Abstract
Metallo-β-lactamase (MBL) production in Gram-negative bacteria is an important contributor to β-lactam antibiotic resistance. Combining β-lactams with β-lactamase inhibitors (BLIs) is a validated route to overcoming resistance, but MBL inhibitors are not available in the clinic. Based on zinc utilization and sequence, MBLs are divided into three subclasses, B1, B2 and B3, whose differing active-site architectures hinder development of BLIs capable of “cross-class” MBL inhibition. We previously described 2-mercaptomethyl thiazolidines (MMTZs) as B1 MBL inhibitors (e.g. NDM-1) and here show that inhibition extends to the clinically relevant B2 (Sfh-I) and B3 (L1) enzymes. MMTZs inhibit purified MBLs in vitro (e.g. Sfh-I, Ki 0.16 μM) and potentiate β-lactam activity against producer strains. X-ray crystallography reveals that inhibition involves direct interaction of the MMTZ thiol with the mono- or di-zinc centers of Sfh-I/L1, respectively. This is further enhanced by sulfur-π interactions with a conserved active site tryptophan. Computational studies reveal that the stereochemistry at chiral centers is critical, showing less potent MMTZ stereoisomers (up to 800-fold) as unable to replicate sulfur-π interactions in Sfh-I, largely through steric constraints in a compact active site. Furthermore, in silico replacement of the thiazolidine sulfur with oxygen (forming an oxazolidine) resulted in less favorable aromatic interactions with B2 MBLs, though the effect is less than previously observed for the subclass B1 enzyme NDM-1. In the B3 enzyme L1, these effects are offset by additional MMTZ interactions with the protein main chain. MMTZs can therefore inhibit all MBL classes by maintaining conserved binding modes through different routes.
Keywords: antibiotic resistance, β-lactamases, inhibitors, carbapenemase
Graphical Abstract

Introduction
β-Lactams are the most prescribed antibiotic class worldwide, but their efficacy is increasingly challenged by the growing problem of antibiotic resistance.1 The production of β-lactamases in Gram-negative bacteria is the major resistance mechanism in the clinic, as members of this large enzyme family are capable of inactivating all β-lactam antibiotics.2 β-Lactamases can be divided into two mechanistic groups, the serine-β-lactamases (Ambler classes A, C and D3, 4, SBLs) and the zinc ion dependent metallo-β-lactamases (class B4, MBLs). SBL-catalyzed hydrolysis involves attack of a nucleophilic serine on the β-lactam ring, and occurs with formation and resolution of a covalent acyl-enzyme intermediate via labile tetrahedral species.5 Conversely, MBLs utilize an active site water/hydroxide to hydrolyze the β-lactam ring without formation of a covalent intermediate (Figure 1A).5
Figure 1. MBL-catalyzed β-lactam breakdown and MBL inhibitors.
(A) Penicillin breakdown by MBLs, resulting in penicilloic acid product formation via a proposed high-energy tetrahedral intermediate. (B) Inhibitors based on proposed species in β-lactam hydrolysis. Left, bisthiazolidine (BTZ) scaffold, R=H/CH3; middle, bicyclic boronates (e.g. taniborbactam, QPX7728); right, 2-mercaptomethyl thiazolidine (MMTZ) scaffold, R=H/CH3. (C) 2-Mercaptomethyl thiazolidines studied here. The absolute R/S configurations of the chiral carbon centers (C-2 and C-4, labelled) are shown. L-anti-1a (brown), D-anti-1a (orange), L-anti-1b (green), L-syn-1b (cyan), D-anti-1b (purple) and D-syn-1b (pink). (C) is adapted from Rossi et al.6
MBLs can be further subdivided into the B1, B2 and B3 subclasses, based on active site architecture and zinc utilisation.5, 7 The B1 enzymes,8 such as NDM-1, VIM-2 and IMP-1, are plasmid mediated and widespread, produced in Gram-negative bacteria such as Klebsiella pneumoniae, Escherichia coli and Pseudomonas aeruginosa. B1 enzymes have a wide substrate specificity and can hydrolyze almost all β-lactam antibiotics, including the carbapenems, such as imipenem and meropenem, that were once considered antibiotics of last-resort. The active site of B1 enzymes is conserved, containing two Zn(II) ions, termed Zn1 and Zn2, that are coordinated by three amino acids each. Zn1 is coordinated by three histidine residues (His116, His118 and His 196; standard MBL numbering throughout9) and Zn2 by Asp120, Cys221 and His263. The Zn(II) ions are bridged by a water/hydroxide that is thought to act as the nucleophile in β-lactam hydrolysis. B2 MBLs are mono-zinc enzymes, containing a zinc ion coordinated as in the Zn2 site of B1 MBLs. To date, they have been identified on the chromosomes of the Serratia fonticola (Sfh-I10), Aeromonas sp. (CphA11), and Pseudomonas sp. (PFM12), organisms capable of causing a range of opportunistic infections in the clinic. B2 MBLs are carbapenemases but have a narrower substrate profile than the B1 enzymes, as they are unable to hydrolyze most penicillins and cephalosporins.13–16 In contrast, in the di-zinc B3 enzymes, Zn1 is coordinated as in B1 MBLs, but Zn2 by Asp120, His121 and His263, resulting in a different active site architecture to the B1 and B2 MBLs. They are found on the chromosomes of numerous environmental and pathogenic Gram-negative bacteria, including Stenotrophomonas maltophilia (L117, 18), Elizabethkingia meningoseptica (GOB19, 20), and Pseudomonas otitidis (POM-121). Some have now been shown to be incorporated into the chromosomes of pathogenic bacteria via a mobile genetic element,22, 23 indicating the potential for wide dissemination. Taken together with their wide substrate spectrum, that includes efficient carbapenem hydrolysis, B3 enzymes could pose an increasing clinical concern.
Inhibitors closely mimicking β-lactam molecules (e.g. clavulanic acid and tazobactam), diazabicyclooctanes (DBOs, e.g. avibactam) and vaborbactam (a cyclic boronate) are now used in the clinic to inactivate some serine-β-lactamases through covalent attachment to the nucleophilic serine.24–26 However, the mechanistically distinct MBLs are not inhibited and can hydrolyze the β-lactam based inhibitors.27 This has prompted extensive investigations into MBL inhibitor development, with bicyclic boronates currently representing the most promising compounds (Figure 1B).28, 29 For example, taniborbactam (formerly VNRX-5133), currently in late-stage clinical development, is a potent inhibitor of most SBLs and B1 MBLs.28–31 In addition, QPX7728 is in phase 1 clinical trials and has an improved spectrum of activity compared to taniborbactam, particularly against SBLs and the B1 MBL IMP-1.32, 33 Despite these significant advances in MBL inhibitor development, there remains a need for further exploration of compounds active against all MBL subclasses that will increase our available armamentarium to combat antibiotic resistance and guard against the possible future dissemination of enzymes beyond subclass B1.
We considered that common features of species populated during antibiotic hydrolysis by diverse MBLs could be exploited for inhibitor development.34, 35 To this end, we previously reported on the synthesis of a series of bisthiazolidines (BTZs) that were designed as bicyclic substrate mimics, containing a carboxylate and additional free thiol as a zinc-binding moiety (Figure 1B).36 We demonstrated that BTZ stereoisomers inhibited all MBL subclasses by adopting multiple binding modes in the structurally different active sites.37
More recently, we described a series of 2-mercaptomethyl thiazolidines (MMTZs) that were designed to exploit common features of the binding of hydrolyzed β-lactam products to the MBL active site (Figure 1B).6 These compounds contain a thiazolidine ring and two chiral carbon centers with carboxylate and free thiol groups (Figure 1C). Some MMTZs also include a gem-dimethyl group that is also present in penicillins and some BTZs. They inhibit the clinically important B1 enzymes NDM-1, VIM-2 and IMP-1 by adopting consistent binding modes that exploit thiol-Zn coordination, and an interaction of the thiazolidine sulfur with a conserved aromatic residue in the active site.
Here, we utilize enzyme kinetics, microbiology, X-ray crystallography and computational chemistry techniques to demonstrate that MBL inhibition by MMTZs extends to the subclass B2 and B3 enzymes, Sfh-I and L1. These data further explore the roles of conserved interactions in the active site, including the sulfur-π interaction, in maintaining a conserved binding mode. The results highlight that exploiting common features of the MBL active site can represent a tractable route towards achieving cross-class inhibition and new drug development.
Results and Discussion
MMTZs are cross-class MBL inhibitors in vitro.
To determine the inhibition profile of MMTZs against MBLs, we first measured inhibitory constants (Ki values) against the clinically relevant B2 and B3 enzymes Sfh-I and L1 by monitoring imipenem hydrolysis in the presence of MMTZs (Figures S1 and S2). Inhibition profiles could then be compared with our previously collected Ki data for the B1 MBLs (Table 1).
Table 1.
MMTZ inhibition constants (Ki, μM) against purified, recombinant metallo-β-lactamases
| B2 MBL | B3 MBL | B1 MBLsa,b | |||
|---|---|---|---|---|---|
|
| |||||
| Inhibitor | Sfh-I | L1 | NDM-1 | VIM-2 | IMP-1 |
|
| |||||
| L-anti-1a | 0.16 ± 0.03 | 10.0 ± 0.8 | 5.2 ± 0.7 | 0.38 ± 0.05 | 1.0 ± 0.2 |
| D-anti-1a | 20 ± 2 | 4.0 ± 0.5 | 2.5 ± 0.5 | 0.39 ± 0.04 | 1.3 ± 0.1 |
| L-anti-1b | 1.3 ± 0.1 | 20 ± 2 | 0.44 ± 0.06 | 0.75 ± 0.09 | 0.46 ± 0.05 |
| L-syn-1b | 100 ± 10 | 28 ± 3 | 8 ± 1 | 3.6 ± 0.4 | 6.0 ± 0.6 |
| D-anti-1b | 130 ± 10 | 1.4 ± 0.2 | 3.1 ± 0.3 | 0.9 ± 0.1 | 0.93 ± 0.08 |
| D-syn-1b | 1.0 ± 0.1 | 4.0 ± 0.6 | 0.6 ± 0.05 | 1.9 ± 0.1 | 2.0 ± 0.2 |
| L-BTZ-1b | 0.26 ± 0.03 | 12 ± 1 | 7 ± 1 | 2.9 ± 0.4 | 8 ± 2 |
| D-BTZ-1b | 26 ± 3 | 10 ± 1 | 19 ± 3 | 3.2 ± 0.4 | 6 ± 1 |
| L-BTZ-2b | 0.36 ± 0.04 | 11 ± 2 | 18 ± 3 | 6 ± 1 | 15 ± 3 |
| D-BTZ-2b | 29 ± 3 | 10 ± 1 | 12 ± 1 | 14 ± 3 | 10 ± 2 |
B1 MMTZ data are from Rossi et al6 following the same protocol as used here.
All BTZ data are from Hinchliffe et al, calculated following the same protocol as used here.37 BTZs −1 and −2 are with and without a gem-dimethyl group, respectively. See Figure S3A for chemical structures of the L-/D-BTZ stereoisomers.
Our results show that MMTZs potently inhibit all MBL subclasses, although there are clear differences between different enantiomers. This is most apparent with Sfh-I for which Ki values range from 0.16 μM to 130 μM, an ∼800-fold difference. In particular, there are ∼100-fold reductions in Ki between L/D-anti-1a, L/D-anti-1b, and L/D-syn-1b. Further, compounds in which C-2 of the thiazolidine ring is in the (R)-configuration have poor potency compared to the (S)-isomer, with Ki values of 20, 100 or 130 μM (D-anti-1a, L-syn-1b and D-anti-1b, respectively). We also found BTZs to be 100-fold less potent against Sfh-I when the chiral carbon center bearing the mercaptomethyl group is in the S-configuration, i.e. the D-enantiomer (Figure S3A). However, unlike the BTZs, the presence of a gem-dimethyl group in MMTZs also adversely affects their potency against Sfh-I, with a c. 10-fold reduction in Ki between L-anti-1a/1b and D-anti-1a/1b. Variation in the inhibitory activity of MMTZs against L1 is less apparent (as is also the case for BTZs), although a slight preference for the D-over the L-isomer is observed (2.5-/7-/14-fold Ki increases for anti-1a, anti-1b and syn-1b L/D stereoisomers, respectively). The two compounds most active against the range of MBLs tested here, and previously,6 are L-anti-1a and D-syn-1b. In contrast, D- and L-captopril, compounds with a free thiol and pyrrolidine ring (Figure S3B), display particularly poor potency against a B2 enzyme (Ki’s 72/950 μM against CphA), and are up to 20-fold less potent than MMTZs against B3 enzymes (Ki’s 20/8.8 μM against L1/SMB-1) (Table S1).
We next sought to determine whether MMTZs could potentiate antibiotic activity against recombinant E. coli expressing NDM-1, Sfh-I and L1, by measuring their effect on the minimal inhibitory concentration (MIC) of imipenem (chosen as Sfh-I predominantly hydrolyzes carbapenems, and for consistency with our previous work6) in broth microdilution assays (Figure 2).
Figure 2. Effect of MMTZs on imipenem minimum inhibitory concentrations of MBL-expressing E. coli.
The minimum inhibitory concentration of imipenem (IMI) +/− 100 μg/mL MMTZ was determined for the E. coli strain DH10B expressing NDM-1, Sfh-I or L1 (see Methods for details). At this concentration, in the absence of antibiotic, MMTZs do not have a detrimental effect on bacterial growth. An asterisk denotes the inhibitor had a four-fold effect on MIC (i.e. two dilutions). A two-fold effect (i.e. one dilution factor) is not considered significant. Results are the mode of three biological replicates.
At a concentration of 100 μg/mL, MMTZs display variable effectiveness in lowering the imipenem MIC of MBL-expressing E. coli DH10B. Although some of the compounds only cause a two-fold reduction in imipenem susceptibility, a number result in a more significant four-fold reduction (i.e. a factor of two dilutions, see asterisks in Figure 2). Indeed, compound L-anti-1a was the most effective, causing a 4-fold MIC reduction against E. coli producing NDM-1, Sfh-I or L1. The compound that was most potent against purified, recombinant MBLs, D-syn-1b, surprisingly showed the least effect against these MBL-expressing bacteria. To test whether lower concentrations of inhibitor could potentiate activity, we tested the effect of 50 and 75 μg/mL on the imipenem MIC against the L1-expressing E. coli DH10B (Table S2). These data indicate that 50 μg/mL of L-anti-1b, D-syn-1b and L-syn-1b all caused a four-fold reduction in imipenem MIC, indicating lower concentrations of inhibitor can potentiate imipenem activity against a lab strain carrying an efficient carbapenemase. However, we note that concentrations of 100 μg/mL were required for effective imipenem potentiation against a number of clinical strains expressing B1 MBLs.6
We have previously reported that MMTZs are relatively stable in solution (PBS), with 12% forming disulfides after 6 hours, and 88% remaining intact.6 It is possible that, over the time course of an MIC assay, this may have a small, but limited, effect on the MIC profiles, contributing to the discrepancy between Ki values and MICs, but we consider it unlikely that this is the sole explanation. We note that penetration into the bacterial periplasm may also be a contributing factor, as our previously determined in cell IC50 values against NDM-1 were also slightly elevated compared to in vitro Ki values.6
MMTZs therefore inhibit all MBL subclasses and can potentiate carbapenem activity against recombinant E. coli expressing enzymes from all three MBL classes. These in vitro data also highlight the importance of considering inhibitor stereochemistry for cross-class inhibition, with stereochemical preferences particularly marked in the case of the B2 enzymes which have a narrower substrate specificity. In contrast, inhibition of the B1 MBLs was less affected by stereochemistry, with the B3 enzymes displaying moderate selectivity towards the different stereoisomers.
X-ray crystallography.
Having previously established that MMTZs adopt a consistent binding mode to B1 MBLs,6 we next sought to determine how they interact with the active sites of both Sfh-I and L1 by X-ray crystallography. We therefore soaked pre-formed MBL crystals in MMTZs and obtained high-resolution structures of L-anti-1a bound to Sfh-I and D-syn-1b to L1 (Table S3). These represent the two compounds that can be considered the most potent over all three MBL subclasses tested (Table 1). Sfh-I crystallized in space group P21 with two molecules in the asymmetric unit (ASU) and L1 in space group P6422 with one molecule in the ASU. In both cases, bound MMTZ could be modelled into clearly defined Fo-Fc electron density (Figure S4) with real space correlation coefficients calculated by the PDB after refinement of 0.96/0.97 (Sfh-I, chains A and B) and 0.95 (L1).
MMTZ binding involves interaction of the thiol directly with the mono-zinc center of Sfh-I, or bridging the di-zinc center of L1 and displacing the catalytic water/hydroxide (Figure 3). In Sfh-I, there is a weak hydrogen bond between the ethyl ester carbonyl of L-anti-1a and the side chain nitrogen of Asn233 (3.2 A). In contrast, D-syn-1b binding to L1 is stabilized by stronger hydrogen bonds with the side chain oxygens of Ser223 (2.5 Å) and Tyr33 (2.9 Å).
Figure 3. Interactions of MMTZs in the active site of L1 and Sfh-I.
Left, thiol-Zn interactions and hydrogen bonds in MMTZ:MBL complexes (distance labelled). Right, interactions with hydrophobic residues lining the active site (blue sticks). (A) L-anti-1a (brown) binding to Sfh-I; (B) D-syn-1b (pink) binding to L1.
The binding mode of D-syn-1b to L1 has similarities with its binding to the di-zinc B1 enzymes NDM-1 and VIM-2, particularly with respect to the positioning of the thiol and carboxylate groups (Figure S5). However, the C-2 ethyl ester side chain is rotated 180°, and therefore pointing in the opposite direction in L1. Furthermore, in the B1 enzymes there is only a single (c. 2.6 – 3 Å) interaction with the backbone nitrogen of Asn233, in comparison to the hydrogen bonds D-syn-1b makes with two amino acid residues in L1.
Structure-activity relationships and conserved sulfur-π interactions.
X-ray-crystallographic data, both presented here and previously,6 therefore indicate that MMTZs employ a consistent binding mode to the active sites of diverse MBLs, especially when compared with the diverse modes of binding previously observed for complexes of the bicyclic BTZs. This is highlighted by the identification that, in B1 MBLs, MMTZ binding utilizes a conserved sulfur-π interaction between the thiazolidine sulfur atom and an active site aromatic residue (Trp87 in NDM-1/VIM-2 or Phe87 in IMP-1). We note that this interaction is conserved in the B2 and B3 enzymes Sfh-I and L1, with distances comparable to those observed in the B1 enzymes (Figure 4).
Figure 4. Sulfur-π interactions of MMTZs in MBL active sites.
Views from the active sites of MBLs with dashes showing the interaction (distances labelled) of the MMTZ sulfur with an active site tryptophan (blue sticks) in (A) NDM-1:L-anti-1b (PDB 6zyp6), (B) NDM-1:D-syn-1b (PDB 6zyq6), (C) Sfh-I:L-anti-1a (PDB 7bj9) and (D) L1:D-syn-1b (PDB 7bj8).
Notably, the sulfur-π interaction is not conserved for the bicyclic thiazolidine stereoisomers (bisthiazolidines, BTZs), for which diverse binding modes allowed cross-class inhibition.37 Indeed, in comparison with their respective MMTZ analogues L-/D-BTZs bind very differently to both Sfh-I and L1 (Figure S6) – the BTZ carboxylate, rather than the thiol, interacts with the Zn(II) ion in Sfh-I, and in L1 the BTZ thiazolidine ring is flipped in comparison to that of the MMTZ. These differences, that are also observed for the B1 enzymes,6 perhaps reflect the more constrained bicyclic ring scaffold of BTZs which does not afford these compounds the flexibility to adopt consistent binding modes that exploit the sulfur-π interaction.
Furthermore, based on comparisons with X-ray crystal structures of biapenem bound to the B2 enzyme CphA (PDB 1×8i38), and hydrolyzed penicillin G bound to L1 (PDB 6u0z39) (Figure S7), MMTZ binding does not closely reflect that of β-lactam derived hydrolysis products. To date, there are no crystal structures available for B2 MBLs in complex with penicillin-derived products as penicillin is not hydrolyzed by these enzymes, so direct comparisons cannot be made between the orientations of the respective thiazolidine rings. However, the carboxylate of the biapenem-derived product in the CphA complex points in the opposite direction to the MMTZ carboxylate in Sfh-I, indicative of likely differing binding modes adopted by MMTZs and β-lactams (Figure S7A). The pyrroline ring of hydrolyzed biapenem lies in the same plane as the MMTZ thiazolidine, although interactions of the pyrroline N with Zn2 place the ring more deeply in the active site and closer to the Zn(II) ion than for the MMTZ. In the case of L1 (Figure S7B), both MMTZ and hydrolyzed penicillin G hydrogen bond with active site residues Tyr33 and Ser223. However, the thiazolidine rings lie at 90° to one another, resulting in a much weaker sulfur-π interaction for the hydrolyzed antibiotic, compared to that possible for the MMTZ. We previously observed the same relationship when comparing MMTZ/antibiotic binding in NDM-1.6
To understand the importance of these conserved interactions, and particularly the structure-activity relationships apparent for Sfh-I, we modelled in silico the MMTZs D-anti-1a, L-syn-1b and D-anti-1b that are ∼125 – 800-fold less potent against Sfh-I than L-anti-1a. We first generated structures of MMTZs docked in the active site as their anion (see Methods) and subsequently optimized these with quantum mechanical/molecular mechanics (QM/MM) simulations. This procedure was validated by closely reproducing the L-anti-1a binding mode to Sfh-I determined crystallographically, particularly maintaining sulfur-π and thiol-Zn interactions (Figure S8). These optimized, crystallographically intractable, complexes resulted in an increase of the sulfur-n interaction from 6.11 Å in the QM/MM optimized crystal structure (PDB 7bj9, L-anti-1a) to 7.50 Å (D-anti-1a) and 6.72 Å (L-syn-1b) (Figure S9A and S9B, respectively). These binding modes most likely arise due to limited space in the active site for the ligand to occupy, preventing conformations that maintain the sulfur-π interaction. This also results in a conformationally constrained L-syn-1b (625-fold less potent than L-anti-1a), in which its carbonyl oxygen is in an energetically unfavorable position, 1.7 Å from the thiol sulfur, further contributing to the poor potency of this MMTZ isomer. In contrast, D-anti-1b binds to the active site with the thiazolidine sulfur pointing away from Trp87, and in the opposite direction to the other docking modes (and the crystallographic structure of the L-anti-1a complex), further highlighting that the sulfur-π interaction cannot be maintained (Figure S9C). These constraints on the ligand geometry are consistent with the narrower substrate profile of Sfh-I compared with most B1 or B3 enzymes. These structural data rationalize the significant structure-activity relationships, particular the impact of the stereochemistry on inhibitory potency, for MMTZs against Sfh-I, where Ki values for compounds with C-2 in the (R)-configuration were significantly poorer than when C-2 is in the (S)-configuration. Furthermore, our crystallographic data support these observations by showing that the addition of a gem-dimethyl group to MMTZs can result in steric clashes with Val67 (Figure 3A, right), contributing to the ∼10-fold reduction in potency of the 1b compounds against Sfh-I (Table 1).
It has previously been shown that potency can be less for inhibitors which have an oxazolidine/triazole rather than thiazolidine/thiadiazole ring,40 highlighting the potential importance of the sulfur atom for inhibition. Therefore, to further explore the role of the sulfur-π interaction in B2 and B3 MBLs we selectively replaced the thiazolidine sulfur with an oxygen atom in silico and optimized the complexes with resultant isosteric 2-mercaptomethyl oxazolidines (MMOZs) with QM/MM. This in silico approach was necessary as the oxazolidine compounds were synthetically intractable due to instability issues. Our previous QM/MM simulations with NDM-16 using the same methodology revealed that MMOZs caused an increase in the interaction distance between the O/S of the MMOZ/MMTZ and the conserved active site aromatic residue Trp87 (Figure 5A and 5B). Here, we note that the effect in Sfh-I and L1 is less prominent than in the B1 NDM-1:L-anti-1a complex (Figure 5 and Table S4). In the case of NDM-1:L-anti-1a, the larger effect was caused by significant movement of the MMOZ in the NDM-1 active site, that is not observed in the Sfh-I and L1 simulations, nor in the NDM-1:D-syn-1b complex.
Figure 5. QM/MM optimized structures of MMTZs and their analagous oxazolidines in MBLs.
Views from the active show overlays of QM/MM optimized complex structures for MMTZs (colored as in Figure 3) and their MMOZ analogues (cyan) bound in MBLs. NDM-1 simulations were performed previously.6 (A) NDM-1:L-anti-1a; (B) NDM-1:D-syn-1b; (C) Sfh-I:L-anti-1a; (D) L1:D-syn-1b.
Although MMOZ movement in the NDM-1:D-syn-1b and Sfh-I:L-anti-1a complexes is limited, the distance of the heteroatom to Trp87 still increases compared to its equivalent in the MMTZ complexes. This can be partly explained by the shorter C-O, compared to C-S, bond distance, resulting in the oxygen atom being further from Trp87 than is the sulfur. In L1, however, the distance increase is minimal (0.13 Å), which can be explained by the increased number of hydrogen bonds formed by D-syn-1b in the active site potentially outweighing the effect of the O/S-aromatic interaction. Substitution of the sulfur with an oxygen therefore has variable effects in silico, but the data are suggestive that the sulfur-π interaction exerts greatest influence in the NDM-1:L-anti-1a complex. Indeed, the S to O substitution appears to only affect structures in which the hydrogen bonds in the active site are minimal (NDM-1 and Sfh-I) but has a less substantive contribution when inhibitor binding is stabilized by stronger hydrogen bonds (L1). Further, inhibitor binding in Sfh-I is more likely driven by constraints imposed by the narrow active site, with a reduced contribution from S-π interactions.
Conclusions
In summary, our analysis demonstrates, and provides the basis for, cross-class inhibition of MBLs by MMTZs. Although S-π interactions contribute to MMTZ:MBL interactions in all cases, our data suggest that different factors are responsible for establishing the broadly similar binding modes in the three structurally different active sites. In B1 MBLs such as NDM-1, S-π interactions most likely dominate; B2 enzymes (e.g. Sfh-I) have a constrained active site that can sterically clash with MMTZs, particularly those with a gem-dimethyl group or C-2 in the (S)-configuration; whereas in B3 enzymes, the contribution of hydrogen bonds to binding appears to override the S-π interaction. Differences in potency, and the potential for steric clashes with the active site, highlight the need to consider alternative inhibitor stereoisomers to achieve cross-class MBL inhibition, and more generally in future inhibitor development. MMTZs therefore represent a promising MBL inhibitor scaffold that is synthetically accessible, allowing for the design of future iterations that improve potency in bacterial cells
Material and Methods
Synthesis of 2-mercaptomethyl thiazolidines.
MMTZs were synthesized as previously described.6 Purity was determined by HPLC to be >95% as described in Rossi et al.6
Inhibition constants.
Inhibition constants were determined as previously described.6 In detail, Ki values were determined under steady-state conditions by following imipenem hydrolysis by Sfh-I and L1 at 30 °C. Imipenem breakdown was measured as a decrease in absorbance at 300 nm (Δε300 = −9000 M−1cm−1) in a Jasco V-670 spectrophotometer. Reactions were setup in a 0.1 cm path length quartz cuvette with a final volume of 300 μL. Final enzyme concentration was 2 nM. Reaction buffer contained 10 mM HEPES pH 7.5, 200 mM NaCl and 50 mg/L BSA. Reaction buffer for L1 was supplemented with 20 μM ZnSO4. MMTZs were dissolved in DMSO at a final concentration of 100 mM and diluted in the reaction buffer to the desired concentration. The DMSO concentration (0.07 %) does not affect the enzyme activity (data not shown). The reactions were started and monitored with the addition of the MBLs to a substrate and inhibitor mixture. Linear time courses were observed for all the conditions and less than 5 % of the substrate was consumed after 300 s (Figure S3). Reaction rates were obtained from the slope of these time courses and fitted with the Competitive Inhibition model in GraphPad Prism 5.0, from which the inhibition constant (Ki) could be calculated (Figure S4).
The same protocol was followed for the previously published Ki determination of the MMTZ compounds with B1 enzymes6 and the BTZ compounds with B136, B2 and B3 enzymes.37
Microbiology.
Minimal inhibitory concentrations (MICs) for imipenem were determined by the broth microdilution method in cation-adjusted Mueller Hinton broth according to Clinical and Laboratory Standards Institute (CLSI) guidelines.41 Imipenem was chosen as a representative carbapenem as Sfh-I is predominantly a carbapenemase. E. coli DH10B cells were transformed with pMBLe (containing blaNDM-1 or blaSfh-I) or pBC-SK (containing blaL1). NDM-1 and Sfh-I expression was induced with 0.1 mM IPTG. MMTZs were dissolved to 250 mM in DMSO before dilution to 100 μg/mL in serial doubling dilutions of imipenem (from 64 mg/L to 0.025 mg/L). Results presented are the mode of three biological replicates.
X-ray crystallography.
L1 and Sfh-1 were produced, purified and crystallized as previously described.37, 42 L1 crystallized by mixing 1 μL protein (23 mg/mL in 10 mM Tris pH 7, 5 mM ZnSO4, 100 mM NaCl) with 1 μL crystallization reagent [0.1 M HEPES pH 7.75, 1.5% PEG400, 2 M (NH4)2SO4].37 Sfh-I was crystallized by mixing, in a 3:2 ratio, protein (15 mg/mL, in 50 mM HEPES pH 7.5, 10% glycerol) with 0.2 M sodium acetate and 27% wt/vol PEG 3350.42 MMTZs were dissolved in crystallization buffer at 5 mM and soaked into pre-formed crystals for 1 h 20 m (L1) or 1 h (Sfh-I). Diffraction data were collected at Diamond I03 (Sfh-I) or I04-1 (L1). Reflections were indexed and integrated in Mosflm (L1) or Dials43 (Sfh-I) and scaled in Aimless.44 Phases were calculated with molecular replacement in Phaser45 using 5evd37 or 5ew037 as search models for L1 and Sfh-I, respectively. Structural models were completed with iterative rounds of refinement in Coot46 and Phenix47 Ligand restraints were calculated in eLBOW47, 48 and modelled into clearly defined Fo-Fc electron density. Figures were created in PyMol.49
Computational calculations.
Starting structure preparation
All ligands were prepared as thiolate anions, with the carboxylate also deprotonated. Oxazolidine complexes were built in silico by replacing the sulfur atom with an oxygen atom in the corresponding L1 and Sfh-I crystal structures, PDBs 7bj8 and 7bj9, respectively.
Sfh-I complexes with crystallographically intractable MMTZs were initially docked in Autodock 4.2.650 using the solvent sites biased Autodock4 docking method (SSBMD).51 Using information from the available crystal structures, we modified the Autodock4 scoring function by adding an energy term for the thiol sulfur atom to the original function. The Sfh-I crystal structure (PDB 7BJ9) was used as the receptor protein with hydrogen atoms added by AutoDockTools4.50 Structures of compounds D-anti-1b, L-syn-1b, D-anti-1a were optimized in vacuum at Hartree–Fock/6-31G* level using Gaussian 0952 and subsequently prepared in pdbqt format using AutoDockTools4.50 The grid map was set to 40 × 40 × 40 points with a grid spacing of 0.375 Å centered on the catalytic site. For each calculation, 100 different docking runs were performed, and the resulting 100 poses were clustered according to a ligand RMSD cut-off of 2 Å. The predicted binding free energy score (ΔG) and population were the criteria used to choose the correct ligand pose, further considering the crystallographic information of similar compounds.
Quantum Mechanics/Molecular Mechanics (QM/MM) optimizations
All starting structures were geometry optimized at the Quantum Mechanics/Molecular Mechanics (QM/MM) level. For hybrid QM/MM calculations, we used a semiempirical method PM653 to describe the QM region and the ff14SB force field54 to describe the MM region as implemented in Amber16.55, 56 Hydrogen atoms were added, and each protein was immersed in a truncated octahedral periodic box with a minimum solute-wall distance of 8 Å, filled with explicit TIP3P water molecules57, using the AMBER16 leap module55. The van der Waals radius, force constants and equilibrium distances, angles and dihedral of the studied inhibitors were taken from the gaff database and partial charges were RESP charges computed using the Hartree–Fock method and 6-31G* basis set.55, 58 To accommodate solvent molecules and possible clashes, an initial minimization at the molecular mechanic level of each complex structure was performed, followed by QM/MM geometry optimization. The QM region consisted of both Zn(II) ions plus the coordinated side chains of residues of the first coordination sphere and the inhibitor compound.
Supplementary Material
Acknowledgements
We thank Diamond Light Source for beamtime (proposals mx12342 and mx17212), and the staff of beamlines I03 and I04-1 for assistance with crystal testing and data collection. This work was supported by the National Institute of Allergy and Infectious Diseases of the National Institutes of Health (NIH) to R. A. B. under Award Numbers R01AI063517 and R01AI072219, to R. A. B., G. M., J. S. and A. J. V. under Award Number R01AI100560 and to B. S. under Award Numbers R01 AI130060 and AI117211. This study was also supported in part by funds and/or facilities provided by the Cleveland Department of Veterans Affairs, Award Number 1I01BX001974 to R. A. B. from the Biomedical Laboratory Research & Development Service of the VA Office of Research and Development, and the Geriatric Research Education and Clinical Center VISN 10. The content is solely the responsibility of the authors and does not necessarily represent the official views of the NIH or the Department of Veterans Affairs. V. M. and V. V. are recipients of a fellowship from Comisión Académica de Posgrado (CAP-Udelar). This work was supported by grant S2021INIC2019 from Comisión Sectorial de Investigación Científica (CSIC) to V. M. and G. M. and grant PICT-2016-1657 from ANPCyT to A. J. V., C. B. and D. M. M. and A. J. V. are staff members from CONICET. M. A. R. is recipient of a fellowship from CONICET.
Abbreviations
- ASU
asymmetric unit
- BLI
β-lactamase inhibitor
- BTZ
bisthiazolidine
- DBO
diazabicyclooctane
- MBL
metallo-β-lactamase
- MIC
minimum inhibitory concentration
- MMOZ
2-mercaptomethyl oxazolidine
- MMTZ
2-mercaptomethyl thiazolidine
- QM/MM
quantum mechanics/molecular mechanics
- SBL
serine-β-lactamase
Footnotes
Supporting Information
Supporting information is available
• Supporting Tables 1–4, Supporting Figures 1–7 and Supporting References.
References
- 1.Bush K; Bradford PA, beta-Lactams and beta-Lactamase Inhibitors: An Overview. Cold Spring Harb Perspect Med 2016, 6 (8). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Bush K, Past and Present Perspectives on beta-Lactamases. Antimicrob Agents Chemother 2018, 62 (10). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Ambler RP, The structure of beta-lactamases. Philos Trans R Soc LondB Biol Sci 1980, 289 (1036), 321–31. [DOI] [PubMed] [Google Scholar]
- 4.Bush K; Jacoby GA, Updated functional classification of beta-lactamases. Antimicrob Agents Chemother 2010, 54 (3), 969–76. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Tooke CL; Hinchliffe P; Bragginton EC; Colenso CK; Hirvonen VHA; Takebayashi Y; Spencer J, beta-Lactamases and beta-Lactamase Inhibitors in the 21st Century. J Mol Biol 2019, 431 (18), 3472–3500. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Rossi M-A; Martinez V; Hinchliffe P; Mojica MF; Castillo V; Moreno DM; Smith R; Spellberg B; Drusano GL; Banchio C; Bonomo RA; Spencer J; Vila AJ; Mahler G, 2-Mercaptomethyl-thiazolidines use conserved aromatic–S interactions to achieve broad-range inhibition of metallo-β-lactamases. Chemical Science 2021, 12 (8), 2898–2908. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Palzkill T, Metallo-beta-lactamase structure and function. Ann N Y Acad Sci 2013, 1277, 91–104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Mojica MF; Bonomo RA; Fast W, B1-Metallo-beta-Lactamases: Where Do We Stand? Curr Drug Targets 2016, 17 (9), 1029–50. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Garau G; Garcia-Saez I; Bebrone C; Anne C; Mercuri P; Galleni M; Frere JM; Dideberg O, Update of the standard numbering scheme for class B beta-lactamases. Antimicrob Agents Chemother 2004, 48 (7), 2347–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Saavedra MJ; Peixe L; Sousa JC; Henriques I; Alves A; Correia A, Sfh-I, a subclass B2 metallo-beta-lactamase from a Serratia fonticola environmental isolate. Antimicrob Agents Chemother 2003, 47 (7), 2330–3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Massidda O; Rossolini GM; Satta G, The Aeromonas hydrophila cphA gene: molecular heterogeneity among class B metallo-beta-lactamases. J Bacteriol 1991, 173 (15), 4611–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Poirel L; Palmieri M; Brilhante M; Masseron A; Perreten V; Nordmann P, PFM-Like Enzymes Are a Novel Family of Subclass B2 Metallo-beta-Lactamases from Pseudomonas synxantha Belonging to the Pseudomonas fluorescens Complex. Antimicrob Agents Chemother 2020, 64 (2). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Walsh TR; Gamblin S; Emery DC; MacGowan AP; Bennett PM, Enzyme kinetics and biochemical analysis of ImiS, the metallo-β-lactamase from Aeromonas sobria 163a. Journal of Antimicrobial Chemotherapy 1996, 37 (3), 423–431. [DOI] [PubMed] [Google Scholar]
- 14.Gonzalez LJ; Stival C; Puzzolo JL; Moreno DM; Vila AJ, Shaping Substrate Selectivity in a Broad-Spectrum Metallo-beta-Lactamase. Antimicrob Agents Chemother 2018, 62 (4). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Rasmussen BA; Bush K, Carbapenem-hydrolyzing beta-lactamases. Antimicrobial agents and chemotherapy 1997, 41 (2), 223–232. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Crowder MW; Spencer J; Vila AJ, Metallo-beta-lactamases: novel weaponry for antibiotic resistance in bacteria. Acc Chem Res 2006, 39 (10), 721–8. [DOI] [PubMed] [Google Scholar]
- 17.Mojica MF; Rutter JD; Taracila M; Abriata LA; Fouts DE; Papp-Wallace KM; Walsh TJ; LiPuma JJ; Vila AJ; Bonomo RA, Population Structure, Molecular Epidemiology, and beta-Lactamase Diversity among Stenotrophomonas maltophilia Isolates in the United States. mBio 2019, 10 (4). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Crossman LC; Gould VC; Dow JM; Vernikos GS; Okazaki A; Sebaihia M; Saunders D; Arrowsmith C; Carver T; Peters N; Adlem E; Kerhornou A; Lord A; Murphy L; Seeger K; Squares R; Rutter S; Quail MA; Rajandream MA; Harris D; Churcher C; Bentley SD; Parkhill J; Thomson NR; Avison MB, The complete genome, comparative and functional analysis of Stenotrophomonas maltophilia reveals an organism heavily shielded by drug resistance determinants. Genome Biol 2008, 9 (4), R74. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Bellais S; Aubert D; Naas T; Nordmann P, Molecular and biochemical heterogeneity of class B carbapenem-hydrolyzing beta-lactamases in Chryseobacterium meningosepticum. Antimicrob Agents Chemother 2000, 44 (7), 1878–86. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Moran-Barrio J; Lisa MN; Larrieux N; Drusin SI; Viale AM; Moreno DM; Buschiazzo A; Vila AJ, Crystal Structure of the Metallo-beta-Lactamase GOB in the Periplasmic Dizinc Form Reveals an Unusual Metal Site. Antimicrob Agents Chemother 2016, 60 (10), 6013–22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Thaller MC; Borgianni L; Di Lallo G; Chong Y; Lee K; Dajcs J; Stroman D; Rossolini GM, Metallo-beta-lactamase production by Pseudomonas otitidis: a species-related trait. Antimicrob Agents Chemother 2011, 55 (1), 118–23. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Wachino J; Yoshida H; Yamane K; Suzuki S; Matsui M; Yamagishi T; Tsutsui A; Konda T; Shibayama K; Arakawa Y, SMB-1, a novel subclass B3 metallo-beta-lactamase, associated with ISCR1 and a class 1 integron, from a carbapenem-resistant Serratia marcescens clinical isolate. Antimicrob Agents Chemother 2011, 55 (11), 5143–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Yong D; Toleman MA; Bell J; Ritchie B; Pratt R; Ryley H; Walsh TR, Genetic and biochemical characterization of an acquired subgroup B3 metallo-beta-lactamase gene, blaAIM-1, and its unique genetic context in Pseudomonas aeruginosa from Australia. Antimicrob Agents Chemother 2012, 56 (12), 6154–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Drawz SM; Bonomo RA, Three decades of beta-lactamase inhibitors. Clin Microbiol Rev 2010, 23 (1), 160–201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Papp-Wallace KM; Bonomo RA, New β-Lactamase Inhibitors in the Clinic. Infectious Disease Clinics of North America 2016, 30 (2), 441–464. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Bush K; Bradford PA, Interplay between beta-lactamases and new beta-lactamase inhibitors. Nat Rev Microbiol 2019, 17 (5), 295–306. [DOI] [PubMed] [Google Scholar]
- 27.Prosperi-Meys C; Llabres G; de Seny D; Soto RP; Valladares MH; Laraki N; Frere JM; Galleni M, Interaction between class B beta-lactamases and suicide substrates of active-site serine beta-lactamases. FEBS Lett 1999, 443 (2), 109–11. [DOI] [PubMed] [Google Scholar]
- 28.Brem J; Cain R; Cahill S; McDonough MA; Clifton IJ; Jimenez-Castellanos JC; Avison MB; Spencer J; Fishwick CW; Schofield CJ, Structural basis of metallo-beta-lactamase, serine-beta-lactamase and penicillin-binding protein inhibition by cyclic boronates. Nat Commun 2016, 7, 12406. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Calvopina K; Hinchliffe P; Brem J; Heesom KJ; Johnson S; Cain R; Lohans CT; Fishwick CWG; Schofield CJ; Spencer J; Avison MB, Structural/mechanistic insights into the efficacy of nonclassical beta-lactamase inhibitors against extensively drug resistant Stenotrophomonas maltophilia clinical isolates. Mol Microbiol 2017, 106 (3), 492–504. [DOI] [PubMed] [Google Scholar]
- 30.Krajnc A; Brem J; Hinchliffe P; Calvopina K; Panduwawala TD; Lang PA; Kamps J; Tyrrell JM; Widlake E; Saward BG; Walsh TR; Spencer J; Schofield CJ, Bicyclic Boronate VNRX-5133 Inhibits Metallo- and Serine-beta-Lactamases. J Med Chem 2019, 62 (18), 8544–8556. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Tooke CL; Hinchliffe P; Krajnc A; Mulholland AJ; Brem J; Schofield CJ; Spencer J, Cyclic boronates as versatile scaffolds for KPC-2 β-lactamase inhibition. RSC Medicinal Chemistry 2020, 11 (4), 491–496. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Hecker SJ; Reddy KR; Lomovskaya O; Griffith DC; Rubio-Aparicio D; Nelson K; Tsivkovski R; Sun D; Sabet M; Tarazi Z; Parkinson J; Totrov M; Boyer SH; Glinka TW; Pemberton OA; Chen Y; Dudley MN, Discovery of Cyclic Boronic Acid QPX7728, an Ultrabroad-Spectrum Inhibitor of Serine and Metallo-beta-lactamases. J Med Chem 2020, 63 (14), 7491–7507. [DOI] [PubMed] [Google Scholar]
- 33.Tsivkovski R; Totrov M; Lomovskaya O, Biochemical Characterization of QPX7728, a New Ultrabroad-Spectrum Beta-Lactamase Inhibitor of Serine and Metallo-Beta-Lactamases. Antimicrob Agents Chemother 2020, 64 (6). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Palacios AR; Rossi MA; Mahler GS; Vila AJ, Metallo-beta-Lactamase Inhibitors Inspired on Snapshots from the Catalytic Mechanism. Biomolecules 2020, 10 (6). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Lisa MN; Palacios AR; Aitha M; Gonzalez MM; Moreno DM; Crowder MW; Bonomo RA; Spencer J; Tierney DL; Llarrull LI; Vila AJ, A general reaction mechanism for carbapenem hydrolysis by mononuclear and binuclear metallo-beta-lactamases. Nat Commun 2017, 8 (1), 538. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Gonzalez MM; Kosmopoulou M; Mojica MF; Castillo V; Hinchliffe P; Pettinati,; Brem J; Schofield CJ; Mahler G; Bonomo RA; Llarrull LI; Spencer J; Vila AJ, Bisthiazolidines: A Substrate-Mimicking Scaffold as an Inhibitor of the NDM-1 Carbapenemase. ACS Infect Dis 2015, 1 (11), 544–54. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Hinchliffe P; Gonzalez MM; Mojica MF; Gonzalez JM; Castillo V; Saiz C; Kosmopoulou M; Tooke CL; Llarrull LI; Mahler G; Bonomo RA; Vila AJ; Spencer J, Cross-class metallo-beta-lactamase inhibition by bisthiazolidines reveals multiple binding modes. Proc Natl Acad Sci U S A 2016, 113 (26), E3745–54. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Garau G; Bebrone C; Anne C; Galleni M; Frere JM; Dideberg O, A metallo-beta-lactamase enzyme in action: crystal structures of the monozinc carbapenemase CphA and its complex with biapenem. J Mol Biol 2005, 345 (4), 785–95. [DOI] [PubMed] [Google Scholar]
- 39.Kim Y; Maltseva N; Wilamowski M; Tesar C; Endres M; Joachimiak A, Structural and biochemical analysis of the metallo-β-lactamase L1 from emerging pathogen Stenotrophomonas maltophilia revealed the subtle but distinct di-metal scaffold for catalytic activity. Protein Sci 2020, 29 (3), 723–743. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Feng L; Yang KW; Zhou LS; Xiao JM; Yang X; Zhai L; Zhang YL; Crowder MW, N-heterocyclic dicarboxylic acids: broad-spectrum inhibitors of metallo-beta-lactamases with co-antibacterial effect against antibiotic-resistant bacteria. Bioorg Med Chem Lett 2012, 22 (16), 5185–9. [DOI] [PubMed] [Google Scholar]
- 41.CLSI, Performance Standards for Antimicrobial Susceptibility Testing, Wayne, PA: Clinical and Laboratory Standards Institute, 2019, vol. 29th ed. [Google Scholar]
- 42.Fonseca F; Arthur CJ; Bromley EH; Samyn B; Moerman P; Saavedra MJ; Correia A; Spencer J, Biochemical characterization of Sfh-I, a subclass B2 metallo-beta-lactamase from Serratia fonticola UTAD54. Antimicrob Agents Chemother 2011, 55 (11), 5392–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Winter G; Waterman DG; Parkhurst JM; Brewster AS; Gildea RJ; Gerstel M; Fuentes-Montero L; Vollmar M; Michels-Clark T; Young ID; Sauter NK; Evans G, DIALS: implementation and evaluation of a new integration package. Acta crystallographica. Section D, Structural biology 2018, 74 (Pt 2), 85–97. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Winn MD; Ballard CC; Cowtan KD; Dodson EJ; Emsley P; Evans PR; Keegan RM; Krissinel EB; Leslie AG; McCoy A; McNicholas SJ; Murshudov GN; Pannu NS; Potterton EA; Powell HR; Read RJ; Vagin A; Wilson KS, Overview of the CCP4 suite and current developments. Acta crystallographica. Section D, Biological crystallography 2011, 67 (Pt 4), 235–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.McCoy AJ; Grosse-Kunstleve RW; Adams PD; Winn MD; Storoni LC; Read RJ, Phaser crystallographic software. Journal of applied crystallography 2007, 40 (Pt 4), 658–674. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Emsley P; Cowtan K, Coot: model-building tools for molecular graphics. Acta crystallographica. Section D, Biological crystallography 2004, 60 (Pt 12 Pt 1), 2126–32. [DOI] [PubMed] [Google Scholar]
- 47.Adams PD; Afonine PV; Bunkoczi G; Chen VB; Davis IW; Echols N; Headd JJ; Hung LW; Kapral GJ; Grosse-Kunstleve RW; McCoy AJ; Moriarty NW; Oeffner R; Read RJ; Richardson DC; Richardson JS; Terwilliger TC; Zwart PH, PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta crystallographica. Section D, Biological crystallography 2010, 66 (Pt 2), 213–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Moriarty NW; Grosse-Kunstleve RW; Adams PD, electronic Ligand Builder and Optimization Workbench (eLBOW): a tool for ligand coordinate and restraint generation. Acta crystallographica. Section D, Biological crystallography 2009, 65 (Pt 10), 1074–80. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Schrödinger L, The PyMOL Molecular Graphics System. Vol. Version 1.8.
- 50.Morris GM; Huey R; Lindstrom W; Sanner MF; Belew RK; Goodsell DS; Olson AJ, AutoDock4 and AutoDockTools4: Automated docking with selective receptor flexibility. J Comput Chem 2009, 30 (16), 2785–91. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Arcon JP; Modenutti CP; Avendano D; Lopez ED; Defelipe LA; Ambrosio F; Turjanski AG; Forli S; Marti MA, AutoDock Bias: improving binding mode prediction and virtual screening using known protein-ligand interactions. Bioinformatics 2019, 35 (19), 3836–3838. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Frisch MJ; Trucks GW; Schlegel HB; Scuseria GE; Robb MA; Cheeseman JR; Scalmani G; Barone V; Petersson GA; Nakatsuji H; Li X; Caricato M; Marenich AV; Bloino J; Janesko BG; Gomperts R; Mennucci B; Hratchian HP; Ortiz JV; Izmaylov AF; Sonnenberg JL; Williams; Ding F; Lipparini F; Egidi F; Goings J; Peng.; Petrone A; Henderson T; Ranasinghe D; Zakrzewski VG; Gao J; Rega N; Zheng G; Liang W; Hada M; Ehara M; Toyota K; Fukuda R; Hasegawa J; Ishida M; Nakajima T; Honda Y; Kitao O; Nakai H; Vreven T; Throssell K; Montgomery JA Jr.; Peralta JE; Ogliaro F; Bearpark MJ; Heyd JJ; Brothers EN; Kudin KN; Staroverov VN; Keith TA; Kobayashi R; Normand J; Raghavachari K; Rendell AP; Burant JC; Iyengar SS; Tomasi J; Cossi M; Millam JM; Klene M; Adamo C; Cammi R; Ochterski JW; Martin RL; Morokuma K; Farkas O; Foresman JB; Fox DJ Gaussian 16 Rev. C.01, Wallingford, CT, 2016. [Google Scholar]
- 53.Stewart JJ, Optimization of parameters for semiempirical methods V: modification of NDDO approximations and application to 70 elements. J Mol Model 2007, 13 (12), 1173–213. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Maier JA; Martinez C; Kasavajhala K; Wickstrom L; Hauser KE; Simmerling C, ff14SB: Improving the Accuracy of Protein Side Chain and Backbone Parameters from ff99SB. J Chem Theory Comput 2015, 11 (8), 3696–713. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Case DA; Betz RM; Cerutti DS; Cheatham TE III; Darden TA; Duke RE; Giese TJ; Gohlke H; Goetz AW; Homeyer N; Izadi S; Janowski P; Kaus J; Kovalenko A; Lee TS; LeGrand S; Li P; Lin C; Luchko T; Luo R; Madej B; Mermelstein D; Merz KM; Monard G; Nguyen H; Nguyen HT; Omelyan I; Onufriev A; Roe DR; Roitberg AE; Sagui C; Simmerling CL; Botello-Smith WM; Swails J; Walker RC; Wang J; Wolf RM; Wu X; Xiao L; Kollman PA, 2016, AMBER 2016, University of California, San Francisco. [Google Scholar]
- 56.de MSG; Walker RC; Elstner M; Case DA; Roitberg AE, Implementation of the SCC-DFTB method for hybrid QM/MM simulations within the amber molecular dynamics package. JPhys Chem A 2007, 111 (26), 5655–64. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Jorgensen WL; Chandrasekhar J; Madura JD; Impey RW; Klein ML, Comparison of simple potential functions for simulating liquid water. Journal of Chemical Physics 1983, 79, 926. [Google Scholar]
- 58.Bayly CI; Cieplak P; Cornell W; Kollman PA, A well-behaved electrostatic potential based method using charge restraints for deriving atomic charges: the RESP model. The Journal of Physical Chemistry 1993, 97 (40), 10269–10280. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.





