and the {HgS3} complex, at an intermediate rate on the NMR time scale, may partially explain the large line widths.70 The role of chemical exchange is further corroborated by the observed gradual shifting of the CysβH2 signals along with the increasing HgII content from 0 to 0.5 and finally to 1.0 equiv per ligand (Figure S13). Besides, the presence of a small amount of other species, i.e., trinuclear complexes, or internal conformational exchange dynamics within the monocomplex(es) might also contribute to line broadening.
Based on our previous experiences with systems of HgII and short, two-Cys-containing oligopeptides,53,71−74 we assumed that the presence of a uniform {HgS2} structure at pH 2.0 would result in a well-resolved 1H NMR spectrum with clearly assignable resonances. The predominance of one of the possible {HgS2} isomers, especially the one with an unbound 9Cys thiol (suggested under ESI-MS conditions; see the Supporting Information), might leave one of the CysβH2 signals unaffected or only weakly affected. As it turned out, the spectrum recorded at pH 2.0 displays broad signals for all amide protons, the aliphatic protons of Cys, and several other residues (Figure S13). This implies that the {HgS2}-type complex presumably exists in the three possible isomeric forms and exchanges the thiolates with the rate falling into an intermediate regime on the NMR time scale.70 Furthermore, the presence of a small fraction of trinuclear species at a 1:1 c(HgII):c(L) ratio cannot be excluded at the concentration of the NMR experiments.
These data, again, suggest fundamental differences in the AsIII- and HgII-binding characteristics of the model peptide. While AsIII is accommodated in a well-defined structure, there is a larger flexibility in HgII coordination, as also indicated by the remarkably different CD features of the AsIII- and HgII-bound species.
Characterization of the HgII Coordination Geometry by 199mHg PAC Spectroscopy
199mHg PAC experiments were conducted in order to directly characterize the metal site structure of the HgII-L complex(es).
The samples contained HgII and the ligand in a 1:1 concentration ratio at different pH values. The PAC data are presented in Table S8 and Figure 7. The data are consistent with essentially only one metal site structure at the starting and end points of a pH titration, i.e., at pH 4.4 and 8.0, while two species are reflected by the data at pH 6.2 and ∼7.0. The pKa is estimated to be in the range from 6.0 to 7.0, in accordance with the UV and CD data, vide supra. The NQI recorded at low pH (NQI1) agrees excellently with the {HgS2} coordination mode.55,73,75,76 The high pH form (NQI2) bears some resemblance to previous 199mHg PAC data for trigonal planar structures, for which the quadrupole frequency, νQ, falls in the range of 1.16–1.18 GHz and the asymmetry parameter, η, falls in the range of 0.22–0.31,55 although the frequency is lower (1.02 GHz) and the NQI is considerably more asymmetric (η = 0.68). A perfect trigonal pyramidal HgS3 structure does not agree with the high asymmetry parameter either: in the ideal case, that would also (like the trigonal planar structure) give η = 0. According to the simple AOM (angular overlap model),77 a T-shaped {HgS3} structure with equal Hg–S bond lengths gives the same NQI frequency, νQ, as trigonal planar species, however, with an asymmetry parameter η = 1.0. Under subequimolar HgII:protein concentration ratios, the HgII-bound CueR metalloregulator also displayed a pair of parameters (νQ = 1.04 or 1.13 GHz, η = 0.77 or 0.56 depending on the absence/presence of DNA)75 that indicated an intermediate state between a perfect trigonal planar and T-shaped structures. Consequently, in the present system, the most likely structural interpretation of the high pH signal (NQI2) is a distorted {HgS3} binding mode. The notion that the structure is distorted agrees well with the coordination by two adjacent cysteinates in the amino acid sequence, imposing geometrical constraints on the positions of the two sulfur atoms in the first coordination sphere. This is substantiated by the optimized structures (Figure S18G), where the S–Hg–S angle for 2Cys and 3Cys is larger than the other S–Hg–S angles.
Figure 7.
Fourier-transformed 199mHg PAC data (thin lines) and fits (bold-faced lines) obtained for the c(HgII): c(L) = 1:1 system at different pH values (at 1 °C, c(L) = 34 μM). The sets of yellow and blue vertical dotted lines indicate the positions of signals corresponding to NQI1 and NQI2, respectively.
Characterization of the AsIII and HgII Coordination Centers by EXAFS Spectroscopy
EXAFS spectra were recorded for the AsIII-L (As K-edge) complex and the HgII-L (Hg LIII-edge) complex, as well as for two reference samples, using glutathione (GSH) as the ligand; see Figures 8 and S15.
Figure 8.
EXAFS spectra and corresponding Fourier transforms of the AsIII-L (top) and HgII-L (bottom) complexes with the simulated models of a single scattering path (As–S3 or Hg–S3) shown in red as solid lines. Experimental conditions: T = 15 K; 25% (V/V) glycerol; c(AsIII): c(L) = 1:1.2, c(AsIII) = 2.8 mM, pH 7.5; c(HgII): c(L) = 1:1.2, c(HgII) = 2.9 mM, pH 8.0.
The spectra recorded for the AsIII-L complex and the AsIII-GSH reference are highly similar. Fitting the data using a model with only sulfur in the first coordination sphere gives a coordination number of 2.9(1) for both data sets and As–S bond lengths of 2.262(6) Å and 2.257(7) Å for the AsIII-L complex and for the AsIII-GSH reference, respectively; see Table S9. The structures optimized using DFT methods, vide infra, indicate that there may be minor differences in the three As–S bond lengths, presumably caused by the constraints imposed by the peptide backbone, and this may be reflected in the line width of the Fourier-transformed EXAFS data for the AsIII–peptide complex, which is slightly larger than that for the reference.
Fitting the data recorded for AsIII-L with two independent As–S bond lengths gives one As–S bond length of 2.24(5) Å and two of 2.27(4) Å. Most importantly, the bond lengths and coordination number are consistent with the EXAFS data and fits reported by Rosen and co-workers for the mother protein AfArsR (As–S bond length of 2.24 Å)23 as well as for another AsIII sensor protein CgArsR (As–S bond length of 2.25 Å),24 indicating that the peptide adequately models the AsIII binding site of the protein. There is a weak signal in the AsIII-L spectrum around 4 Å, which is not accounted for by the fit. This might originate from multiple scattering or atoms at a longer distance from AsIII than the three sulfurs. This additional signal is not present for the AsIII-GSH reference, possibly because the reference is less ordered. This additional weak signal at large values of R cannot unambiguously be discerned for the HgII-L complex because the signal-to-noise ratio is inadequate. This might indicate that the HgII-L complex is more dynamic and disordered than AsIII-L, in agreement with the MD simulations, vide infra, and CD spectroscopy.
Fitting the EXAFS data recorded for the HgII-L complex with only sulfur in the first coordination sphere gives a coordination number of 2.5(3) and a Hg–S bond length of 2.46(1) Å; see Figure 8 and Table S9. The bond length falls in the range expected for HgS3 coordination as determined by X-ray diffraction (2.40–2.51 Å, with an average of 2.44 Å),78 as well as by EXAFS data on HgII-GSH complexes.79,80 The too-low coordination number (<3) may arise due to significant distortion from the trigonal planar geometry of the HgII binding site, possibly accompanied by dynamics, as indicated by the relatively large Debye–Waller factor, and in agreement with the molecular dynamics simulations, vide infra. Alternatively, the coordination number of 2.5(3) reflects a distribution of HgS2 and HgS3 species, and fitting a linear combination of two such sites to the EXAFS data leads to 51(9)% HgS3 and 49(7)% HgS2, with bond lengths of 2.50 and 2.36 Å, respectively (Table S9). Allowing the three-coordinated site to have two different Hg–S bond lengths leads to 69(9)% HgS3 and 31(9)% HgS2. In either case, the HgS3-to-HgS2 ratio is higher than expected given the UV, CD, and PAC spectroscopic analyses in this work, indicating that the EXAFS conditions (25% V/V glycerol and low temperature) induce this shift in the equilibrium between HgS3 and HgS2.
In summary, the EXAFS data for AsIII-L indicate the presence of an AsS3 complex, in agreement with the interpretation of EXAFS data recorded for the AfArsR23 and CgArsR24 proteins, while the HgII-L complex displays more structural diversity, either as a highly distorted trigonal planar site or due to the coexistence of HgS2 and HgS3 species.
DFT-Optimized Structure of the AsIII and HgII Binding Segment of the Coordinated Ligand
The metal site model system excised from the AsIII-bound AfArsR structure, shown in Figures 9A and S17–S19 in a ball and stick representation, was geometry-optimized by DFT calculations. The model with HgII inserted instead of AsIII gave the planar HgII coordination geometry in Figure 9B. In contrast, optimization with AsIII converged to the structure in Figure 9C, which strongly resembles the crystal structure endo-form (Figure 9A).
Figure 9.
Reference crystal structure (A) and DFT-optimized structures of the 2Cys-3Cys fragment of L, complemented with a bound CH3–CH2–S– ion (mimicking the distant 9Cys unit) in complexes with HgII (B) and AsIII in endo-conformation (C) and exo-conformation (D). Models were optimized by ORCA 5.0.381 (see the Experimental Procedures in the Supporting Information for details). The crystal structure (pdb: 6J05(25)) cutout has been saturated with H atoms. S–metal ion–S angles for the models are shown in Table S10.
The structure in Figure 9D is the corresponding exo-form, resulting from geometry optimization of the model with AsIII initially placed on the other side of the plane spanned by the three S atoms. The calculated difference in Gibbs free energy between the optimized exo and endo models is 2.6 kcal/mol, corresponding, respectively, to ∼1.3 and ∼98.7% population of the two structures at 25 °C. This suggests that a low-lying conformer may coexist with the major endo species, providing a possible explanation for the appearance of the additional set of 1H NMR resonances besides that of the main complex conformer. Moreover, the DFT optimizations show a metal-ion-specific orientation of the amide plane connecting 2Cys and 3Cys. The amide plane is parallel to the (near) C3 symmetry axis of the AsS3 structure in both the endo- and exo-models but is tilted in the HgII model, possibly caused by charge–dipole interactions of the negative charge at the HgS3 center and the amide dipole. Consequently, in AsIII models, the distance between the amide hydrogen and the metal ion is 3.46 Å (endo) and 3.39 Å (exo), while it is significantly shorter, 2.86 Å, in the HgII model, reflecting a fundamental difference of the peptide structure in the close surroundings of the AsIII and HgII centers.
There is only one other ArsR protein for which the structure of the AsIII-bound form has been characterized by X-ray diffraction, CgArsR.25 It is interesting how well the crystal structure of the CgArsR protein, offering two of the AsIII binding Cys thiols from one monomer and a third Cys thiol from the end of the α2 helix of the other monomer, overlaps with the structure of AfArsR-As, which displays the three AsIII-bound Cys residues in a short sequence of the same protein chain.25 In the CgArsR structure, helix α2, connected directly to AsIII, can be regarded as a link from the AsIII binding site toward the DNA-binding site.
A more comprehensive account of the signaling pathway, operating in ArsR family member proteins, may be found in recent literature.82−85 This provides a pathway for signal transmission that is a function of how AsIII binding may affect the DNA-binding site. Indeed, a plausible link between the metal site and helix α2 is also present in AfArsR-As via a hydrogen bond connecting the amide NH between 96Cys and 97His and the last carbonyl oxygen of helix α2 (2.7 Å). HgII binding to the metal site and the changes induced in the surrounding protein fold may disrupt this connection to the DNA-binding domain. One may speculate that the total charge at the AsIII binding site (=0) and the HgII-binding site (=–1) also affects metal ion selection. The α2 helix in AfArsR (of the other monomer) points the negative end of the helix dipole toward the metal center in the crystal structure,25 which may affect the structure of the C-terminal peptide fragment in the HgII-loaded form.
One of the advocated structural origins of the metal ion-selective functioning of metalloregulatory proteins is related to the coordination geometry of the bound metal ion,86 which is well-corroborated by our data. It seems likely that the selectivity based on proper folding of part of the protein, induced only by the cognate metal(loid), is operational in other sensor proteins, as well. Indeed, careful comparison of crystal structures of the metal-free and metal-bound forms of various metalloregulators (relying partly on modeled structures using AlphaFold87,88 (Uniprot id: B7J952 (AfArsR), P0ACS5 (ZntR)) and I-TASSER89−91 (I-TASSER server)) indicates a certain level of flexibility in the metal-free or repressor forms near the effector binding site (Figure S20). Similarly, structural data on the DtxR/MntR family member proteins MntR92,93 and ScaR,94 the MerR family sensors CueR,95,96 ZntR,95 and MeR,97 the tetrameric NikR,98 as well as the ArsR/SmtB family sensory proteins CzrA,99 CadC,100,101 and AfArsR25 all show that the coordinating ligands are only partially arranged in the absence of metal ions. However, metal ion binding reorganizes the coordinating donor groups as well as the surrounding protein skeleton in a way that allows the metal ion to adopt its favored coordination geometry. These sometimes subtle structural changes in the protein scaffold around the metal center may serve as the initiation of the signal transmission pathway toward the DNA-binding domain.
MD Simulations of Ac-NCCHGTRDCA-NH2
The Ac-NCCHGTRDCA-NH2 peptide with coordinated AsIII populated two main conformational ensembles in the 2000 ns MD trajectory, as indicated by the backbone RMSD plot (Figure 10A, left). Inset structural snapshots at 250 and 1500 ns highlight the characteristics of these ensembles. The backbone atoms of the snapshots have been aligned to allow their comparison. As expected, the first coordination sphere geometry remains largely unchanged upon comparison of the two snapshots. However, the backbone is more expanded at 1500 ns, mainly as a consequence of the disruption of the backbone hydrogen bond between 2Cys and 7Arg. Indeed, such a hydrogen bond in the AsIII-bound ligand is also suggested by 1H NMR spectra recorded at varying temperatures that reflect a −0.2 ppb/K temperature coefficient for the 7Arg-NH signal (see Figure S14 with text).102,103
Figure 10.

(A) Backbone root-mean-square distance (RMSD) from the initial structure during MD simulations of the Ac-NCCHGTRDCA-NH2 peptide with AsIII (left), with HgII (middle), and without metal ions (right). Insets show structural snapshots for the AsIII and HgII systems at 250 and 1500 ns. Cα atoms are indicated by spheres colored by the sequence position (red, N-terminal; blue, C-terminal). (B) Superposition of MD snapshots within each trajectory to highlight conformational variation.
In the HgII system, the first coordination sphere also remains well-conserved. However, the RMSD curve in Figure 10A (middle) reveals increased backbone dynamics relative to those of the system with AsIII. Snapshots at 250 and 1500 ns are displayed for consistency with the AsIII system, although the RMSD fluctuations indicate more than two conformational ensembles. The metal-ion-free system displays pronounced backbone RMSD fluctuations (Figure 10A, right) due to the vastly increased conformational freedom in the absence of metal ion coordination. The RMSD curves in Figure 10A are further detailed with structural snapshots. The initial and final structures for each simulation are shown in Figure S21. Figure 10B depicts the superposition of frames for each MD trajectory, reduced to 20 frames. This superposition highlights the increase in conformational dynamics, advancing across the subfigures from the left (AsIII) through the middle (HgII) to the right (no metal ion). Thus, the MD simulations demonstrate that the apparent minor local difference in basic inorganic chemical coordination preference for AsIII (trigonal pyramidal) to HgII (trigonal planar) has a profound effect on the structure and dynamics of the entire peptide, with AsIII binding leading to formation of a well-defined structure and presumably the functional form of the AfArsR protein.
Conclusions
Studies on the metalloid binding fragment of the transcriptional regulator AfArsR revealed that the possible origins of the effector recognition selectivity of the protein are manifested already at the peptide model level. The AsIII and HgII binding characteristics of the peptide were compared applying several equilibrium, solution structural, and computational methods at different AsIII/HgII–ligand concentration ratios and pHs. HgII was chosen as a highly thiophilic metal ion, which may also accommodate a tristhiolate-coordinated environment. The HgII binding affinity of the AfArsR model peptide was demonstrated to surpass the affinity for AsIII by an enormous margin, which highlights the significance of the metalloid-selective operation of the ArsR proteins. Titrations of the peptide by AsIII and HgII, followed by UV and CD techniques, show that tristhiolate-coordinated structures dominate for both ions at pH 7.5; however, the peptide is driven into a fundamentally different conformation in the AsIII-bound species as compared to either the free ligand or the HgII complex. 1H NMR data suggest that the structure of the peptide is significantly more defined in the AsIII complex than in the presence of HgII. In agreement with this, conformational disorder in MD simulations was lower for AsIII-peptide models relative to HgII- and metal ion-free models. The DFT-optimized structure of the AsIII-bound ligand, using a small model, reflects that the structure of the AfArsR-As, displaying an endo-conformation, is very closely retained, but the calculations also indicate the possibility for the existence of the exo-type species, in line with the NMR data. This minor isomer represents only a very small fraction of the ligand for the peptide model system, and it is presumably not relevant with regard to the native metalloid site of the protein. Optimization of the minimal model counterpart of the small endo model, i.e., As(CH3S)3 (Figures S18 and S19), leads to a closely overlapping structure to that of the small model. This implies that the amino acid sequence with the adjacent 2Cys and 3Cys residues allows for the organization of part of the metalloid binding site where the optimal trigonal pyramidal structure is not significantly distorted by the constraints imposed by the protein backbone. In contrast, DFT calculations showed that HgII can accommodate a distorted trigonal planar geometry in the energetically favored structure of its peptide complex, in good correlation with 199mHg PAC results, which ultimately means that the metal center is significantly displaced from the position where the metalloid is located in the crystal structure. It is remarkable that the amide group of the Cys–Cys fragment, being perpendicular to the AsS3 plane, is tilted in the HgII-bound peptide possibly because of the electrostatic interaction between the HgS3 center and the amide dipole.
In summary, our spectroscopic and computational results indicate that the local peptide fold, shaping an optimal trigonal pyramidal coordination geometry via the binding of the three Cys thiolate groups, plays an important role in the selection of the cognate effectors. The fundamentally different binding preferences of AsIII and HgII to the effector binding site of AfArsR might lead to a disparate protein structure in the AsIII- vs the HgII-bound forms, at least in the surroundings of the metal center, with a potential impact on the protein function by controlling metal ion recognition. In addition, all of the data in this work imply that for target ions such as AsIII, where it is difficult to establish selectivity over other biologically relevant metal ions by design of a high-affinity binding site, an alternative route to selectivity is found, where only the target ion triggers a conformational change—for AfArsR, it is a change from disorder of the free peptide to order of the AsIII–peptide complex—which leads to the active form of the protein. In a broader perspective, the metal ion-induced organization of the structurally flexible metal binding site, observed in the peptide model of AfArsR, seems to operate also in a number of metal sensor proteins, and this may be a key step in metal ion selection. The flexibility of the metal binding site allows the cognate metal ion to reorganize the local protein fold to achieve the preferred coordination geometry, leading to signal transmission from the metal site to the DNA-binding domain, a pathway that is disrupted by the different coordination structure of the noncognate ions.
Acknowledgments
The authors acknowledge the financial support received from the German Ministry of Education and Research (BMBF) under Grants 05K16PGA, 05K22PGA, and 05K22PGB, the Portuguese Foundation for Science and Technology (FCT, Projects CERN/FIS-TEC/0003/2019 and CERN/FIS-TEC/0003/2021), and from the European Union’s Horizon 2020 Framework Research and Innovation Program under Grant Agreement No. 654002 (ENSAR2). The authors thank CERN and the ISOLDE technical team for beam time as well as EURONS and NICE for financial support. The authors acknowledge MAX IV Laboratory for time on Beamline Balder under Proposal 20240082. Research conducted at MAX IV, a Swedish national user facility, is supported by the Swedish Research council under Contract 2018-07152, the Swedish Governmental Agency for Innovation Systems under Contract 2018-04969, and Formas under Contract 2019-02496. N.J.C. is grateful for the research grant 40932 awarded by VILLUM FONDEN, and Z.H.N. acknowledges support from the College of Pharmacy, University of Babylon, Hillah 51001, Iraq. The work received support from the University of Szeged Open Access Fund, grant number: 7056.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.3c11665.
Detailed description of the applied materials, sample preparation protocols, and experimental methods; additional information and figures of AsIII and HgII titrations of the ligand followed by UV and CD spectroscopies; determination of the AsIII-binding affinities; evaluation of pH-potentiometric data and determination of HgII-binding affinities; characterization of AsIII and HgII binding by ESI-MS; additional NMR data and figures; table of fitted 199mHg PAC parameters; table of fitted EXAFS parameters; figure of the spectra recorded for reference samples; figure of simulated species distribution for HgII-GSH; additional data and figures of DFT and MD studies; and appendix describing semiempirical (SESCA and PDBMD2CD) approaches to predict CD spectra, as well as TD-DFT calculations of the excited states of the AsIII-bound peptide complex and the related electronic CD (PDF)
Author Contributions
The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.
The authors declare no competing financial interest.
Supplementary Material
References
- Hughes M. F. Arsenic toxicity and potential mechanisms of action. Toxicol. Lett. 2002, 133, 1. 10.1016/S0378-4274(02)00084-X. [DOI] [PubMed] [Google Scholar]
 - Kitchin K. T.; Wallace K.; Andrewes P.. Some Chemical Properties Underlying Arsenic’s Biological Activity. In Arsenic Exposure and Health Effects; Chappell W. R.; Abernathy C. O.; Calderon R. L.; Thomas D. J., Eds.; Elsevier: San Diego, 2003; pp 345–354. [Google Scholar]
 - Burford N.; Carpenter Y.-Y.; Conrad E.; Saunders C. D. L.. The Chemistry of Arsenic, Antimony and Bismuth. In Biological Chemistry of Arsenic, Antimony and Bismuth; Sun H., Ed.; John Wiley & Sons Ltd.: Chichester, United Kingdom, 2011; pp 1–15. [Google Scholar]
 - Jomova K.; Jenisova Z.; Feszterova M.; Baros S.; Liska J.; Hudecova D.; Rhodesd C. J.; Valko M. Arsenic: toxicity, oxidative stress and human disease. J. Appl. Toxicol. 2011, 31, 95–107. 10.1002/jat.1649. [DOI] [PubMed] [Google Scholar]
 - Hughes M. F.; Beck B. D.; Chen Y.; Lewis A. S.; Thomas D. J. Arsenic Exposure and Toxicology: A Historical Perspective. Toxicol. Sci. 2011, 123, 305. 10.1093/toxsci/kfr184. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Hu Y.; Li J.; Lou B.; Wu R.; Wang G.; Lu C.; Wang H.; Pi J.; Xu Y. The Role of Reactive Oxygen Species in Arsenic Toxicity. Biomolecules 2020, 10, 240. 10.3390/biom10020240. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Shen S.; Li X.-F.; Cullen W. R.; Weinfeld M.; Le X. C. Arsenic Binding to Proteins. Chem. Rev. 2013, 113, 7769–7792. 10.1021/cr300015c. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Dixon H. B. F. The Biochemical Action of Arsonic Acids Especially As Phosphate Analogues. Adv. Inorg. Chem. 1996, 44, 191–227. 10.1016/S0898-8838(08)60131-2. [DOI] [Google Scholar]
 - Vergara-Gerónimo C. A.; Del Río A. L.; Rodríguez-Dorantes M.; Ostrosky-Wegman P.; Salazar A. M. Arsenic-protein interactions as a mechanism of arsenic toxicity. Toxicol. Appl. Pharmacol. 2021, 431, 115738 10.1016/j.taap.2021.115738. [DOI] [PubMed] [Google Scholar]
 - Bergquist E. R.; Fischer R. J.; Sugden K. D.; Martin B. D. Inhibition by methylated organoarsenicals of the respiratory 2-oxo-acid dehydrogenases. J. Organomet. Chem. 2009, 694, 973–980. 10.1016/j.jorganchem.2008.12.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Schwerdtle T.; Walter I.; Hartwig A. Arsenite and its biomethylated metabolites interfere with the formation and repair of stable BPDE-induced DNA adducts in human cells and impair XPAzf and Fpg. DNA Repair 2003, 2, 1449–1463. 10.1016/j.dnarep.2003.09.001. [DOI] [PubMed] [Google Scholar]
 - Zhou X.; Sun X.; Cooper K. L.; Wang F.; Liu K. J.; Hudson L. G. Arsenite Interacts Selectively with Zinc Finger Proteins Containing C3H1 or C4 Motifs. J. Biol. Chem. 2011, 286, 22855–22863. 10.1074/jbc.M111.232926. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Wang Z.; Zhang H.; Li X. F.; Le X. C. Study of interactions between arsenicals and thioredoxins (human and E. coli) using mass spectrometry. Rapid Commun. Mass Spectrom. 2007, 21, 3658–3666. 10.1002/rcm.3263. [DOI] [PubMed] [Google Scholar]
 - Kitchin K. T.; Wallace K. Arsenite binding to synthetic peptides based on the Zn finger region and the estrogen binding region of the human estrogen receptor-α. Toxicol. Appl. Pharmacol. 2005, 206, 66–72. 10.1016/j.taap.2004.12.010. [DOI] [PubMed] [Google Scholar]
 - Kitchin K. T.; Wallace K. Arsenite Binding to Synthetic Peptides: The Effect of Increasing Length between Two Cysteines. J. Biochem. Mol. Toxicol. 2006, 20, 35–38. 10.1002/jbt.20112. [DOI] [PubMed] [Google Scholar]
 - Kitchin K. T.Arsenic’s Interactions with Macromolecules and its Relationship to Carcinogenesis. In Biological Chemistry of Arsenic, Antimony and Bismuth; Sun H., Ed.; John Wiley & Sons Ltd.: Chichester, United Kingdom, 2011; pp 19–51. [Google Scholar]
 - Zhang X.-W.; Yan X.-J.; Zhou Z.-R.; Yang F.-F.; Wu Z.-Y.; Sun H.-B.; Liang W.-X.; Song A.-X.; Lallemand-Breitenbach V.; Jeanne M.; Zhang Q.-Y.; Yang H.-Y.; Huang Q.-H.; Zhou G.-B.; Tong J.-H.; Zhang Y.; Wu J.-H.; Hu H.-Y.; de Thé H.; Chen S.-J.; Chen Z. Arsenic Trioxide Controls the Fate of the PML-RARa Oncoprotein by Directly Binding PML. Science 2010, 328, 240–243. 10.1126/science.1183424. [DOI] [PubMed] [Google Scholar]
 - Yuan A. T.; Stillman M. J. Arsenic binding to human metallothionein-3. Chem. Sci. 2023, 14, 5756. 10.1039/D3SC00400G. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Yang J.; Rawat S.; Stemmler T. L.; Rosen B. P. Arsenic Binding and Transfer by the ArsD As(III) Metallochaperone. Biochemistry 2010, 49, 3658–3666. 10.1021/bi100026a. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Ruan X.; Bhattacharjee H.; Rosen B. P. Characterization of the metalloactivation domain of an arsenite/antimonite resistance pump. Mol. Microbiol. 2008, 67, 392–402. 10.1111/j.1365-2958.2007.06049.x. [DOI] [PubMed] [Google Scholar]
 - Packianathan C.; Kandavelu P.; Rosen B. P. The Structure of an As(III) S-Adenosylmethionine Methyltransferase with 3-Coordinately Bound As(III) Depicts the First Step in Catalysis. Biochemistry 2018, 57, 4083–4092. 10.1021/acs.biochem.8b00457. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Shi W.; Dong J.; Scott R. A.; Ksenzenko M. Y.; Rosen B. P. The Role of Arsenic-Thiol Interactions in Metalloregulation of the ars Operon. J. Biol. Chem. 1996, 271, 9291–9297. 10.1074/jbc.271.16.9291. [DOI] [PubMed] [Google Scholar]
 - Qin J.; Fu H.-L.; Ye J.; Bencze K. Z.; Stemmler T. L.; Rawlings D. E.; Rosen B. P. Convergent Evolution of a New Arsenic Binding Site in the ArsR/SmtB Family of Metalloregulators. J. Biol. Chem. 2007, 282, 34346–34355. 10.1074/jbc.M706565200. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Ordóñez E.; Thiyagarajan S.; Cook J. D.; Stemmler T. L.; Gil J. A.; Mateos L. M.; Rosen B. P. Evolution of Metal(loid) Binding Sites in Transcriptional Regulators. J. Biol. Chem. 2008, 283, 25706–25714. 10.1074/jbc.M803209200. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Prabaharan C.; Kandavelu P.; Packianathan C.; Rosen B. P.; Thiyagarajan S. Structures of two ArsR As(III)-responsive transcriptional repressors: Implications for the mechanism of derepression. J. Struct. Biol. 2019, 207, 209–217. 10.1016/j.jsb.2019.05.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Zhang J.; Li Y.-N.; Shen J.; Nadar V. S.; Chen J. Characterization of a novel ArsR regulates divergent ars operon in Ensifer adhaerens strain ST2. FEMS Microbiol. Lett. 2023, 370, fnad113 10.1093/femsle/fnad113. [DOI] [PubMed] [Google Scholar]
 - Busenlehner L. S.; Pennella M. A.; Giedroc D. P. The SmtB/ArsR family of metalloregulatory transcriptional repressors: structural insights into prokaryotic metal resistance. FEMS Microbiol. Rev. 2003, 27, 131–143. 10.1016/S0168-6445(03)00054-8. [DOI] [PubMed] [Google Scholar]
 - Xu C.; Zhou T.; Kuroda M.; Rosen B. P. Metalloid Resistance Mechanisms in Prokaryotes. J. Biochem. 1998, 123, 16–23. 10.1093/oxfordjournals.jbchem.a021904. [DOI] [PubMed] [Google Scholar]
 - Rosen B. P. Biochemistry of arsenic detoxi¢cation. FEBS Lett. 2002, 529, 86–92. 10.1016/S0014-5793(02)03186-1. [DOI] [PubMed] [Google Scholar]
 - Arruda L. M.; Monteiro L. M. O.; Silva-Rocha R. The Chromobacterium violaceum ArsR Arsenite Repressor Exerts Tighter Control on Its Cognate Promoter Than the Escherichia coli System. Front. Microbiol. 2016, 7, 1851. 10.3389/fmicb.2016.01851. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Wu J.; Rosen B. P. Metalloregulated Expression of the ars Operon. J. Biol. Chem. 1993, 268, 52–58. 10.1016/S0021-9258(18)54113-2. [DOI] [PubMed] [Google Scholar]
 - Tauriainen S.; Karp M.; Chang W.; Virta M. Recombinant Luminescent Bacteria for Measuring Bioavailable Arsenite and Antimonite. Appl. Environ. Microbiol. 1997, 63, 4456–4461. 10.1128/aem.63.11.4456-4461.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Liao V. H.; Ou K.-L. Development and testing of a green fluorescent protein-based bacterial biosensor for measuring bioavailable arsenic in contaminated groundwater samples. Environ. Toxicol. Chem. 2005, 24, 1624–1631. 10.1897/04-500R.1. [DOI] [PubMed] [Google Scholar]
 - Lee W.; Kim H.; Jang G.; Kim B.-G.; Yoon Y. Antimony sensing whole-cell bioreporters derived from ArsR genetic engineering. Appl. Microbiol. Biotechnol. 2020, 104, 2691–2699. 10.1007/s00253-020-10413-5. [DOI] [PubMed] [Google Scholar]
 - Chen J.; Zhu Y.-G.; Rosen B. P. A Novel Biosensor Selective for Organoarsenicals. Appl. Environ. Microbiol. 2012, 78, 7145–7147. 10.1128/AEM.01721-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Chen J.; Rosen B. P.. Metalloid-Sensing Transcriptional Regulatory Proteins. In Molecular Bio-Sensors and the Role of Metal Ions, Metal Ions in Life Sciences; Sigel A.; Sigel H.; Freisinger E.; Sigel R., Eds.; CRC Press: Boca Raton, 2022; Vol. 23, pp 1–21. [Google Scholar]
 - Matzapetakis M.; Farrer B. T.; Weng T.-C.; Hemmingsen L.; Penner-Hahn J. E.; Pecoraro V. L. Comparison of the Binding of Cadmium(II), Mercury(II), and Arsenic(III) to the de Novo Designed Peptides TRI L12C and TRI L16C. J. Am. Chem. Soc. 2002, 124, 8042–8054. 10.1021/ja017520u. [DOI] [PubMed] [Google Scholar]
 - Touw D. S.; Nordman C. E.; Stuckey J. A.; Pecoraro V. L. Identifying important structural characteristics of arsenic resistance proteins by using designed three-stranded coiled coils. Proc. Natl. Acad. Sci. U.S.A. 2007, 104, 11969–11974. 10.1073/pnas.0701979104. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Chakraborty S.; Touw D. S.; Peacock A. F. A.; Stuckey J.; Pecoraro V. L. Structural Comparisons of Apo- and Metalated Three-Stranded Coiled Coils Clarify Metal Binding Determinants in Thiolate Containing Designed Peptides. J. Am. Chem. Soc. 2010, 132, 13240–13250. 10.1021/ja101812c. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Zampella G.; Neupane K. P.; De Gioia L.; Pecoraro V. L. The Importance of Stereochemically Active Lone Pairs For Influencing PbII and AsIII Protein Binding. Chem. – Eur. J. 2012, 18, 2040–2050. 10.1002/chem.201102786. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Szekeres L. I.; Maldivi P.; Lebrun C.; Gateau C.; Mesterhazy E.; Delangle P.; Jancso A. Tristhiolato Pseudopeptides Bind Arsenic(III) in an AsS3 Coordination Environment Imitating Metalloid Binding Sites in Proteins. Inorg. Chem. 2023, 62, 6817–6824. 10.1021/acs.inorgchem.3c00563. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Spuches A. M.; Kruszyna H. G.; Rich A. M.; Wilcox D. E. Thermodynamics of the As(III)-Thiol Interaction: Arsenite and Monomethylarsenite Complexes with Glutathione, Dihydrolipoic Acid, and Other Thiol Ligands. Inorg. Chem. 2005, 44, 2964–2972. 10.1021/ic048694q. [DOI] [PubMed] [Google Scholar]
 - Kolozsi A.; Lakatos A.; Galbacs G.; Madsen A. O.; Larsen E.; Gyurcsik B. A pH-Metric, UV, NMR, and X-ray Crystallographic Study on Arsenous Acid Reacting with Dithioerythritol. Inorg. Chem. 2008, 47, 3832–3840. 10.1021/ic7024439. [DOI] [PubMed] [Google Scholar]
 - Szekeres L. I.; Gyurcsik B.; Kiss T.; Kele Z.; Jancso A. Interaction of Arsenous Acid with the Dithiol-Type Chelator British Anti-Lewisite (BAL): Structure and Stability of Species Formed in an Unexpectedly Complex System. Inorg. Chem. 2018, 57, 7191–7200. 10.1021/acs.inorgchem.8b00894. [DOI] [PubMed] [Google Scholar]
 - Zékány L.; Nagypál I.; Peintler G.. PSEQUAD for Chemical Equilibria; Technical Software Distributors: Baltimore, MD, 1991.
 - Sénèque O.; Rousselot-Pailley P.; Pujol A.; Boturyn D.; Crouzy S.; Proux O.; Manceau A.; Lebrun C.; Delangle P. Mercury Trithiolate Binding (HgS3) to a de Novo Designed Cyclic Decapeptide with Three Preoriented Cysteine Side Chains. Inorg. Chem. 2018, 57, 2705–2713. 10.1021/acs.inorgchem.7b03103. [DOI] [PubMed] [Google Scholar]
 - Mesterházy E.; Lebrun C.; Crouzy S.; Jancsó A.; Delangle P. Short oligopeptides with three cysteine residues as models of sulphur-rich Cu(I)- and Hg(II)-binding sites in proteins. Metallomics 2018, 10, 1232–1244. 10.1039/C8MT00113H. [DOI] [PubMed] [Google Scholar]
 - Pujol A. M.; Lebrun C.; Gateau C.; Manceau A.; Delangle P. Mercury-Sequestering Pseudopeptides with a Tris(cysteine) Environment in Water. Eur. J. Inorg. Chem. 2012, 2012, 3835–3843. 10.1002/ejic.201200484. [DOI] [Google Scholar]
 - Jullien A. S.; Gateau C.; Lebrun C.; Delangle P. Mercury Complexes with Tripodal Pseudopeptides Derived from D-Penicillamine Favour a HgS3 Coordination. Eur. J. Inorg. Chem. 2015, 2015, 3674–3680. 10.1002/ejic.201500421. [DOI] [Google Scholar]
 - Dieckmann G. R.; McRorie D. K.; Tierney D. L.; Utschig L. M.; Singer C. P.; O’Halloran T. V.; Penner-Hahn J. E.; DeGrado W. F.; Pecoraro V. L. De No Design of Mercury-Binding Two- and Three-Helical Bundles. J. Am. Chem. Soc. 1997, 119, 6195–6196. 10.1021/ja964351i. [DOI] [Google Scholar]
 - Rousselot-Pailley P.; Sénèque O.; Lebrun C.; Crouzy S.; Boturyn D.; Dumy P.; Ferrand M.; Delangle P. Model Peptides Based on the Binding Loop of the Copper Metallochaperone Atx1: Selectivity of the Consensus Sequence MxCxxC for Metal Ions Hg(II), Cu(I), Cd(II), Pb(II), and Zn(II). Inorg. Chem. 2006, 45, 5510–5520. 10.1021/ic060430b. [DOI] [PubMed] [Google Scholar]
 - Pires S.; Habjanič J.; Sezer M.; Soares C. M.; Hemmingsen L.; Iranzo O. Design of a Peptidic Turn with High Affinity for HgII. Inorg. Chem. 2012, 51, 11339–11348. 10.1021/ic3008014. [DOI] [PubMed] [Google Scholar]
 - Szekeres L. I.; Bálint S.; Galbács G.; Kálomista I.; Kiss T.; Larsen F. H.; Hemmingsen L.; Jancsó A. Hg2+ and Cd2+ binding of a bioinspired hexapeptide with two cysteine units constructed as a minimalistic metal ion sensing fluorescent probe. Dalton Trans. 2019, 48, 8327–8339. 10.1039/C9DT01141B. [DOI] [PubMed] [Google Scholar]
 - Łuczkowski M.; Zeider B. A.; Hinz A. V. H.; Stachura M.; Chakraborty S.; Hemmingsen L.; Huffman D. L.; Pecoraro V. L. Probing the Coordination Environment of the Human Copper Chaperone HAH1: Characterization of HgII-Bridged Homodimeric Species in Solution. Chem. – Eur. J. 2013, 19, 9042–9049. 10.1002/chem.201204184. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Iranzo O.; Thulstrup P. W.; Ryu S.-B.; Hemmingsen L.; Pecoraro V. L. The Application of 199Hg NMR and 199mHg PerturbedAngular Correlation (PAC) Spectroscopy to Define the Biological Chemistry of HgII: A Case Study with Designed Two- and Three-Stranded Coiled Coils. Chem. – Eur. J. 2007, 13, 9178–9190. 10.1002/chem.200701208. [DOI] [PubMed] [Google Scholar]
 - Li Y. J.; Weser U. Circular Dichroism, Luminescence, and Electronic Absorption of Copper Binding Sites in Metallothionein and Its Chemically Synthesized α and β Domains. Inorg. Chem. 1992, 31, 5526–5533. 10.1021/ic00052a031. [DOI] [Google Scholar]
 - Kelly S. M.; Price N. C. The Use of Circular Dichroism in the Investigation of Protein Structure and Function. Curr. Protein Pept. Sci. 2000, 1, 349–384. 10.2174/1389203003381315. [DOI] [PubMed] [Google Scholar]
 - Greenfield N. J.; Fasman G. D. Computed Circular Dichroism Spectra for the Evaluation of Protein Conformation. Biochemistry 1969, 8, 4108–4116. 10.1021/bi00838a031. [DOI] [PubMed] [Google Scholar]
 - Venyaminov S.; Baikalov I.; Shen Z.; Wu C.; Yang T. Circular dichroic analysis of denatured proteins: inclusion of denatured proteins in the reference set. Anal. Biochem. 1993, 214, 17–24. 10.1006/abio.1993.1450. [DOI] [PubMed] [Google Scholar]
 - Uversky V. N. Natively unfolded proteins: A point where biology waits for physics. Protein Sci. 2002, 11, 739–756. 10.1110/ps.4210102. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Cheng Y.; Yan Y.-B.; Liu J. Spectroscopic characterization of metal bound phytochelatin analogue (Glu–Cys)4–Gly. J. Inorg. Biochem. 2005, 99, 1952–1962. 10.1016/j.jinorgbio.2005.06.016. [DOI] [PubMed] [Google Scholar]
 - Perczel A.; Hollósi M.. Turns. In Circular Dichroism and the Conformational Analysis of Biomolecules; Fasman G. D., Ed.; Plenum Press: New York, 1996; pp 285–380. [Google Scholar]
 - Micsonai A.; Wien F.; Kernya L.; Lee Y.-H.; Goto Y.; Réfrégiers M.; Kardos J. Accurate secondary structure prediction and fold recognition for circular dichroism spectroscopy. Proc. Natl. Acad. Sci. U.S.A. 2015, 112, E3095–E3103. 10.1073/pnas.1500851112. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Pujol A. M.; Gateau C.; Lebrun C.; Delangle P. A Series of Tripodal Cysteine Derivatives asWater-Soluble Chelators that are Highly Selective for Copper(I). Chem. – Eur. J. 2011, 17, 4418–4428. 10.1002/chem.201003613. [DOI] [PubMed] [Google Scholar]
 - Filella M.; May P. M. Critical appraisal of available thermodynamic data for the complexation of antimony(III) and antimony(V) by low molecular mass organic ligands. J. Environ. Monit. 2005, 7, 1226–1237. 10.1039/b511453e. [DOI] [PubMed] [Google Scholar]
 - Magyar J. S.; Weng T.-C.; Stern C. M.; Dye D. F.; Rous B. W.; Payne J. C.; Bridgewater B. M.; Mijovilovich A.; Parkin G.; Zaleski J. M.; Penner-Hahn J. E.; Godwin H. A. Reexamination of Lead(II) Coordination Preferences in Sulfur-Rich Sites: Implications for a Critical Mechanism of Lead Poisoning. J. Am. Chem. Soc. 2005, 127, 9495–9505. 10.1021/ja0424530. [DOI] [PubMed] [Google Scholar]
 - Lukács M.; Pálinkás D. C.; Szunyog G.; Várnagy K. Metal Binding Ability of Small Peptides Containing Cysteine Residues. ChemistryOpen 2021, 10, 451–463. 10.1002/open.202000304. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Sun H.; Yan S. C.; Cheng W. S. Interaction of antimony tartrate with the tripeptide glutathione. Implication for its mode of action. Eur. J. Biochem. 2000, 267, 5450–5457. 10.1046/j.1432-1327.2000.01605.x. [DOI] [PubMed] [Google Scholar]
 - Rey N. A.; Howarth O. W.; Pereira-Maia E. C. Equilibrium characterization of the As(III)–cysteine and the As(III)–glutathione systems in aqueous solution. J. Inorg. Biochem. 2004, 98, 1151–1159. 10.1016/j.jinorgbio.2004.03.010. [DOI] [PubMed] [Google Scholar]
 - Charlier C.; Cousin S. F.; Ferrage F. Protein dynamics from nuclear magnetic relaxation. Chem. Soc. Rev. 2016, 45, 2410–2422. 10.1039/C5CS00832H. [DOI] [PubMed] [Google Scholar]
 - Jancsó A.; Gyurcsik B.; Mesterházy E.; Berkecz R. Competition of zinc(II) with cadmium(II) or mercury(II) in binding to a 12-mer peptide. J. Inorg. Biochem. 2013, 126, 96–103. 10.1016/j.jinorgbio.2013.05.019. [DOI] [PubMed] [Google Scholar]
 - Szunyogh D.; Szokolai H.; Thulstrup P. W.; Larsen F. H.; Gyurcsik B.; Christensen N. J.; Stachura M.; Hemmingsen L.; Jancsó A. Specificity of the Metalloregulator CueR for Monovalent Metal Ions: Possible Functional Role of a Coordinated Thiol?. Angew. Chem., Int. Ed. 2015, 54, 15756–15761. 10.1002/anie.201508555. [DOI] [PubMed] [Google Scholar]
 - Szunyogh D.; Gyurcsik B.; Larsen F. H.; Stachura M.; Thulstrup P. W.; Hemmingsen L.; Jancsó A. ZnII and HgII binding to a designed peptide that accommodates different coordination geometries. Dalton Trans. 2015, 44, 12576–12588. 10.1039/C5DT00945F. [DOI] [PubMed] [Google Scholar]
 - Szunyogh D.Study of the Metal Binding Properties of Cu-Efflux Regulator CueR Protein Via Model Peptides, Ph.D. Thesis; University of Szeged: Hungary, 2016. [Google Scholar]
 - Balogh R. K.; Gyurcsik B.; Hunyadi-Gulyás É.; Schell J.; Thulstrup P. W.; Hemmingsen L.; Jancsó A. C-terminal Cysteines of CueR Act as Auxiliary Metal Site Ligands upon HgII Binding—A Mechanism To Prevent Transcriptional Activation by Divalent Metal Ions?. Chem. – Eur. J. 2019, 25, 15030–15035. 10.1002/chem.201902940. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Jancsó A.; Correia J. G.; Balogh R. K.; Schell J.; Jensen M. L.; Szunyogh D.; Thulstrup P. W.; Hemmingsen L. A reference compound for 199mHg perturbed angular correlation of γ-rays spectroscopy. Nucl. Instrum. Methods Phys. Res., Sect. A 2021, 1002, 165154 10.1016/j.nima.2021.165154. [DOI] [Google Scholar]
 - Bauer R.; Jensen S. J.; Schmidt-Nielsen B. The angular overlap model applied to the calculation of nuclear quadrupole interactions. Hyperfine Interact. 1988, 39, 203–234. 10.1007/BF02394319. [DOI] [Google Scholar]
 - Manceau A.; Nagy K. L. Relationships between Hg(II)–S bond distance and Hg(II) coordination in thiolates. Dalton Trans. 2008, 1421–1425. 10.1039/b718372k. [DOI] [PubMed] [Google Scholar]
 - Mah V.; Jalilehvand F. Glutathione Complex Formation with Mercury(II) in Aqueous Solution at Physiological pH. Chem. Res. Toxicol. 2010, 23, 1815–1823. 10.1021/tx100260e. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Mah V.; Jalilehvand F. Mercury(II) complex formation with glutathione in alkaline aqueous solution. J. Biol. Inorg. Chem. 2008, 13, 541–553. 10.1007/s00775-008-0342-2. [DOI] [PubMed] [Google Scholar]
 - Neese F. Software update: The ORCA program system—Version 5.0. WIREs Comput. Mol. Sci. 2022, 12, e1606 10.1002/wcms.1606. [DOI] [Google Scholar]
 - Chakravorty D. K.; Wang B.; Lee C. W.; Giedroc D. P.; Merz K. M. Jr. Simulations of Allosteric Motions in the Zinc Sensor CzrA. J. Am. Chem. Soc. 2012, 134, 3367–3376. 10.1021/ja208047b. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Chakravorty D. K.; Parker T. M.; Guerra A. J.; Sherrill C. D.; Giedroc D. P.; Merz K. M. Jr. Energetics of Zinc-Mediated Interactions in the Allosteric Pathways of Metal Sensor Proteins. J. Am. Chem. Soc. 2013, 135, 30–33. 10.1021/ja309170g. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Capdevila D. A.; Braymer J. J.; Edmonds K. A.; Wu H.; Giedroc D. P. Entropy redistribution controls allostery in a metalloregulatory protein. Proc. Natl. Acad. Sci. U.S.A. 2017, 114, 4424–4429. 10.1073/pnas.1620665114. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Capdevila D. A.; Edmonds K. A.; Campanello G. C.; Wu H.; Gonzalez-Gutierrez G.; Giedroc D. P. Functional Role of Solvent Entropy and Conformational Entropy of Metal Binding in a Dynamically Driven Allosteric System. J. Am. Chem. Soc. 2018, 140, 9108–9119. 10.1021/jacs.8b02129. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Pennella M. A.; Giedroc D. P. Structural determinants of metal selectivity in prokaryotic metal-responsive transcriptional regulators. BioMetals 2005, 18, 413–428. 10.1007/s10534-005-3716-8. [DOI] [PubMed] [Google Scholar]
 - Jumper J.; Evans R.; Pritzel A.; Green T.; Figurnov M.; Ronneberger O.; Tunyasuvunakool K.; Bates R.; Žídek A.; Potapenko A.; Bridgland A.; Meyer C.; Kohl S. A. A.; Ballard A. J.; Cowie A.; Romera-Paredes B.; Nikolov S.; Jain R.; Adler J.; Back T.; Petersen S.; Reiman D.; Clancy E.; Zielinski M.; Steinegger M.; Pacholska M.; Berghammer T.; Bodenstein S.; Silver D.; Vinyals O.; Senior A. W.; Kavukcuoglu K.; Kohli P.; Hassabis D. Highly accurate protein structure prediction with AlphaFold. Nature 2021, 596, 583–589. 10.1038/s41586-021-03819-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Varadi M.; Anyango S.; Deshpande M.; Nair S.; Natassia C.; Yordanova G.; Yuan D.; Stroe O.; Wood G.; Laydon A.; Žídek A.; Green T.; Tunyasuvunakool K.; Petersen S.; Jumper J.; Clancy E.; Green R.; Vora A.; Lutfi M.; Figurnov M.; Cowie A.; Hobbs N.; Kohli P.; Kleywegt G.; Birney E.; Hassabis D.; Velankar S. AlphaFold Protein Structure Database: massively expanding the structural coverage of protein-sequence space with high-accuracy models. Nucleic Acids Res. 2022, 50, D439–D444. 10.1093/nar/gkab1061. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Zheng W.; Zhang C.; Li Y.; Pearce R.; Bell E. W.; Zhang Y. Folding non-homology proteins by coupling deep-learning contact maps with I-TASSER assembly simulations. Cell Rep. Methods 2021, 1, 100014 10.1016/j.crmeth.2021.100014. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Zhang C.; Freddolino P. L.; Zhang Y. COFACTOR: improved protein function prediction by combining structure, sequence and protein-protein interaction information. Nucleic Acids Res. 2017, 45, W291–299. 10.1093/nar/gkx366. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Yang J.; Zhang Y. I-TASSER server: new development for protein structure and function predictions. Nucleic Acids Res. 2015, 43, W174–W181. 10.1093/nar/gkv342. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Cong X.; Yuan Z.; Wang Z.; Wei B.; Xu S.; Wang J. Crystal structures of manganese-dependent transcriptional repressor MntR (Rv2788) from Mycobacterium tuberculosis in apo and manganese bound forms. Biochem. Biophys. Res. Commun. 2018, 501, 423e427 10.1016/j.bbrc.2018.05.005. [DOI] [PubMed] [Google Scholar]
 - Lee M. Y.; Lee D. W.; Joo H. K.; Jeong K. H.; Lee J. Y. Structural analysis of the manganese transport regulator MntR from Bacillus halodurans in apo and manganese bound forms. PLoS One 2019, 14, e0224689 10.1371/journal.pone.0224689. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Stoll K. E.; Draper W. E.; Kliegman J. I.; Golynskiy M. V.; Brew-Appiah R. A. T.; Phillips R. K.; Brown H. K.; Breyer W. A.; Jakubovics N. S.; Jenkinson H. F.; Brennan R. G.; Cohen S. M.; Glasfeld A. Characterization and Structure of the Manganese-Responsive Transcriptional Regulator ScaR. Biochemistry 2009, 48, 10308–10320. 10.1021/bi900980g. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Changela A.; Chen K.; Xue Y.; Holschen J.; Outten C. E.; O’Halloran T. V.; Mondragón A. Molecular Basis of Metal-Ion Selectivity and Zeptomolar Sensitivity by CueR. Science 2003, 301, 1383–1387. 10.1126/science.1085950. [DOI] [PubMed] [Google Scholar]
 - Philips S. J.; Canalizo-Hernandez M.; Yildirim I.; Schatz G. C.; Mondragón A.; O’Halloran T. V. Allosteric transcriptional regulation via changes in the overall topology of the core promoter. Science 2015, 349, 877–881. 10.1126/science.aaa9809. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Chang C.-C.; Lin1 L.-Y.; Zou X.-W.; Huang C.-C.; Chan N.-L. Structural basis of the mercury(II)-mediated conformational switching of the dual-function transcriptional regulator MerR. Nucleic Acids Res. 2015, 43, 7612–7623. 10.1093/nar/gkv681. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Schreiter E. R.; Wang S. C.; Zamble D. B.; Drennan C. L. NikR–operator complex structure and the mechanism of repressor activation by metal ions. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 13676–13681. 10.1073/pnas.0606247103. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Eicken C.; Pennella M. A.; Chen X.; Koshlap K. M.; VanZile M. L.; Sacchettini J. C.; Giedroc D. P. A Metal–Ligand-mediated Intersubunit Allosteric Switch in Related SmtB/ArsR Zinc Sensor Proteins. J. Mol. Biol. 2003, 333, 683–695. 10.1016/j.jmb.2003.09.007. [DOI] [PubMed] [Google Scholar]
 - Ye J.; Kandegedara A.; Martin P.; Rosen B. P. Crystal Structure of the Staphylococcus aureus pI258 CadC Cd(II)/Pb(II)/Zn(II)-Responsive Repressor. J. Bacteriol. 2005, 187, 4214–4221. 10.1128/JB.187.12.4214-4221.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Kandegedara A.; Thiyagarajan S.; Kondapalli K. C.; Stemmler T. L.; Rosen B. P. Role of Bound Zn(II) in the CadC Cd(II)/Pb(II)/Zn(II)-responsive Repressor. J. Biol. Chem. 2009, 284, 14958–14965. 10.1074/jbc.M809179200. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - Baxter N. J.; Williamson M. P. Temperature dependence of 1H chemical shifts in proteins. J. Biomol. NMR 1997, 9, 359–369. 10.1023/A:1018334207887. [DOI] [PubMed] [Google Scholar]
 - Cierpicki T.; Otlewski J. Amide proton temperature coefficients as hydrogen bond indicators in proteins. J. Biomol. NMR 2001, 21, 249–261. 10.1023/A:1012911329730. [DOI] [PubMed] [Google Scholar]
 
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.








