Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2023 Apr 24;63(13):4056–4069. doi: 10.1021/acs.jcim.2c01156

Unraveling the Reaction Mechanism of Russell’s Viper Venom Factor X Activator: A Paradigm for the Reactivity of Zinc Metalloproteinases?

Juliana Castro-Amorim , Ana Oliveira , Ashis K Mukherjee , Maria J Ramos , Pedro A Fernandes †,*
PMCID: PMC10336966  PMID: 37092784

Abstract

graphic file with name ci2c01156_0009.jpg

Snake venom metalloproteinases (SVMPs) are important drug targets against snakebite envenoming, the neglected tropical disease with the highest mortality worldwide. Here, we focus on Russell’s viper (Daboia russelii), one of the “big four” snakes of the Indian subcontinent that, together, are responsible for ca. 50,000 fatalities annually. The “Russell’s viper venom factor X activator” (RVV-X), a highly toxic metalloproteinase, activates the blood coagulation factor X (FX), leading to the prey’s abnormal blood clotting and death. Given its tremendous public health impact, the WHO recognized an urgent need to develop efficient, heat-stable, and affordable-for-all small-molecule inhibitors, for which a deep understanding of the mechanisms of action of snake’s principal toxins is fundamental. In this study, we determine the catalytic mechanism of RVV-X by using a density functional theory/molecular mechanics (DFT:MM) methodology to calculate its free energy profile. The results showed that the catalytic process takes place via two steps. The first step involves a nucleophilic attack by an in situ generated hydroxide ion on the substrate carbonyl, yielding an activation barrier of 17.7 kcal·mol–1, while the second step corresponds to protonation of the peptide nitrogen and peptide bond cleavage with an energy barrier of 23.1 kcal·mol–1. Our study shows a unique role played by Zn2+ in catalysis by lowering the pKa of the Zn2+-bound water molecule, enough to permit the swift formation of the hydroxide nucleophile through barrierless deprotonation by the formally much less basic Glu140. Without the Zn2+ cofactor, this step would be rate-limiting.

Introduction

Snakebite Epidemiology

Snakebite envenoming represents a public health hazard affecting tropical countries’ rural populations.1 Tropical and subtropical regions of the world, like Asia, sub-Saharan Africa, Latin America, and parts of Oceania, suffer the most significant impact.24 Although the actual number of people bitten by snakes is not known accurately,5 estimates suggest that, per year, approximately 5.4 million snakebites occur, with up to 2.7 million cases (one-half) leading to snakebite envenoming, which causes 81,000–138,000 deaths and around 400,000 amputations and permanent sequelae.1,3,6 Despite being globally distributed, snakebite envenoming in India is the highest compared to any other country,4 accounting for nearly half of global snakebite deaths.5,7

In India, most of the snakebites are inflicted by the “big four” snakes, which include the saw-scaled viper (Echis carinatus), the spectacled cobra (Naja naja), the common krait (Bungarus caeruleus), and the Russell’s viper (Daboia russelii).4,5,8,9 The latter, which will be further discussed, is one of the major venomous snakes responsible for the high rates of morbidities and mortality in the Indian subcontinent and Southeast Asia,7,1013 representing a severe medical threat to the local population. Unfortunately, access to health services is limited in the remote rural areas where most snakebites occur.6,14 The only reasonably efficient therapy is antibody-based antivenom,4,6,14,15 which is very expensive and needs a cold chain of transportation and storage, and inpatient administration as it is frequently associated with anaphylactic reactions.14,15 Therefore, the therapy is of meager availability to those in desperate need.4,1416

Following several requests, in 2017, the World Health Organization (WHO) recognized snakebite envenoming as a neglected tropical disease (NTD),3,4,14,17 presently the one leading to the most significant number of annual deaths. Therefore, in May 2019, the WHO launched a work plan outlining the goal of halving mortality and disabilities caused by envenoming by 2030.3,14,18 Among other measures, this action plan promotes the research and development of inexpensive, heat-stable, affordable-for-all small-molecule inhibitors that can be stored in local villages and administered outside a hospital settlement.3,18 Some chelators have offered great potential for snakebite treatment by inhibiting snake venom metalloproteinase (SVMP) toxins, including the peptidomimetic hydroxamate inhibitor (WR2), ilomastat, marimastat, and batimastat. Although initially designed for inhibiting human extracellular matrix metalloproteinases (MMPs), these were used for repurposing based on the high degree of sequence identity shared by MMPs and SVMPs. However, all failed in the clinical trials due to toxicity problems brought on by off-target effects.19,20

Russell’s Viper and Its Venom

This paper focuses on one of the leading actors of the problem, Russell’s viper (RV), a large and beautiful snake that is responsible for most snakebites that occur in the vicinity of the agricultural lands and villages4,9,12,21,22 where its primary dietary preference, rats and mice, grow.5,8 The RV is a severe issue for rice cultivators due to its abundance in rice paddies and for other farmers that work in its habitat and therefore are its primary victims,4,5,8 as the snake usually reacts to human presence by concealing itself (Daboia means “the lurker” in Hindi) but bites if the farmer incidentally steps on it.4

Snake venoms have an essentially proteinaceous nature.5 They are composed of up to 200 different proteins, and Russell’s viper venom (RVV) is no exception. The enzymatic toxins phospholipases A2, metalloproteinases, and serine proteases constitute most of Russell’s viper venom (Figure 1; see ref (1) for a review on snake venom composition). These enzymes are responsible for several local and systemic clinical manifestations, including edema, tissue necrosis, hemorrhage, acute renal failure, and hypotension.8,11,14,23 Besides those, some events of neurotoxicity and myotoxicity post RV envenomation have been documented in Sri Lankan and South Indian populations11 and Myanmar, respectively.8,10,24 Envenomation further reduces male sex hormones; the complete loss of libido and secondary sexual characteristics (for example, loss of facial hair) and even reduction of the size of the sexual organs are not uncommon. Recovery is rarely complete.25

Figure 1.

Figure 1

Average proteomic profile of Russell’s Viper venom. The principal protein families are the l-amino acid oxidase (LAAO), natriuretic peptides (NP), serine protease (SVSP), phospholipase A2 (PLA2), snake C-type lectins (CTL/Snaclec), metalloproteinases (SVMP), and disintegrin (DIS).1

Structure and Function of Metalloproteinases

SVMPs are zinc-dependent metalloproteinases with a modular structure that might include in the most complex isoforms (such as in RVV-X) a metalloproteinase, a disintegrin-like, a cysteine-rich, and a C-type lectin-like domain.2630 The metalloproteinase domains are characterized by a strictly conserved catalytic consensus sequence HEXXHXXGXXH, whose three histidine residues coordinate to the zinc ion through its Nε2 atoms.19,27,30,31 A water molecule completes the tetrahedral coordination of the cofactor and is hydrogen-bonded to Glu140.3234

RVV-X

In some geographic regions, RVV exhibits potent anticoagulant effects, whereas, in others, it exhibits potent pro-coagulant activity.8,10,12,35 The latter is mainly due to a specific toxin of its venom, RVV-X,8,10,36 which is the most potent venom coagulation activator known that causes fatal envenomation.36 It is a non-hemorrhagic class III (subclass d) snake venom metalloproteinase10 (Supporting Information, Figure S1) that interferes in the victim’s hemostasis by triggering uncontrolled coagulation through activation of the coagulation factor X (FX).8,36 The activation is triggered by the RVV-X catalytic cleavage of a specific peptide bond between Arg152 and Ile153 (human numbering) of FX, resulting in the release of a 52-residue peptide named as “activation peptide” (Figure 2).8,21,34,37 The cleavage leads to the conversion of the zymogen FX into its active form FXα and, consequently, to the coagulation cascade’s activation.21,34,3638 Thus, just like for other vipers, SVMPs play a determinant role in the pathogenesis of RV bites.39

Figure 2.

Figure 2

Schematic model of the FX conversion to its active form, FXa, by the RVV-X. (I) The FX activation peptide binds to the active site of the RVV-X; (II) the Arg152–Ile153 scissile bond is cleaved by the RVV-X, resulting in the release of the 52-residue activation peptide; (III) consequently, the common coagulation pathway is activated, leading to the formation of blood clots. The stereo structure of the RVV-X is colored by chain (HC: dark blue, LC1: green, LC2: cyan) and shown in VDW representation. The orange circle represents the zinc ion. Created with BioRender.com.

The structure of RVV-X has already been solved by X-ray crystallography. Takeda and co-workers determined its complete structure in 2007 with a resolution of 2.91 Å.40 RVV-X is a 93 kDa heterotrimeric glycoprotein composed of three disulfide-linked glycosylated polypeptide chains, one heavy chain (HC), and two light chains (LC1 and LC2). The backbone structure of the HC (α-chain, 57.6 kDa) follows the characteristics of class III SVMPs with its three metalloproteinase, disintegrin-like, and Cys-rich domains. These domains are organized in a C-shape configuration where the metalloproteinase domain interacts with the Cys-rich domain. The latter interacts with the LC1 via a disulfide bond.8,27,36,40 The LC subunit (β- and γ-chains, 19.4 and 16.4 kDa) forms a domain-swapped dimer. The light chains share significant sequence identity with snake venom C-type lectin-like proteins. The dimeric interface formed by the two LC is a concave structure that may function as an exosite that confers both affinity and binding specificity for the Gla domain of FX in the presence of Ca2+ ions.36,39,41

Due to its high specificity, RVV-X is a remarkable biotechnological tool widely used in coagulation research and medical diagnostics.5,31,42 For example, the Stypven time is determined by a clotting assay that measures the FX to FXa conversion by RVV-X and subsequent prothrombin activation that initiates clot formation.4345 In particular, a long Stypven time is a diagnostic of FX deficiency, also called Stuart–Prower factor deficiency.44

Catalytic Mechanism of SVMP

The reaction mechanism of SVMPs has not been elucidated so far.26,46 Typically, snake venom metalloproteinase X-ray structures show the Zn2+ cofactor coordinated by the three His residues of the catalytic motif and sometimes by a water molecule33,4750 or a doubly coordinated peptidomimetic inhibitor40,5154 replacing both the water molecule and the substrate. Computational studies of the reaction mechanism of the homologous human matrix metalloproteinases (MMPs) carried out by Pelmenschikov and Siegbahn,55 Díaz and Suárez,56 and Vasilevskaya et al.57 shed some light on the possible reaction mechanism of SVMP. The former group studied the hydrolysis of N-methylacetamide catalyzed by the matrix metalloproteinase-3 using a two-layered hybrid of our own n-layered Integrated molecular orbital and molecular mechanics (ONIOM) method, while the last two groups studied the hydrolysis of peptides by the matrix metalloproteinase-2. These studies point to a mechanism for the human MMPs named the “water-promoted pathway”5860 shown in Scheme 1, providing a starting point for the understanding of the reaction mechanism of the SVMPs.

Scheme 1. “Water-Promoted Pathway,” A Plausible Catalytic Mechanism for the SVMPs.

Scheme 1

(1) The conserved Glu acts as a base and deprotonates the Zn2+-bound water molecule. The generated hydroxide ion attacks the Zn2+-bound peptide carbonyl, generating a Zn2+-stabilized gem-diolate; the oxygen from the Zn2+-bound substrate carbonyl becomes negatively charged; collapse of TS1 into INT1 and hydrogen-bond network rearrangement of the conserved Glu; (2) the neutral Glu protonates the amine nitrogen of the scissile substrate bond, breaking the peptide bond; the proton from the hydroxyl group of the peptide is also shuttled to the amine product by the Glu residue or, alternatively, the proton comes from the bulk solvent after the product release; releasing the first fragment of the products, an amine-terminated peptide; (3) the gem-diolate unbinds the Zn2+ ion, through replacement by a water molecule, generating the second fragment of the product, a carboxylic-acid-terminated peptide and regenerating the active site. Alternatively, the water deprotonation can precede the hydrolysis, with the active site ground state consisting of a hydroxide-bound Zn2+ ion and a neutral Glu, which is plausible given the lowering of the water pKa that binding the metal provides. In this scenario, the peptide bond’s attack and the protonation of the amine nitrogen can occur in a concerted fashion or sequentially.

Mutations in the highly conserved catalytic residues affect one or more stages of the catalytic process.32,33 For example, in 1994, Crabbe et al.61 concluded through mutation experiments that the glutamate residue plays an essential role in the MMP-2 reaction, as its substitution by an alanine led to a marked decrease in the enzymatic activity.

Nevertheless, metalloproteinases are versatile in terms of their catalytic mechanisms. Besides the water-promoted pathway, there is evidence of an alternative mechanism in other related zinc proteases (thermolysin62 and carboxypeptidase-A58,63), named “anhydride pathway”, also called “glutamate-mediated nucleophilic pathway”. In this pathway, the sidechain carboxylate of the conserved active-site Glu residue (Glu140 in RVV-X numbering) directly attacks the scissile carbonyl carbon, producing an anhydride acyl-enzyme intermediate. Then, this intermediate is hydrolyzed by an incoming water molecule to release the products (Supporting Information, Scheme S1).58,62 However, X-ray diffraction and 18O isotopic labeling58,64 indicated that the Glu-mediated nucleophilic attack might only be viable for the hydrolysis of ester substrates but not for peptides.59 This is thought to be due to a misalignment of the reactant residues and the existence of an H-bond between the Glu and the substrate amide, which reduces the Glu nucleophilicity.58 Thus, the water-promoted mechanism for peptide substrates should be structurally and kinetically favored over the anhydride mechanism in RVV-X.58,59 Nevertheless, we have analyzed both mechanistic scenarios in this work.

In summary, the atomic-level details of the SVMP mechanism are still controversial, mainly because earlier studies relied on different types of metalloproteinases, and the reactions were studied with different substrates.5559,62 Additionally, no studies have been made on snake venom orthologues. Therefore, a quantitative description of the free energy profiles and detailed atomic-level mechanisms is needed for the SVMPs, particularly given its central role in the snakebite envenoming pathophysiology. These facts together demand an urgent need for a deep understanding of the chemical transformations in the RVV-X active site. This is essential for designing new and more efficient tools, that is, readily accessible and affordable-for-all small-molecule inhibitors for early treatment of snakebite envenoming.

In this study, computational DFT:MM calculations were used to determine the reaction mechanism of RVV-X. We present the RVV-X proteolytic mechanism against FX at the DFT:MM level of theory, with atomic detail, including the geometry of all minima and transition states and the reaction free energy profile. In addition, the crucial role played by the Zn2+ cofactor in catalysis has also been elucidated.

As referred before, the metalloproteinases are highly conserved proteolytic enzymes. The catalytic sites of the SVMP members are structurally very similar, and all share the consensus signature sequence. SVMP structures have a typical topology, consisting of six α-helices and five stranded β-sheets in the M domain where the zinc ion is localized. Besides, sequence comparison between the M domains of SVMPs reveals quite high identity values of around 66%.65 This chemical and stereochemical match, particularly at the active site region, allows us to hypothesize that many of the members of this family of enzymes share a common catalytic mechanism. Therefore, our results concerning the intricacies of the RVV-X mechanism might be extrapolated to many other members of this vast family.

Computational Methods

Homology Modeling

Structural information on snake venom proteins (toxins) is limited and, therefore the structure of our target, RVV-X from D. russelii, is not available in the protein data bank. However, the structure of the Eastern RV (Daboia siamensis) RVV-X has been determined (PDB ID: 2E3X).40 The taxonomic proximity between the two vipers is so significant that some researchers still consider the latter a subspecies of the former. Consistently, the sequence identity between the enzymes of both species is nearly complete: the Eastern RV RVV-X structure showed 97.7% of sequence identity to RV HC and 100 and 91.0% to LC1 and LC2, respectively. This structure was co-crystallized with a clinical inhibitor, Ilomastat, with a resolution of 2.91 Å. Thus, a three-dimensional (3D) model of D. russelii RVV-X was trivially built, resorting to a homology modeling technique using the SWISS-MODEL web server,66,67 using the crystallographic structure from D. Siamensis RVV-X as a template. A list with the mutations and respective positions required to modify the RVV-X primary structure of D. siamensis to D. russelii is provided in the Supporting Information.

The superimposition of the backbone of both the Eastern’s RV template and RV target revealed a remarkable similarity, with an RMSD difference of 0.1 Å, which was expected due to the very high sequence identity. The HC was organized in four long α-helices, one of which contains the active site and a four-stranded β-sheet without the antiparallel β-sheet (β4). The catalytic zinc ion made a tetrahedral coordination sphere with His139, His143, His149, and the added water molecule (SI, Figure S2). The metalloproteinase domain of this structure was solvated and used for the calculations. The protonation states of titratable residues were predicted by the H++ web server (http://biophysics.cs.vt.edu/H++)68,69 and are presented in Supporting Table S1. In summary, all residues were predicted to be in their typical protonation states except for Glu90 and Glu140, which were predicted to be neutral at physiological pH. The first is exposed to the solvent and does not establish strong interactions or salt bridges that justify the neutral form; thus, we decided to model it in the anionic form. The second, Glu140, was interestingly predicted to have a pKa slightly above 7.4, emphasizing its considerable basicity that is materialized through the barrierless and spontaneous abstraction of the nucleophilic water just at the beginning of the reaction, as will be seen below.

Parametrization of the Metalloproteinase’s Metal Center

The python-based Metal Center Parameter Builder (MCPB.py)70 module of the Assisted Model Building with Energy Refinement (AMBER) Tools18 package was used to parametrize the metalloproteinase’s metal center (details in the SI).

To carry out classical molecular dynamics simulations of the RVV-X, the modeled protein was solvated in an octahedral box of TIP3P water molecules71 so that the box boundaries were at least 15 Å away from any protein atom. Finally, eight Cl counterions were added to achieve electroneutrality on the modeled system. Missing hydrogen atoms were added by the tLeap software from the AMBER18 package. The geometry of the entire system was minimized through four sequential steps using the AMBER18 package and the GAFF72 and ff14SB73 force fields (details in the SI). Considering that the chemical reaction occurs in the metalloproteinase domain, which is located in the HC, the studies focused only on its HC, i.e., the LC (snaclecs) were erased for the subsequent simulations.

Protein–Peptide Docking

As the activation peptide structure is undetermined and disordered,74 it had to be modeled before the docking calculation. Due to its disordered nature, only an eight-residue peptide was used to model the cleavage region of the substrate, which is the portion that fits into the RVV-X binding site. To this end, the last 4 residues of the activation peptide (Asn-Leu-Thr-Arg) and the first 4 residues of the FXa serine protease N-terminal domain (Ile-Val-Gly-Gly) were modeled with tLeap. The peptide minimizations were carried out according to the same procedure used for the RVV-X model. With the substrate and target minimized, the next step consisted of protein–peptide docking. The latter was performed with the HPepDock (Hierarchical Flexible Peptide Docking)75 online server. A set of restraints was defined: the receptor residues (His143, Zn2+, Glu140) had to interact with the ligand residues (Thr3, Arg4, and Ile5) with a minimum and maximum distance of 2 and 6 Å. Finally, the most favorable solution was selected and further minimized (details in the SI). The final structure formed the obligatory interactions for the reaction: The scissile bond carbonyl coordinated to the Zn2+ cofactor, and the scissile peptide amine lay close to the Glu140 sidechain to allow for the amine protonation by Glu140. These observations further confirm that the position of the substrate is correct. In addition, the C-terminal portion of the scissile peptide segment (Val-Gly-Gly-Ile) adopted the same conformation as the peptide-like inhibitor GM6 co-crystallized with D. siamensis RVV-X40 (SI, Figure S3). A water molecule was introduced to complete the tetrahedral shell of the Zn2+ cofactor. It lay in the same position as the N3 atom of the GM6 inhibitor. Altogether, these observations showed that a correct geometry for the Michaelis complex was achieved.

The substrate was subsequently parameterized using AMBER’s ff14SB force field. We relaxed the modeled structure with a 170 ns molecular dynamics (MD) simulation (details in the SI). The results have shown that the modeled and optimized structure did not deviate from the ones in the MD ensemble, apart from the sidechain rotamer of Arg219, which goes deeper into the pocket S1 in the MD simulation. In this scenario, we decided to use the modeled and optimized structure to keep it as close as possible to the X-ray structure. The latter represents an average over all proteins in the crystal and thus is the most important and informative of all structures. Our experience is that the x-ray structure can provide results similar to ensemble averages because it is an ensemble average itself.76,77

QM/MM Model

The minimized structure of the protein-substrate complex was loaded on the molUP plugin78 of the VMD program package, which provides a full-featured graphical user interface (GUI) to the software Gaussian.79 Only the metalloproteinase domain and the substrate were included in the model to reduce the computational burden. To simulate the aqueous environment, 3489 water molecules and one chloride ion were added, corresponding to a solvent shell thickness of 6 Å around the protein. The whole system consisted of 6954 atoms. Two ONIOM layers were defined within the “our own n-layered Integrated molecular Orbital and Molecular mechanics” (ONIOM) formalism: the quantum mechanics (QM) layer, described with density functional theory (DFT), and the molecular mechanics (MM) layer, described by the AMBER force field. The QM region included all of the atoms that undergo bond formation or breaking processes and those which made first-shell interactions with the reacting atoms or whose hybridization changed during the reaction:8082 the Zn2+ ion, the catalytic water molecule, the side chains of His139, His143, and His149 until the β-carbon and the Glu140 residue, the substrate’s Arg219 backbone up to the β -carbon, the Ile220- α- and β-carbon plus the amine part, and the Thr218 α-carbon. This layer contained 73 atoms and a net charge of +1 (Figure 3). As typical with metalloenzymes, a relatively small QM layer encompassed all atoms relevant for the reaction. The QM part was described at the DFT level with the B3LYP hybrid functional,83 in combination with the 6-31G(d) basis set84 for geometry optimizations (larger basis sets were used for single-point energy calculations). The MM part included the rest of the protein and the solvent. It was treated with the AMBER ff14SB force field. Incomplete valences at the boundary between the two layers were completed with link hydrogen atoms.80,85,86

Figure 3.

Figure 3

QM/MM model used in the calculations. (Left) The metalloproteinase domain (surface and ribbon), the solvent (CPK), and the cofactor (yellow sphere). (Right) Close-up view of the QM layer. Carbon is shown in green, nitrogen in blue, oxygen in red, hydrogen in white, and zinc in copper.

Determination of the Stationary Points

Linear transit scans were initially performed to study the energy profile of the reaction’s steps and provide good guesses for the unrestrained optimization of minima and transition states. These were subsequently fully optimized (apart from the frozen atoms), their vibrational frequencies were calculated to confirm that the structures were indeed minima or transition states (i.e., having none or a single imaginary frequency, respectively), and IRC calculations87 were performed to confirm that the transition states were indeed the ones of interest. Furthermore, Grimme’s D3(BJ) dispersion corrections were added. In addition, the zero-point (ZPE), thermal energy, and rigid rotor/harmonic oscillator entropy were calculated at 298.15 K and 1.0 bar. Finally, single-point energy calculations at the ONIOM (B3LYP/6-311++G(2d,2p):AMBER), ONIOM(B3LYP/6-311++G(3df,2pd):AMBER), ONIOM(M06/6-311++G(3df,2pd):AMBER) and ONIOM(M06L/6-311++G(3df,2pd):AMBER) levels of theory were carried out.

Additionally, a set of different initial structures was retrieved from a 100 ns MD simulation (details in the SI) in the NPT ensemble with the same setup as the ones explained above, to evaluate the influence of the starting conformation on the free energy barrier of the rate-limiting step. Ten distinct structures that possessed suitable interatomic distances for the catalysis to occur were then selected. Following optimization of each retrieved structure, potential energy surface scans were performed using the B3LYP/6-31G(d) level of theory. Zero-point (ZPE), thermal energy, and rigid rotor/harmonic oscillator entropy were calculated and single-point energy calculations at the B3LYP/6-311++G(3df,2pd) were used to determine the energetic barrier for each conformation. Grimme’s D3(BJ) dispersion corrections were further added.

Images were created with PyMol and BioRender.com.

Results and Discussion

Michaelis Complex

In the structure of the protein–peptide complex obtained from the docking of the eight-segment peptide in the RVV-X binding site (SI, Figure S4A,B), the metal ion has a distorted tetrahedral coordination sphere, with the three histidine residues of the catalytic motif (SI, Figure S4A), the catalytic water, and the substrate’s carbonyl oxygen atom bound to it, with distances ranging from 1.89 to 2.62 Å. The zinc ion is well coordinated, and the backbone of the scissile residues—Arg219 and Ile220—is correctly oriented for the cleavage process. The Zn2+-coordinated water molecule is hydrogen-bonded to Glu140 (1.5 Å) and close to the Arg219 carbonyl carbon atom. In summary, the Michaelis complex structure is suitable for the reaction to proceed through the “water-promoted pathway” mechanism (Figure 4).

Figure 4.

Figure 4

Michaelis complex as predicted by restrained docking and modeling, in which the carbonyl oxygen (Opep) of the scissile peptide bond (Arg219-Ile220), the catalytic water (Owat), and the three histidine residues (His139, His143, His149) are coordinated to the Zn ion. Distances between the catalytic residues are presented. The yellow shadow highlights the peptide bond being broken.

Moreover, with a distance of 2.85 Å, the zinc-bound water nucleophile is ideally positioned to attack the scissile peptide bond’s carbonyl carbon (Cpep). This rearrangement is sustained by two hydrogen bonds given by the Glu140 sidechain (H1wat-Oε2Glu) and the Thr218 (H2wat-OThr) of the peptide substrate, with distances of 1.53 and 1.66 Å. During the subsequent QM/MM geometry optimization, the metal ion maintained its five-coordination environment. However, Opep moved slightly away from the Zn2+ ion (2.65 Å), and the water molecule got closer to the Zn2+ ion (Zn2+–Owat distance of 1.90 Å). One of the water protons was immediately transferred to Glu140 (Oε2Glu-H1wat bond length of 1.07 Å), with Owat keeping a hydrogen bond with the transferred proton (1.44 Å). The nucleophilic water molecule is very polarized due to the interactions with the negatively charged carboxyl group of Glu140 base and the acidic Zn2+ ion. In these conditions, the pKa of the water molecule drops considerably,88,89 and in this case, it dropped enough to allow for the barrierless deprotonation by Glu140. Thus, the nucleophile for the water-mediated pathway, i.e., the Zn2+-bound hydroxide ion was swiftly generated.

On the other hand, the RVV-X structure indicates that the anhydride catalytic pathway is unlikely because the glutamate residue and the substrate’s carbonyl group are not in a favorable position for the reaction to take place with low barriers. The distance between the Glu140 sidechain oxygen atom closest to the carbonyl carbon of the scissile bond is 4.28 Å. Given the difficulty in dragging the enzyme backbone to approach the tightly coordinated substrate, the free energy barriers for navigating through this mechanism might be significant. Moreover, the nucleophilicity of Glu140 is severely compromised by the water molecule’s spontaneous and barrierless protonation.

In summary, an enzyme–substrate complex (REACT) is formed in which the hydroxide ion is ready to attack the scissile peptide bond (Figure 5A). This Michaelis complex differs from the proposals for the human matrix metalloproteinases; it is much more reactive and prone to carry out the water-promoted pathway.

Figure 5.

Figure 5

Fully optimized stationary states along the catalytic cycle of RV RVV-X. Only the QM layer is shown; the MM layer has been deleted for clarity. (A) Enzyme–substrate complex, (REACT); (B) first transition state, TS1, corresponding to the nucleophilic attack of the Zn2+-bound hydroxide ion to the peptide carbonyl carbon; (C) first intermediate (INT1), which corresponds to a Zn2+-bound doubly coordinated gem-diolate; (D) second, rate-limiting transition state (TS2) resulting from a proton transfer from Glu140 to the amine nitrogen of the scissile bond and partial cleavage of the Cpep-Npep peptide bond; (E) second intermediate, INT2, with the peptide amine nitrogen protonated and peptide bond severely weakened but not broken; (F) product, PROD, with the peptide bond fully broken and the product formed. To complete the cycle, a bulk solvent water molecule enters the binding site, coordinates to the Zn2+ ion, and displaces the C-terminal fraction of the product from the cofactor and active site. The most relevant catalytic distances are presented in Ångström.

Revealing the Catalytic Mechanism of RVV-X

Glutamate-Assisted Nucleophilic Pathway

Even though the active site-substrate architecture naturally favors the water-promoted pathway, we performed a linear transit scan along the distance between the Glu140 Oε2 atom and the Cpep. As expected, the Glu140 nucleophilic attack led to a steep rise in energy up to prohibitive values. The Glu140 has an unproductive position for the attack as it is far from Cpep (5.1 Å), and thus the approximation implies high reorganization energy. In addition, it is a poor nucleophile due to the water molecule’s Zn2+-promoted spontaneous and barrierless protonation. In line with all previous observations here, the “Glutamate-assisted nucleophilic pathway” cannot explain the reactivity of RVV-X. As this pathway led to a very high energy rise for the first reaction step and a stable intermediate was not formed, we considered this mechanism unfeasible and continued by studying the water-promoted pathway.

Water-Promoted Pathway, First Step—Nucleophilic Attack

The first step of the proposed reaction is the deprotonation of the Zn2+-bound water molecule by Glu140. However, this step is barrierless at the active site of RVV-X, and thus the Michaelis complex already has the hydroxide ion formed at the active site. Thus, the first step of the mechanism of RVV-X is the nucleophilic attack of the Zn2+-bound hydroxide ion on the carbonyl carbon of the substrate’s scissile bond. The Owat–Cpep distance was chosen as the reaction coordinate. This step resulted in the formation of the first transition state (TS1, Figure 5B) and the first intermediate (INT1, Figure 5C). These stationary states were subsequently entirely and unconstrainedly optimized (after IRC calculations for the two minima). At TS1, the Owat–Cpep distance shortened to 1.84 Å, and the Zn2+–OH distance increased to 2.18 Å. Simultaneously, the distance between Opep and Zn2+ also decreased from 2.65 to 2.03 Å, which is expectable as Opep receives the negative charge of the hydroxide and becomes anionic, along with the change of hybridization of Cpep to sp3. This first step yielded a free energy barrier of 17.7 kcal·mol–1. The vibrational frequency involving the reaction coordinate for TS1 resulted in a single imaginary value of 218i cm–1. Furthermore, the distance between the Glu140 proton and the nitrogen atom of the scissile peptide bond (Npep) decreased from 3.75 to 3.08 Å when moving from the reactant to TS1. In the first intermediate (INT1, Figure 5C), the Owat–Cpep distance is 1.49 Å, and Opep gains an sp3 hybridization and becomes negatively charged.

Second Step—Proton Transfer and Peptide Bond Cleavage

In the second step, the sidechain of Glu140, which acts as a proton shuttle, donates the proton taken from the catalytic water to the scissile-bond nitrogen (Npep), triggering the breaking of the peptide bond. Thus, the distance between the Glu140 hydrogen and the scissile-bond nitrogen (HGlu–Npep) was chosen as the reaction coordinate, whose scan led to the identification of the second transition state (TS2, Figure 5D), which was subsequently fully and freely optimized. The product of this step and the second intermediate of the cycle (INT2, Figure 5E) was determined through IRC calculations and subsequently entirely and unconstrained optimized. At the TS2, the proton taken from the catalytic water molecule by the Glu140 was transferred to Npep, leading to a sharp drop in the HGlu–Npep distance, from 3.12 Å at INT1 to 1.35 Å at TS2. Furthermore, the Cpep–Npep peptide bond distance increased from 1.45 at INT1 to 1.57 Å at TS2 (SI, Figure S5).

The hydrogen-bond network is rearranged along this step: the hydrogen bond between the peptide’s hydroxyl group of the scissile bond and the backbone oxygen of Thr218 (OHpep–OThr) is replaced by a new H-bond between the protonated carboxyl group of the peptide’s scissile bond and the Oε1 of Glu140 (OHpep–Oε1Glu), with a distance of 1.75 Å (SI, Figure S6). This spontaneous H-bond rearrangement greatly facilitates the protonation of Glu140 by the substrate’s C-terminal carboxylic acid, which is very acidic at this stage due to the pKa drop induced by its coordination to the Zn2+ cofactor. Moreover, the proton of the Glu140 sidechain carboxylate is no longer bonded to the carboxyl group but is instead halfway between the carboxylate and the peptide bond nitrogen. These events led to forming a Zn2+-bound gem-diolate transition state with both oxygens of the C-terminal carboxylic acid coordinated to the metal ion in a bidentate fashion with distances of 2.35 and 1.95 Å, the shorter distance corresponding to the anionic oxygen. The free energy barrier for this step is the highest of the cycle, 23.1 kcal·mol–1 above the reactants, becoming the rate-limiting transition state. The vibrational frequency for the TS2 structure is 1046i cm–1.

Interestingly, this coordination mode is mimicked by hydroxamate-based inhibitors such as batimastat or marimastat, which were developed as antineoplastic drug candidates targeting the human matrix metalloproteinases but failed at clinical trials. These drugs are under investigation to treat snakebite, as they are highly efficient in inhibiting the SVMPs in vitro and in vivo.14,20,90 The drug candidate’s carbonyl group resembles the substrate’s carbonyl group, and the hydroxamic acid hydroxyl group resembles the hydroxide ion (TS2, Figure 5D and SI, Figure S3). The fact that these drug candidates precisely mimic the rate-limiting transition state of the whole cycle reinforces the strategy of inspiring drug discovery on the rate-limiting transition states determined through accurate QM/MM calculations.

The rearrangements resulted in a gem-diolate intermediate INT2 (Figure 5E) with a severely weakened Cpep–Npep bond (1.68 Å). At this stage, the Glu140 proton has been transferred entirely to Npep. Simultaneously, there was a decrease in the distance between the Zn2+ and the substrate hydroxyl group (2.33 Å) (SI, Figure S5). The Zn2+–Opep distance increased to 1.95 Å as it moved from an oxyanion (C–O) to neutral oxygen (C=O). The INT2 structure retained the distorted penta-coordinated shell around the Zn2+ cofactor (Figure 5E), composed of the oxygen atoms of the C-product carboxyl group (2.33 and 1.98 Å) and the histidines 139, 143, and 149.

As the peptide bond was not fully broken at the end of the second step, the Cpep–Npep distance was used as the reaction coordinate for the last step. The energy barrier of the definitive breaking of the peptide bond was less than 1 kcal·mol–1. The transition state between the INT2 and the products, with the peptide bond completely broken (PROD, Figure 5F), was not further optimized because the slope of the potential energy surface in the respective region was very flat. The associated barrier is very shallow (close to kBT)91 at the biologically relevant range of temperatures (25–37 °C), i.e., the range of body temperatures of the principal RV prey, for example, rodents. Therefore, the crossing rate from INT2 to TS3 is within the same timescale as bond vibration. This very fast step results in the formation of PROD. This reaction is exergonic, with a sharp free energy drop at PROD to −0.9 kcal·mol–1 (Figure 6), resulting from the peptide bond’s cleavage and leading to a Cpep–Npep distance of 2.93 Å. At the end of this step, the proton of the substrate’s carboxyl group and the proton of the Oε2Glu are entirely transferred to the Glu140 and the amine-free nitrogen, respectively. The Zn2+–Owat distance increases from 2.33 to 2.86 Å, which leads to a short decrease in the Zn2+–Opep distance (SI, Figure S5). These events determine the end of the reactive part of the enzyme cycle.

Figure 6.

Figure 6

Water-promoted pathway: Gibbs energy profile along the reaction progress at the B3LYP/6-311++G(3df,2pd) with added Grimme’s D3-BJ dispersion corrections. The critical residues for each reaction step are shown in CPK representation. The first and second steps of the reaction yielded activation free energies of 17.7 and 23.1 kcal·mol–1, respectively. The reaction is exergonic; the overall reaction free energy is −0.9 kcal·mol–1.

Albeit the product is formed, the enzyme is not entirely reconstituted. The reactant state’s complete regeneration still involves replacing the bound product with a water molecule from the bulk solvent; the carboxylate of the leaving C-terminal portion of the product probably deprotonates Glu140 and fully regenerates the enzyme for a new cycle. However, the product’s release is very challenging to simulate, and as it is a physical process that is well understood but demands extensive sampling to simulate, it is generally not included in mechanistic studies of enzyme catalysis. Figure 6 shows the free energy profile for the complete reaction and the evolution of the critical distances along the complete reaction cycle is presented in the SI, Figure S5. See Supporting information, Scheme S2 for a schematic diagram of the final reaction mechanism for the RVV-X.

The rate-limiting step (23.1 kcal·mol–1) is higher than the one in experimental studies for MMP-2, 15–17 kcal·mol–1, obtained by kinetic studies.56,92,93 However, these experimental studies have been made with different enzymes, hindering the direct comparison of values. Zinc hydrolases generally have rate-limiting free energies of 16.9 ± 2.1 (mean ± SD)94 The value found here is consistent with this interval, and the difference is within the error of the computational method (3–5 kcal·mol–1). In addition, it is well known that the accurate description of energy barriers demands a significant fraction of HF exchange, but the description of transition metals demands a low fraction of HF exchange or no HF exchange. Therefore, there is a conflict in the need for HF exchange in studying barriers involving transition metals, such as the case here. In the face of this, and to evaluate the influence of the density functional in the results, we recalculated the ONIOM energy with the M06 and M06-L density functionals and the 6-311++G (3df, 2pd) basis set (results in Table S2). The results were very similar with all functionals, with an MUE between B3LYP and M06 of 2.4 kcal·mol–1 and between B3LYP and M06-L of 0.8 kcal·mol–1. The rate-limiting step with M06 and M06-L was lower, 17.8 and 21.1 kcal·mol–1, respectively.

The potential energy surface of 10 additional starting structures taken from a 100 ns MD simulation was analyzed, from which the lowest and highest obtained free barriers (TS2 in relation to ES) ranged from 15.85 to 26.98 kcal·mol–1 (Table 1), at the B3LYP/6-311++G(3df,2pd)-D3:ff14SB level of theory. Many of these free energy barriers ranged from 18 to 22 kcal·mol–1. Although the 10 starting structures exhibited a quite similar QM region with almost the same interatomic distances, these nonetheless account for nearly 11 kcal·mol–1 energy difference. No pronounced differences in the high-level region of the structures were found that could justify the variation of the above-mentioned barrier, meaning that the differences come from medium- and long-range interactions. In other studies of enzyme reaction mechanisms, a significant spread in the activation energy values originated from different starting structures has also been observed, with a low-tail close to the experimental value and a spread to larger values corresponding to enzyme structures that are not perfectly pre-organized to react.9597 Here, the scenario is not different. Finally, as the observed rate is exponentially dominated by the lowest barriers,98 the calculations further confirm that the rate-limiting step is close to the experimental studies for MMP-2, 15–17 kcal·mol–1, as previously mentioned.

Table 1. Rate-limiting Activation Free Energies at the ONION(B3LYP/6-311++G(3df,2dp)-D3:MM) Level Starting from 10 Initial Structures.

conformation rate-limiting ΔGact/kcal·mol–1 conformation rate-limiting ΔGact/kcal·mol–1
X-ray 23.1 conf. 6 27.0
conf. 1 21.5 conf. 7 24.4
conf. 2 17.7 conf. 8 22.8
conf. 3 21.4 conf. 9 22.2
conf. 4 19.7 conf. 10 15.8
conf. 5 17.0    

Role of the Zn2+ Cofactor

The Zn2+ cofactor has many different roles. It is obviously important for coordinating the substrate and the organization of the active site. Nevertheless, we wanted to understand its role in the first-order reaction rate, i.e., in lowering the activation free energy of the chemical steps, with an emphasis on the rate-limiting step. To do so, we calculated the contribution of the Zn2+ ion to the barrier of each reaction step (SI, Figure S7). For that purpose, we subtracted the Zn2+ ion energy and interactions (i.e., recalculating the energy in single-point calculations with the 6-31G(d) basis set but deleting the Zn2+ ion) and only the electronic energy of the QM layer was used. When comparing the energy profile without the Zn2+ cofactor with that obtained for the wild-type using the same basis set and including only the QM layer, the rate-limiting barrier for the wild-type reaction was 13 kcal·mol–1 lower than the value obtained without the cofactor (22.9 kcal·mol–1).

Without the cofactor, the REACT state is not the ground state anymore. Instead, the ground state has an anionic Glu140 and a neutral water molecule, as the Glu140 does not spontaneously and barrierlessly deprotonate the nucleophilic water molecule. Without the cofactor, the rate-limiting step of the whole cycle becomes precisely the deprotonation of the nucleophilic water, and the rate-limiting barrier increases by 9.0 kcal·mol–1. Thus, it becomes clear that the principal role of the metal ion is the in situ generation of the hydroxide nucleophile.

The problem of generating a strong nucleophile at the active site is transversal for the protease superfamily.99 The nucleophile has to be generated in situ to avoid side reactions. For example, most serine proteases rely on a Ser–His–Asp/Glu triad to overcome the burden of deprotonating the very basic serine residue, whose pKa in aqueous solution is very high.100 At the active site, the slightly basic histidine residue efficiently deprotonates the serine because the Asp/Glu residue of the triad is pre-organized to stabilize the positive form of the histidine that results from the serine deprotonation. The same principles and catalytic machinery can be found in the large family of cysteine proteases.101 In both families, the oxyanion generated upon nucleophilic attack is generally stabilized by an oxyanion hole constituted by two hydrogen-bond donors, often backbone amines.99101

In the case of aspartic proteases, the existence of a pair of active-site Asp residues, one neutral and one negative, allows overcoming the burden of deprotonating the very basic water molecule. While one of the Asp residues deprotonates the hydrolytic water molecule, the second Asp residue protonates the very basic aliphatic oxyanion that is generated by the water attack on the substrate’s carbonyl98 so the burden of the deprotonation is compensated by the simultaneous very exothermic protonation of the growing oxyanion.99,102 In some cases, the protonation and deprotonation can be carried out by a single Asp residue.103

The strategy of RVV-X for generating the nucleophile is to coordinate the water molecule to the Zn2+ cofactor, lowering the water pKa through the much stronger coordination of the hydroxide ion in relation to the neutral water. The water deprotonation molecule moves the negative charge from the non-Zn2+-coordinated Glu140 to the Zn2+-coordinated water molecule, with an obvious electrostatic gain. In the subsequent reactions, the number of ionic oxygen atoms coordinated to the Zn2+ ion is kept unchanged—there are always two Zn2+-bound oxygen atoms, one negative and one neutral, except for the Michaelis complex before the water deprotonation, where both Zn2+-coordinated oxygen atoms are neutral. Thus, it makes sense that the cofactor does not make a decisive contribution to lowering the subsequent energy barriers.

Given the very significant sequence identity among zinc metalloproteases, particularly at the active site, it is tempting to speculate that the role of the Zn2+ cofactor might be transversal to the family of metalloproteases. However, such generalization, albeit logical and even probable, needs a solid analysis ground in further family members.

Comparison with Other Studies

The obtained stationary geometries are close to those found in studies of human matrix metalloproteinases.55,56,59,104 The catalytic reaction follows a two-step mechanism consistent with the study carried out by Pelmenschikov and Siegbahn on human Matrix metalloprotease-3 (MMP-3).55 However, unlike the proposed mechanism for MMP-3, the rate-limiting step in the RVV-X mechanism is not the nucleophilic attack but the protonation of the leaving amino group. This last finding agrees with the study carried out by Díaz and Suárez on the human MMP-2 enzyme toward two peptide substrates.56

Overall, the QM/MM results provide solid support for the water-promoted pathway mechanism proposed for the human MMPs. Furthermore, we show that the pre-reactive complex is well organized for the reaction to occur according to this proposal. In 2015, Vasilevskaya and their colleagues demonstrated that pulling out the C-product from the active site (estimated free energy barrier of 20 kcal·mol–1) causes the product’s NH2 group to rotate, allowing the Glutamate carboxyl group also to rotate and, in turn, protonate the N-product.104 However, Pelmenschikov and Siegbahn showed in 2002 that the thermolysin N-product could be released in its neutral state without needing a second proton transfer.62

Conclusions

We present the first simulation of the reaction mechanism of RVV-X, an SVMP. This toxin is of keen interest due to its pharmacological effects following envenomation and biotechnological applications. The findings of this study on the RVV-X mechanism are generally consistent with the proposed “water-promoted pathway” mechanism of human metalloproteinases. The most evident difference is that the RVV-X ground state already has the nucleophilic hydroxide ready for action. The latter is generated in situ through spontaneous and barrierless deprotonation due to the extensive lowering of the pKa of the Zn2+-bound water molecule and the pre-organization of Glu140 in an ideal position for deprotonation. The results further indicate that the most significant role of the Zn2+ cofactor is precise in generating the nucleophile; its role in stabilizing the following transition states is relatively minor. This role is well framed within the challenges that the large and ubiquitous protease superfamily face to hydrolyze peptide bonds. The high active site conservation might be a general trait of zinc metalloproteinases. In addition, we show that the coordination mode of therapeutic inhibitors of the human MMPs, such as batimastat and marimastat, presently under study for snakebite treatment, perfectly mimics the one of the rate-limiting transition of RVV-X state found in this study. Thus, understanding the reaction mechanism and catalysis of this fantastic enzyme shall open avenues for the rational development of other high-affinity RVV-X inhibitors with relevant therapeutic potential to combat the RV envenomation.

Acknowledgments

The authors acknowledge financial support from FCT/MCTES—the Portuguese Fundação para a Ciência e Tecnologia, through project PTDC/QUI-OUT/1401/2020 and from the Laboratório Associado para a Química Verde (LAQV), which is financed by FCT/MCTES within the scope of project UIDB/50006/2020.

Data Availability Statement

The cartesian coordinates and the force field parameters of all stationary points discussed in this paper are available in the Supporting Information as a ZIP file. The system was prepared using AMBER18 software. The AMBER package can be purchased at https://ambermd.org/index.php. H++ web server (http://biophysics.cs.vt.edu/H++) was used for the prediction of the protein titratable residues pKa. VMD v1.9.4 and PyMOL v2.3.0 were used for visualization and analysis of the system. VMD can be downloaded from https://www.ks.uiuc.edu/Research/vmd/ and PyMOL can be downloaded from https://pymol.org/. The VMD MolUP extension was used to prepare Gaussian input files and to visualize results from Gaussian output files. The molUP extension can be downloaded from https://biosim.pt/molup/. Gaussian 09 D01 was used for the QM/MM calculations. Gaussian software can be purchased from https://gaussian.com/.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jcim.2c01156.

  • SVMP different classes; model construction; pKa values; minimization and molecular dynamics simulations; stationary points; energetic data; representation of the anhydride pathway, and representation of the final mechanism obtained for RVV-X (PDF)

  • PDB files and Gaussian input files used for the optimization of the stationary points, parameter files of the initial system, and restart file of the minimized structure (ZIP)

The authors declare no competing financial interest.

Supplementary Material

ci2c01156_si_001.pdf (2.1MB, pdf)
ci2c01156_si_002.zip (2.9MB, zip)

References

  1. Oliveira A. L.; Viegas M. F.; da Silva S. L.; Soares A. M.; Ramos M. J.; Fernandes P. A. The chemistry of snake venom and its medicinal potential. Nat. Rev. Chem. 2022, 6, 451–469. 10.1038/s41570-022-00393-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Gutiérrez J. M.; Escalante T.; Rucavado A.; Herrera C.; Fox J. W. A Comprehensive View of the Structural and Functional Alterations of Extracellular Matrix by Snake Venom Metalloproteinases (SVMPs): Novel Perspectives on the Pathophysiology of Envenoming. Toxins 2016, 8, 304 10.3390/toxins8100304. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Preciado L. M.; Pereañez J. A.; Comer J. Potential of matrix metalloproteinase inhibitors for the treatment of local tissue damage induced by a type PI snake venom metalloproteinase. Toxins 2020, 12, 8 10.3390/toxins12010008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Kalita B.; Mackessy S. P.; Mukherjee A. K. Proteomic analysis reveals geographic variation in venom composition of Russell’s Viper in the Indian subcontinent: implications for clinical manifestations post-envenomation and antivenom treatment. Expert Rev. Proteomics 2018, 15, 837–849. 10.1080/14789450.2018.1528150. [DOI] [PubMed] [Google Scholar]
  5. Thakur R.; Mukherjee A.; Biotechnology M. A brief appraisal on Russell’s Viper venom (Daboia russelii russelii) proteinases. Snake Venoms 2015, 1–18. 10.1007/978-94-007-6648-8_19-1. [DOI] [Google Scholar]
  6. Xie C.; Albulescu L.; Bittenbinder M. A.; Somsen G.; Vonk F.; Casewell N. R.; Kool J. Neutralizing effects of small molecule inhibitors and metal chelators on coagulopathic Viperinae snake venom toxins. Biomedicines 2020, 8, 297 10.3390/biomedicines8090297. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Laxme R. S.; Khochare S.; de Souza H. F.; Ahuja B.; Suranse V.; Martin G.; Whitaker R.; Sunagar K. Beyond the ‘big four’: Venom profiling of the medically important yet neglected Indian snakes reveals disturbing antivenom deficiencies. PLoS Neglected Trop. Dis. 2019, 13, e0007899 10.1371/journal.pntd.0007899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Faisal T.; Tan K. Y.; Sim S. M.; Quraishi N.; Tan N. H.; Tan C. H. Proteomics, functional characterization and antivenom neutralization of the venom of Pakistani Russell’s viper (Daboia russelii) from the wild. J. Proteomics 2018, 183, 1–13. 10.1016/j.jprot.2018.05.003. [DOI] [PubMed] [Google Scholar]
  9. Kalita B.; Patra A.; Das A.; Mukherjee A. K. Proteomic Analysis and Immuno-Profiling of Eastern India Russell’s Viper (Daboia russelii) Venom: Correlation between RVV Composition and Clinical Manifestations Post RV Bite. J. Proteome Res. 2018, 17, 2819–2833. 10.1021/acs.jproteome.8b00291. [DOI] [PubMed] [Google Scholar]
  10. Pla D.; Sanz L.; Quesada-Bernat S.; Villalta M.; Baal J.; Chowdhury M. A. W.; León G.; Gutiérrez J. M.; Kuch U.; Calvete J. J. Phylovenomics of Daboia russelii across the Indian subcontinent. Bioactivities and comparative in vivo neutralization and in vitro third-generation antivenomics of antivenoms against venoms from India, Bangladesh and Sri Lanka. J. Proteomics 2019, 207, 103443 10.1016/j.jprot.2019.103443. [DOI] [PubMed] [Google Scholar]
  11. Kalita B.; Singh S.; Patra A.; Mukherjee A. Quantitative proteomic analysis and antivenom study revealing that neurotoxic phospholipase A 2 enzymes, the major toxin class of Russell’s viper venom from southern India, shows the least immuno-recognition and neutralization by commercial polyvalent antivenom. Int. J. Biol. Macromol. 2018, 118, 375–385. 10.1016/j.ijbiomac.2018.06.083. [DOI] [PubMed] [Google Scholar]
  12. Laxme R. R. S.; Khochare S.; Attarde S.; Suranse V.; Iyer A.; Casewell N. R.; Whitaker R.; Martin G.; Sunagar K. Biogeographic venom variation in Russell’s viper (Daboia russelii) and the preclinical inefficacy of antivenom therapy in snakebite hotspots. PLOS Negl. Trop. Dis. 2021, 15, e0009247 10.1371/journal.pntd.0009247. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Thorpe R.; Pook C.; Malhotra A. Phylogeography of the Russell’s viper (Daboia russelii) complex in relation to variation in the colour pattern and symptoms of envenoming. Herpetol. J. 2007, 17, 209–218. [Google Scholar]
  14. Albulescu L.-O.; Xie C.; Ainsworth S.; Alsolaiss J.; Crittenden E.; Dawson C. A.; Softley R.; Bartlett K. E.; Harrison R. A.; Kool J.; Casewell N. R. A therapeutic combination of two small molecule toxin inhibitors provides broad preclinical efficacy against viper snakebite. Nat. Commun. 2020, 11, 6094 10.1038/s41467-020-19981-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Perales J.; Domont G. B.; Neves-Ferreira A. G. C.; Valente R. H.. Natural inhibitors: innate immunity to snake venoms. In Handbook of Venoms and Toxins of Reptiles; Mackessy S. P.; CRC Press, 2009; pp 259–284.
  16. Liu C.-C.; Wu C.-J.; Hsiao Y.-C.; Yang Y.-H.; Liu K.-L.; Huang G.-J.; Hsieh C.-H.; Chen C.-K.; Liaw G.-W. Snake venom proteome of Protobothrops mucrosquamatus in Taiwan: Delaying venom-induced lethality in a rodent model by inhibition of phospholipase A2 activity with varespladib. J. Proteomics 2021, 234, 104084 10.1016/j.jprot.2020.104084. [DOI] [PubMed] [Google Scholar]
  17. Slagboom J.; Kool J.; Harrison R. A.; Casewell N. R. Haemotoxic snake venoms: their functional activity, impact on snakebite victims and pharmaceutical promise. Br. J. Haematol. 2017, 177, 947–959. 10.1111/bjh.14591. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Williams D. J.; Faiz M. A.; Abela-Ridder B.; Ainsworth S.; Bulfone T. C.; Nickerson A. D.; Habib A. G.; Junghanss T.; Fan H. W.; Turner M.; et al. Strategy for a globally coordinated response to a priority neglected tropical disease: Snakebite envenoming. PLoS Neglected Trop. Dis. 2019, 13, e0007059 10.1371/journal.pntd.0007059. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Aoki-Shioi N.; Koh C. Y.; Kini R. M. Natural inhibitors of snake venom metalloproteinases. Aust. J. Chem. 2020, 73, 277–286. 10.1071/CH19414. [DOI] [Google Scholar]
  20. Puzari U.; Fernandes P. A.; Mukherjee A. K. Advances in the Therapeutic Application of Small-Molecule Inhibitors and Repurposed Drugs against Snakebite. J. Med. Chem. 2021, 64, 13938–13979. 10.1021/acs.jmedchem.1c00266. [DOI] [PubMed] [Google Scholar]
  21. Suntravat M.; Yusuksawad M.; Sereemaspun A.; Pérez J. C.; Nuchprayoon I. Effect of purified Russell’s viper venom-factor X activator (RVV-X) on renal hemodynamics, renal functions, and coagulopathy in rats. Toxicon 2011, 58, 230–238. 10.1016/j.toxicon.2011.05.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Mukherjee A. K.; Kalita B.; Mackessy S. P. A proteomic analysis of Pakistan Daboia russelii russelii venom and assessment of potency of Indian polyvalent and monovalent antivenom. J. Proteomics 2016, 144, 73–86. 10.1016/j.jprot.2016.06.001. [DOI] [PubMed] [Google Scholar]
  23. Mukherje A. K.; Ghosal S. K.; Maity C. R. Some biochemical properties of Russell’s viper (Daboia russelli) venom from Eastern India: correlation with clinico-pathological manifestation in Russell’s viper bite. Toxicon 2000, 38, 163–175. 10.1016/s0041-0101(99)00125-7. [DOI] [PubMed] [Google Scholar]
  24. Sharma M.; Das D.; Iyer J. K.; Kini R. M.; Doley R. Unveiling the complexities of Daboia russelii venom, a medically important snake of India, by tandem mass spectrometry. Toxicon 2015, 107, 266–281. 10.1016/j.toxicon.2015.06.027. [DOI] [PubMed] [Google Scholar]
  25. Bhattacharya S.; Krishnamurthy A.; Gopalakrishnan M.; Kalra S.; Kantroo V.; Aggarwal S.; Surana V. Endocrine and Metabolic Manifestations of Snakebite Envenoming. Am. J. Trop. Med. Hyg. 2020, 103, 1388–1396. 10.4269/ajtmh.20-0161. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Cardoso F. C.; Ferraz C. R.; Arrahman A.; Xie C.; Casewell N. R.; Lewis R. J.; Kool J. Multifunctional toxins in snake venoms and therapeutic implications: from pain to hemorrhage and necrosis. Front. Ecol. Evol. 2019, 7, 218 10.3389/fevo.2019.00218. [DOI] [Google Scholar]
  27. Olaoba O. T.; dos Santos P. K.; Selistre-de-Araujo H. S.; de Souza D. H. F. Snake Venom Metalloproteinases (SVMPs): A structure-function update. Toxicon: X 2020, 7, 100052 10.1016/j.toxcx.2020.100052. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Valente R. H.; Dragulev B.; Perales J.; Fox J. W.; Domont G. B. BJ46a, a snake venom metalloproteinase inhibitor: Isolation, characterization, cloning and insights into its mechanism of action. Eur. J. Biochem. 2001, 268, 3042–3052. 10.1046/j.1432-1327.2001.02199.x. [DOI] [PubMed] [Google Scholar]
  29. Takeda S. ADAM and ADAMTS family proteins and snake venom metalloproteinases: A structural overview. Toxins 2016, 8, 155 10.3390/toxins8050155. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Markland F. S. Jr; Swenson S. Snake venom metalloproteinases. Toxicon 2013, 62, 3–18. 10.1016/j.toxicon.2012.09.004. [DOI] [PubMed] [Google Scholar]
  31. Takeda S.; Takeya H.; Iwanaga S. Snake venom metalloproteinases: structure, function and relevance to the mammalian ADAM/ADAMTS family proteins. Biochim. Biophys. Acta, Proteins Proteomics 2012, 1824, 164–176. 10.1016/j.bbapap.2011.04.009. [DOI] [PubMed] [Google Scholar]
  32. Camacho E.; Sanz L.; Escalante T.; Pérez A.; Villalta F.; Lomonte B.; Neves-Ferreira A. G. C.; Feoli A.; Calvete J. J.; Gutiérrez J. M.; Rucavado A. Novel catalytically-inactive PII metalloproteinases from a viperid snake venom with substitutions in the canonical zinc-binding motif. Toxins 2016, 8, 292 10.3390/toxins8100292. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Watanabe L.; Shannon J. D.; Valente R. H.; Rucavado A.; Alape-Girón A.; Kamiguti A. S.; Theakston R. D. G.; Fox J. W.; Gutiérrez J. M.; Arni R. K. Amino acid sequence and crystal structure of BaP1, a metalloproteinase from Bothrops asper snake venom that exerts multiple tissue-damaging activities. Protein Sci. 2009, 12, 2273–2281. 10.1110/ps.03102403. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Kini R. M.; Koh C. Y. Metalloproteases affecting blood coagulation, fibrinolysis and platelet aggregation from snake venoms: Definition and nomenclature of interaction sites. Toxins 2016, 8, 284 10.3390/toxins8100284. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Chen H. S.; Chen J. M.; Lin C. W.; Khoo K. H.; Tsai I. H. New insights into the functions and N-glycan structures of factor X activator from Russell’s viper venom. FEBS J. 2008, 275, 3944–3958. 10.1111/j.1742-4658.2008.06540.x. [DOI] [PubMed] [Google Scholar]
  36. Takeda S.Structural Aspects of the Factor X Activator RVV-X from Russell’s Viper Venom. In Toxins and Hemostasis; Springer: Dordrecht, 2010; pp 465–484. [Google Scholar]
  37. Takeya H.; Nishida S.; Miyata T.; Kawada S.; Saisaka Y.; Morita T.; Iwanaga S. Coagulation factor X activating enzyme from Russell’s viper venom (RVV-X). A novel metalloproteinase with disintegrin (platelet aggregation inhibitor)-like and C-type lectin-like domains. J. Biol. Chem. 1992, 267, 14109–14117. 10.1016/S0021-9258(19)49685-3. [DOI] [PubMed] [Google Scholar]
  38. Mohamed Abd El-Aziz T.; Soares A. G.; Stockand J. D. Snake venoms in drug discovery: valuable therapeutic tools for life saving. Toxins 2019, 11, 564 10.3390/toxins11100564. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Yee K. T.; Pitts M.; Tongyoo P.; Rojnuckarin P.; Wilkinson M. C. Snake venom metalloproteinases and their peptide inhibitors from Myanmar Russell’s viper venom. Toxins 2017, 9, 15 10.3390/toxins9010015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Takeda S.; Igarashi T.; Mori H. Crystal structure of RVV-X: An example of evolutionary gain of specificity by ADAM proteinases. FEBS Lett. 2007, 581, 5859–5864. 10.1016/j.febslet.2007.11.062. [DOI] [PubMed] [Google Scholar]
  41. Venkateswarlu D.; Perera L.; Darden T.; Pedersen L. G. Structure and dynamics of zymogen human blood coagulation factor X. Biophys. J. 2002, 82, 1190–1206. 10.1016/S0006-3495(02)75476-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Vonk F. J.; Jackson K.; Doley R.; Madaras F.; Mirtschin P.; Vidal N. Snake venom: From fieldwork to the clinic. Bioessays 2011, 33, 269–279. 10.1002/bies.201000117. [DOI] [PubMed] [Google Scholar]
  43. Kini R. M.; Banerjee Y. Dissection approach: a simple strategy for the identification of the step of action of anticoagulant agents in the blood coagulation cascade. J. Thromb. Haemostasis 2005, 3, 170–171. 10.1111/j.1538-7836.2004.01056.x. [DOI] [PubMed] [Google Scholar]
  44. Marsh N. A. Diagnostic Uses of Snake Venom. Pathophysiol. Haemostasis Thromb. 2001, 31, 211–217. 10.1159/000048065. [DOI] [PubMed] [Google Scholar]
  45. Alvarez-Flores M.; Faria F.; de Andrade S.; Chudzinski-Tavassi A.. Snake Venom Components Affecting the Coagulation System. In Snake venoms; Springer: Netherlands, 2016; pp 1–20. [Google Scholar]
  46. Herrera C.; Voisin M.-B.; Escalante T.; Rucavado A.; Nourshargh S.; Gutiérrez J. M. Effects of PI and PIII Snake Venom Haemorrhagic Metalloproteinases on the Microvasculature: A Confocal Microscopy Study on the Mouse Cremaster Muscle. PLoS One 2016, 11, e0168643. 10.1371/journal.pone.0168643. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Zhu X.-Y.; Teng M. K.; Niu L. W. Structure of acutolysin-C, a haemorrhagic toxin from the venom of Agkistrodon acutus, providing further evidence for the mechanism of the pH-dependent proteolytic reaction of zinc metalloproteinases. Acta Crystallogr., Sect. D: Biol. Crystallogr. 1999, 55, 1834–1841. 10.1107/S0907444999010306. [DOI] [PubMed] [Google Scholar]
  48. Gong W.; Zhu X.; Liu S.; Teng M.; Niu L. Crystal structures of acutolysin A, a three-disulfide hemorrhagic zinc metalloproteinase from the snake venom of Agkistrodon acutus11Edited by R. Huber. J. Mol. Biol. 1998, 283, 657–668. 10.1006/jmbi.1998.2110. [DOI] [PubMed] [Google Scholar]
  49. Chou T.-L.; Wu C.-H.; Huang K.-F.; Wang A. H. J. Crystal structure of a Trimeresurus mucrosquamatus venom metalloproteinase providing new insights into the inhibition by endogenous tripeptide inhibitors. Toxicon 2013, 71, 140–146. 10.1016/j.toxicon.2013.05.009. [DOI] [PubMed] [Google Scholar]
  50. Gomis-Rüth F. X.; Kress L. F.; Bode W. First structure of a snake venom metalloproteinase: a prototype for matrix metalloproteinases/collagenases. EMBO J. 1993, 12, 4151–4157. 10.1002/j.1460-2075.1993.tb06099.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Lingott T.; Schleberger C.; Gutiérrez J. M.; Merfort I. High-resolution crystal structure of the snake venom metalloproteinase BaP1 complexed with a peptidomimetic: insight into inhibitor binding. Biochemistry 2009, 48, 6166–6174. 10.1021/bi9002315. [DOI] [PubMed] [Google Scholar]
  52. Gomis-Rüth F. X.; Meyer E. F.; Kress L. F.; Politi V. Structures of adamalysin II with peptidic inhibitors. Implications for the design of tumor necrosis factor α convertase inhibitors. Protein Sci. 1998, 7, 283–292. 10.1002/pro.5560070207. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Cirilli M.; Gallina C.; Gavuzzo E.; Giordano C.; Gomis-Rüth F.; Gorini B.; Kress L.; Mazza F.; Paradisi M. P.; Pochetti G. 2 Å X-ray structure of adamalysin II complexed with a peptide phosphonate inhibitor adopting a retro-binding mode. FEBS Lett. 1997, 418, 319–322. 10.1016/S0014-5793(97)01401-4. [DOI] [PubMed] [Google Scholar]
  54. Igarashi T.; Araki S.; Mori H.; Takeda S. Crystal structures of catrocollastatin/VAP2B reveal a dynamic, modular architecture of ADAM/adamalysin/reprolysin family proteins. FEBS Lett. 2007, 581, 2416–2422. 10.1016/j.febslet.2007.04.057. [DOI] [PubMed] [Google Scholar]
  55. Pelmenschikov V.; Siegbahn P. E. M. Catalytic Mechanism of Matrix Metalloproteinases: Two-Layered ONIOM Study. Inorg. Chem. 2002, 41, 5659–5666. 10.1021/ic0255656. [DOI] [PubMed] [Google Scholar]
  56. Díaz N.; Suárez D.; Suárez E. Kinetic and binding effects in peptide substrate selectivity of matrix metalloproteinase-2: Molecular dynamics and QM/MM calculations. Proteins: Struct., Funct., Bioinf. 2010, 78, 1–11. 10.1002/prot.22493. [DOI] [PubMed] [Google Scholar]
  57. Vasilevskaya T.; Khrenova M. G.; Nemukhin A. V.; Thiel W. Methodological aspects of QM/MM calculations: A case study on matrix metalloproteinase-2. J. Comput. Chem. 2016, 37, 1801–1809. 10.1002/jcc.24395. [DOI] [PubMed] [Google Scholar]
  58. Wu S.; Zhang C.; Xu D.; Guo H. Catalysis of Carboxypeptidase A: Promoted-Water versus Nucleophilic Pathways. J. Phys. Chem. B 2010, 114, 9259–9267. 10.1021/jp101448j. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Feliciano G. T.; da Silva A. J. R. Unravelling the reaction mechanism of matrix metalloproteinase 3 using QM/MM calculations. J. Mol. Struct. 2015, 1091, 125–132. 10.1016/j.molstruc.2015.02.079. [DOI] [Google Scholar]
  60. Christianson D. W.; Lipscomb W. N. Comparison of carboxypeptidase A and thermolysin: inhibition by phosphonamidates. J. Am. Chem. Soc. 1988, 110, 5560–5565. 10.1021/ja00224a047. [DOI] [Google Scholar]
  61. Crabbe T.; Zucker S.; Cockett M. I.; Willenbrock F.; Tickle S.; O’Connell J. P.; Scothern J. M.; Murphy G.; Docherty A. J. P. Mutation of the Active Site Glutamic Acid of Human Gelatinase A: Effects on Latency, Catalysis, and the Binding of Tissue Inhibitor of Metalloproteinases-1. Biochemistry 1994, 33, 6684–6690. 10.1021/bi00187a039. [DOI] [PubMed] [Google Scholar]
  62. Pelmenschikov V.; Blomberg M. R.; Siegbahn P. E. A theoretical study of the mechanism for peptide hydrolysis by thermolysin. JBIC, J. Biol. Inorg. Chem. 2002, 7, 284–298. 10.1007/s007750100295. [DOI] [PubMed] [Google Scholar]
  63. Christianson D. W.; Lipscomb W. N. Carboxypeptidase A. Acc. Chem. Res. 1989, 22, 62–69. 10.1021/ar00158a003. [DOI] [Google Scholar]
  64. Breslow R.; Wernick D. L. Unified picture of mechanisms of catalysis by carboxypeptidase A. Proc. Natl. Acad. Sci. U.S.A. 1977, 74, 1303–1307. 10.1073/pnas.74.4.1303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Mahmudul I.; Roly Z. Y.; Ruhullah M.; Islam Z.; Sharia A.; Tanvir R. Z.; Foyasal K. Identification of accelerated evolution in the metalloproteinase domain of snake venom metalloproteinase sequences (SVMPs) through comparative analysis. Afr. J. Biotechnol. 2016, 15, 252–263. 10.5897/AJB2015.14818. [DOI] [Google Scholar]
  66. Biasini M.; Bienert S.; Waterhouse A.; Arnold K.; Studer G.; Schmidt T.; Kiefer F.; Cassarino T.; Bertoni M.; Bordoli L.; Schwede T. SWISS-MODEL: Modelling protein tertiary and quaternary structure using evolutionary information. Nucleic Acids Res. 2014, 42, W252–W258. 10.1093/nar/gku340. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Waterhouse A.; Bertoni M.; Bienert S.; Studer G.; Tauriello G.; Gumienny R.; Heer F. T.; de Beer T. A. P.; Rempfer C.; Bordoli L.; et al. SWISS-MODEL: homology modelling of protein structures and complexes. Nucleic Acids Res. 2018, 46, W296–W303. 10.1093/nar/gky427. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Anandakrishnan R.; Aguilar B.; Onufriev A. V. H++ 3.0: automating p K prediction and the preparation of biomolecular structures for atomistic molecular modeling and simulations. Nucleic Acids Res. 2012, 40, W537–W541. 10.1093/nar/gks375. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Gordon J. C.; Myers J. B.; Folta T.; Shoja V.; Heath L. S.; Onufriev A. H++: a server for estimating p K as and adding missing hydrogens to macromolecules. Nucleic Acids Res. 2005, 33, W368–W371. 10.1093/nar/gki464. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Li P.; Merz K. M. Jr. MCPB. py: A Python Based Metal Center Parameter Builder; ACS Publications, 2016. [DOI] [PubMed] [Google Scholar]
  71. Mahoney M. W.; Jorgensen W. L. A five-site model for liquid water and the reproduction of the density anomaly by rigid, nonpolarizable potential functions. J. Chem. Phys. 2000, 112, 8910–8922. 10.1063/1.481505. [DOI] [Google Scholar]
  72. Wang J.; Wolf R. M.; Caldwell J. W.; Kollman P. A.; Case D. A. Development and testing of a general amber force field. J. Comput. Chem. 2004, 25, 1157–1174. 10.1002/jcc.20035. [DOI] [PubMed] [Google Scholar]
  73. Tian C.; Kasavajhala K.; Belfon K. A. A.; Raguette L.; Huang H.; Migues A. N.; Bickel J.; Wang Y.; Pincay J.; Wu Q.; Simmerling C. ff19SB: Amino-Acid-Specific Protein Backbone Parameters Trained against Quantum Mechanics Energy Surfaces in Solution. J. Chem. Theory Comput. 2020, 16, 528–552. 10.1021/acs.jctc.9b00591. [DOI] [PubMed] [Google Scholar]
  74. Stojanovski B. M.; Pelc L.; Cera E. Role of the activation peptide in the mechanism of protein C activation. Sci. Rep. 2020, 10, 11079 10.1038/s41598-020-68078-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Zhou P.; Jin B.; Li H.; Huang S. Y. HPEPDOCK: a web server for blind peptide-protein docking based on a hierarchical algorithm. Nucleic Acids Res. 2018, 46, W443–w450. 10.1093/nar/gky357. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Neves R. P. P.; Cunha A. V.; Fernandes P. A.; Ramos M. J. Towards the Accurate Thermodynamic Characterization of Enzyme Reaction Mechanisms. ChemPhysChem 2022, 23, e202200159 10.1002/cphc.202200159. [DOI] [PubMed] [Google Scholar]
  77. Coimbra J. T. S.; Neves R. P. P.; Cunha A. V.; Ramos M. J.; Fernandes P. A. Different Enzyme Conformations Induce Different Mechanistic Traits in HIV-1 Protease. Eur. J. Chem. 2022, 28, e202201066 10.1002/chem.202201066. [DOI] [PubMed] [Google Scholar]
  78. Fernandes H. S.; Ramos M. J.; Cerqueira N. M. F. S. A. molUP: A VMD plugin to handle QM and ONIOM calculations using the gaussian software. J. Comput. Chem. 2018, 39, 1344–1353. 10.1002/jcc.25189. [DOI] [PubMed] [Google Scholar]
  79. Humphrey W.; Dalke A.; Schulten K. VMD: visual molecular dynamics. J. Mol. Graphics 1996, 14, 33–38. 10.1016/0263-7855(96)00018-5. [DOI] [PubMed] [Google Scholar]
  80. van der Kamp M. W.; Mulholland A. J. Combined Quantum Mechanics/Molecular Mechanics (QM/MM) Methods in Computational Enzymology. Biochemistry 2013, 52, 2708–2728. 10.1021/bi400215w. [DOI] [PubMed] [Google Scholar]
  81. Ranaghan K. E.; Mulholland A. J. Investigations of enzyme-catalysed reactions with combined quantum mechanics/molecular mechanics (QM/MM) methods. Int. Rev. Phys. Chem. 2010, 29, 65–133. 10.1080/01442350903495417. [DOI] [Google Scholar]
  82. Lundberg M.; Sasakura Y.; Zheng G.; Morokuma K. Case studies of ONIOM (DFT: DFTB) and ONIOM (DFT: DFTB: MM) for enzymes and enzyme mimics. J. Chem. Theory Comput. 2010, 6, 1413–1427. 10.1021/ct100029p. [DOI] [Google Scholar]
  83. Yanai T.; Tew D. P.; Handy N. C. A new hybrid exchange–correlation functional using the Coulomb-attenuating method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 51–57. 10.1016/j.cplett.2004.06.011. [DOI] [Google Scholar]
  84. Hariharan P. C.; Pople J. A. The influence of polarization functions on molecular orbital hydrogenation energies. Theory Chim. Acta 1973, 28, 213–222. 10.1007/BF00533485. [DOI] [Google Scholar]
  85. Svensson M.; Humbel S.; Morokuma K. A multilayered integrated MO MM method for geometry optimizations and single point energy predictions. A test for Diels-Alder reactions and Pt (P (t-Bu) 3) 2 H2 oxidative addition. J. Chem. Phys. 1996, 105, 3654–3661. 10.1021/jp962071j. [DOI] [Google Scholar]
  86. Field M. J.; Bash P. A.; Karplus M. A combined quantum mechanical and molecular mechanical potential for molecular dynamics simulations. J. Comput. Chem. 1990, 11, 700–733. 10.1002/jcc.540110605. [DOI] [Google Scholar]
  87. Fukui K. The path of chemical reactions-the IRC approach. Acc. Chem. Res. 1981, 14, 363–368. 10.1021/ar00072a001. [DOI] [Google Scholar]
  88. Lipscomb W. N.; Sträter N. Recent Advances in Zinc Enzymology. Chem. Rev. 1996, 96, 2375–2434. 10.1021/cr950042j. [DOI] [PubMed] [Google Scholar]
  89. Cross J. B.; Duca J. S.; Kaminski J. J.; Madison V. S. The Active Site of a Zinc-Dependent Metalloproteinase Influences the Computed p K a of Ligands Coordinated to the Catalytic Zinc Ion. J. Am. Chem. Soc. 2002, 124, 11004–11007. 10.1021/ja0201810. [DOI] [PubMed] [Google Scholar]
  90. Layfield H. J.; Williams H.; Ravishankar D.; Mehmi A.; Sonavane M.; Salim A.; Vaiyapuri R.; Karthik L.; Vallance T.; Bicknell A.; et al. Repurposing Cancer Drugs Batimastat and Marimastat to Inhibit the Activity of a Group I Metalloprotease from the Venom of the Western Diamondback Rattlesnake, Crotalus atrox. Toxins 2020, 12, 309 10.3390/toxins12050309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Xu D.; Guo H. Quantum Mechanical/Molecular Mechanical and Density Functional Theory Studies of a Prototypical Zinc Peptidase (Carboxypeptidase A) Suggest a General Acid–General Base Mechanism. J. Am. Chem. Soc. 2009, 131, 9780–9788. 10.1021/ja9027988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Fasciglione G. F.; Marini S.; D’Alessio S.; Politi V.; Coletta M. pH-and temperature-dependence of functional modulation in metalloproteinases. A comparison between neutrophil collagenase and gelatinases A and B. Biophys. J. 2000, 79, 2138–2149. 10.1016/S0006-3495(00)76461-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Netzel-Arnett S.; Sang Q. X.; Moore W. G.; Navre M.; Birkedal-Hansen H.; Van Wart H. E. Comparative sequence specificities of human 72-and 92-kDa gelatinases (type IV collagenases) and PUMP (matrilysin). Biochemistry 1993, 32, 6427–6432. 10.1021/bi00076a016. [DOI] [PubMed] [Google Scholar]
  94. Sousa S. F.; Ramos M. J.; Lim C.; Fernandes P. A. Relationship between enzyme/substrate properties and enzyme efficiency in hydrolases. ACS Catal. 2015, 5, 5877–5887. 10.1021/acscatal.5b00923. [DOI] [Google Scholar]
  95. Teixeira L. M. C.; Coimbra J. T. S.; Ramos M. J.; Fernandes P. A. Transmembrane Protease Serine 2 Proteolytic Cleavage of the SARS-CoV-2 Spike Protein: A Mechanistic Quantum Mechanics/Molecular Mechanics Study to Inspire the Design of New Drugs To Fight the COVID-19 Pandemic. J. Chem. Inf. Model. 2022, 62, 2510–2521. 10.1021/acs.jcim.1c01561. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Sousa J. P. M.; Neves R. P. P.; Sousa S. F.; Ramos M. J.; Fernandes P. A. Reaction Mechanism and Determinants for Efficient Catalysis by DszB, a Key Enzyme for Crude Oil Bio-desulfurization. ACS Catal. 2020, 10, 9545–9554. 10.1021/acscatal.0c03122. [DOI] [Google Scholar]
  97. Santos-Martins D.; Calixto A. R.; Fernandes P. A.; Ramos M. J. A Buried Water Molecule Influences Reactivity in α-Amylase on a Subnanosecond Time Scale. ACS Catal. 2018, 8, 4055–4063. 10.1021/acscatal.7b04400. [DOI] [Google Scholar]
  98. Ribeiro A. J. M.; Santos-Martins D.; Russo N.; Ramos M. J.; Fernandes P. A. Enzymatic Flexibility and Reaction Rate: A QM/MM Study of HIV-1 Protease. ACS Catal. 2015, 5, 5617–5626. 10.1021/acscatal.5b00759. [DOI] [Google Scholar]
  99. Elsässer B.; Goettig P. Mechanisms of Proteolytic Enzymes and Their Inhibition in QM/MM Studies. Int. J. Mol. Sci. 2021, 22, 3232 10.3390/ijms22063232. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Hedstrom L. Serine Protease Mechanism and Specificity. Chem. Rev. 2002, 102, 4501–4524. 10.1021/cr000033x. [DOI] [PubMed] [Google Scholar]
  101. Oanca G.; Asadi M.; Saha A.; Ramachandran B.; Warshel A. Exploring the Catalytic Reaction of Cysteine Proteases. J. Phys. Chem. B 2020, 124, 11349–11356. 10.1021/acs.jpcb.0c08192. [DOI] [PubMed] [Google Scholar]
  102. Feliciano M. A. M.; Gold B. Unique, yet Typical Oxyanion Holes in Aspartic Proteases. ACS Catal. 2020, 10, 14201–14209. 10.1021/acscatal.0c03624. [DOI] [Google Scholar]
  103. Coimbra J. T.; Neves R. P.; Cunha A. V.; Ramos M. J.; Fernandes P. A. Different Enzyme Conformations Induce Different Mechanistic Traits on HIV-1 Protease. Chem. - Eur. J. 2022, 28, e202201066 10.1002/chem.202201066. [DOI] [PubMed] [Google Scholar]
  104. Vasilevskaya T.; Khrenova M.; Nemukhin A.; Thiel W. Mechanism of proteolysis in matrix metalloproteinase-2 revealed by QM/MM modeling. J. Comput. Chem. 2015, 36, 1621–1630. 10.1002/jcc.23977. [DOI] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

ci2c01156_si_001.pdf (2.1MB, pdf)
ci2c01156_si_002.zip (2.9MB, zip)

Data Availability Statement

The cartesian coordinates and the force field parameters of all stationary points discussed in this paper are available in the Supporting Information as a ZIP file. The system was prepared using AMBER18 software. The AMBER package can be purchased at https://ambermd.org/index.php. H++ web server (http://biophysics.cs.vt.edu/H++) was used for the prediction of the protein titratable residues pKa. VMD v1.9.4 and PyMOL v2.3.0 were used for visualization and analysis of the system. VMD can be downloaded from https://www.ks.uiuc.edu/Research/vmd/ and PyMOL can be downloaded from https://pymol.org/. The VMD MolUP extension was used to prepare Gaussian input files and to visualize results from Gaussian output files. The molUP extension can be downloaded from https://biosim.pt/molup/. Gaussian 09 D01 was used for the QM/MM calculations. Gaussian software can be purchased from https://gaussian.com/.


Articles from Journal of Chemical Information and Modeling are provided here courtesy of American Chemical Society

RESOURCES