key: cord-0838194-wv7xqc9s authors: Carli, Matteo; Sormani, Giulia; Rodriguez, Alex; Laio, Alessandro title: Candidate Binding Sites for Allosteric Inhibition of the SARS-CoV-2 Main Protease from the Analysis of Large-Scale Molecular Dynamics Simulations date: 2020-12-11 journal: J Phys Chem Lett DOI: 10.1021/acs.jpclett.0c03182 sha: dd8ae65ec22e2068f412e924bc3cc157782fa814 doc_id: 838194 cord_uid: wv7xqc9s [Image: see text] We analyzed a 100 μs MD trajectory of the SARS-CoV-2 main protease by a non-parametric data analysis approach which allows characterizing a free energy landscape as a simultaneous function of hundreds of variables. We identified several conformations that, when visited by the dynamics, are stable for several hundred nanoseconds. We explicitly characterize and describe these metastable states. In some of these configurations, the catalytic dyad is less accessible. Stabilizing them by a suitable binder could lead to an inhibition of the enzymatic activity. In our analysis we keep track of relevant contacts between residues which are selectively broken or formed in the states. Some of these contacts are formed by residues which are far from the catalytic dyad and are accessible to the solvent. Based on this analysis we propose some relevant contact patterns and three possible binding sites which could be targeted to achieve allosteric inhibition. T he severe acute respiratory syndrome which broke out in December 2019 is caused by coronavirus 2 (SARS-CoV-2). 1,2 Its main protease (M pro or 3CL pro ) was the first protein of SARS-CoV-2 to be crystallized, in complex with a covalent inhibitor, in January 2020. 3 It is essential in the viral life cycle since it operates at least 11 cleavage sites on large viral polyproteins that are required for replication and transcription, 3, 4 so it is an attractive target for the design of antiviral drugs. 5 Since there is no known human protease having a cleavage specificity similar to the one of M pro , it may be possible to design molecules that do not interact with human enzymes. 3, 4 M pro is a homodimer. Each monomer has 306 residues and is composed of three domains. Domains I and II (residues 10−99 and 100−182, respectively) have an antiparallel β-barrel structure. The binding site of the substrate is enclosed between these β-sheets. 4 Domain III (residues 198−303) contains five αhelices and has a role in the regulation of the protein dimerization. 4 The two residues His 41 and Cys 145 form the catalytic dyad. The structure and way of functioning of the SARS-CoV-2 M pro are similar to those of the SARS-CoV M pro . 6, 7 This is expected, due to a 96% sequence identity between them. The most direct strategy to block the action of the M pro is through small molecules that directly interact with the catalytic site. The first in silico trials were made with covalent inhibitors known to be interacting with the catalytic site of SARS-CoV M pro , such as N3 3 or 11r 4 . Many efforts followed in the field of virtual screening. In this kind of studies, computational docking of millions of molecules is performed, and the behavior of the best candidates is usually then tested through MD simulation. 8−13 Another possible route that can be followed to stop the action of the M pro is allosteric inhibition. 14, 15 The functional definition of allosteric regulation implies the energetic coupling between two binding events. 16, 17 The binding of the allosteric ligands affects orthosteric pockets by altering protein dynamics, either through large-scale structural changes or through more subtle changes in correlated residue motions. 18, 19 Following the idea of conformational selection, 20 allosteric effectors will act as inhibitors by stabilizing configurations in which the access to the active pocket is at least partially closed. In short, the idea is to block the protease in one of its metastable conformations, in which the catalytic dyad cannot regularly operate, inhibiting in this way the whole protein functionality. This approach, at least in principle, has several advantages. First of all, it offers the possibility to drug sites far from the catalytic pocket, thus enlarging the chance to discover active compounds and to obtain non-competitive inhibition. If an allosteric site is identified and targeted, using this strategy, one can develop drugs which are highly specific since they do not bind in active sites, which are typically conserved in protein families. 21 Owing to these advantages, allostery has been established as a mechanism for drug discovery, for example to target Gprotein-coupled receptors (GPCRs) 22, 23 or protein kinases. 24−26 We here propose a strategy to identify candidate binding sites for allosteric inhibition which is fully based on the analysis of a long molecular dynamics (MD) trajectory. We analyze a 100 μs MD trajectory of the M pro generated in the D. E. Shaw Lab. 27 Our scope is to search for possible metastable states of the protease, namely configurations which do not change significantly on the scale of several tens of ns. These configurations are important for developing drugs for allosteric inhibition, since they are already (marginally) stable, and by designing a ligand which increases their stability, they can become kinetic traps. 21 These metastable states are searched by an approach, developed by us, which allows estimating the free energy landscape of a system in a high dimensional space. 28, 29 The local minima of the free energy, if deep enough, correspond to the metastable states, approximately the same that would be found by performing a much more expensive Markov State Modeling analysis. 30 The competitive advantage of our approach is that it allows performing the analysis in very high-dimensional spaces, taking into account at the same time several hundred different variables. This allows finding the free energy minima, and thus the metastable states, with no prejudice on their structure. In the case of the M pro , we carry out our analysis in two different spaces: the space defined by all the ψ backbone dihedrals of the protease, and the space defined by the contacts between pairs of residues which break or form during the dynamics. Both spaces consider the enzyme globally, not limiting the analysis to the catalytic dyad or to the binding pocket, which is essential to unveil possible allosteric states. Based on a characterization of global and local properties of these states, we propose a few possible targets which could serve as binding sites for drug-like compounds with the purpose of allosteric inhibition. We extract from the 100 μs MD trajectory 10 000 equally spaced frames, one every 10 ns. Since the enzyme is a homodimer, we consider the 20 000 total frames of the two monomer trajectories as a sample of the conformational space of a single monomer. However, the trajectories of the two monomers are considered and analyzed separately, in order to verify a posteriori whether the configurations they explore are similar or not. In both metric spaces in which we perform our analysis, we neglect the 10 residues at the C-terminus of the peptide, since they are highly mobile in both monomers and might introduce noise in the analysis. The two metrics are • the ψ-backbone-dihedral distance: 31 such distance between configurations t and t′ is defined as where ψ i,t is the value at time t of the ψ dihedral angle that involves the α-carbon of residue i of the monomer, index i runs between 1 and 296, and the notation ((·)) stands for 2π-periodicity within the brackets; • the contact-map distance, 31 restricted only to contacts which vary significantly during the simulation. To define these mobile contacts, we first compute the contact-map matrix C for each frame, restricted to residues 1−296. For each couple of residues ij, we first evaluate the distances between all the couples of heavy atoms, with one atom belonging to i and the second one belonging to j. C ij is then equal to σ(d min ), where d min is the smallest distance between the couples of atoms and σ is the sigmoidal function: σ = (1 − (d/r 0 ) 10 )/((1 − (d/r 0 ) 20 )), with r 0 = 4.5 Å. We consider as mobile the contacts which are completely formed (C ij > 0.8) in at least 5% of the frames and completely broken (C ij < 0.2) in at least 5% of the frames. Moreover, we neglect those contacts which have a value between 0.2 and 0.8 (i.e., close to r 0 ) in more than 50% of the frames. This procedure selects 155 relevant mobile contacts for the first monomer (m1) and 184 for the second (m2). Most of these contacts are in common, which is reasonable since the two monomers are chemically identical; the union of the two sets has 235 elements. Denoting by the set of mobile contacts of a monomer, the contact-map distance between configu- Our two metrics are both sensitive to local and global conformational changes in the peptide but capture different details: the ψ coordinates keep track of the changes in the protein backbone, whereas the mobile contacts metrics, instead, also keep track of the side-chain rearrangements while neglecting fluctuations around the completely formed or completely unformed contacts. The free energy landscape of each dataset is estimated following the procedure introduced in ref 32 . First of all, the intrinsic dimension (ID) of the manifold containing the configurations is calculated. 33 In the spaces of the ψ dihedrals we get an ID of 28 for m1 and of 26 for m2. In the spaces of the mobile contacts, we get an ID of 17 for both monomers. The free energy F of each configuration is then calculated using the PAk estimator, 28 which also provides an estimate of the uncertainty of F. The core of the approach is the calculation of the radius of the neighborhood in which the free energy can be considered constant within a fixed statistical confidence. Importantly, this algorithm requires the knowledge of the ID of the space in which the data points are lying, but it does not require knowing explicitly which variables define the reduced space. Finally, using Density Peak (DP) clustering 34 in its unsupervised variant, 32 we build a topography of the free energy landscape. We first find the free energy minima, and we assign all the frames to one of these minima according to the DP procedure. The set of configurations assigned to a single free energy minimum defines a free energy basin. Then, following ref 32, we find the saddle point between each pair of basins. The core set (CS) of a basin is the set of configurations whose free energy is lower than the free energy of the lowest saddle point of the basin. The described approach requires choosing the metric and a single metaparameter, the statistical confidence Z at which a basin is considered meaningful. A basin s is considered meaningful if (F ab − F a ) > Z(ε F a + ε F ab ) for all the basins b which share a border with a. Here, F a is the free energy minimum of basin a, ε F a is its uncertainty, F ab is the free energy of the saddle point between basin a and b, and ε F ab is its uncertainty. In our analysis Z is set to the value Z = 1.4, which corresponds to a confidence level of approximately 85%. This means that we expect to have nearly 15% of artificially split free energy basins. We verified that, by varying Z around this value, the description does not change significantlythe most populated free energy basins remain approximately unchanged. In the following analysis we call a state a set of configurations which belong to the core set of the same free energy basins according to both metrics. If, for example, a given basin number found using the dihedral metric is split in two different basins according to the contact metric, in our analysis we will consider two states. As a consequence, our states are structurally uniform The Journal of Physical Chemistry Letters pubs.acs.org/JPCL Letter according to both metrics. We consider in our analysis only states with a population of at least 8 core state configurations. With this criterion, we identify 11 relevant states in the trajectory of m1 and 7 in the trajectory of m2, for a total of 18 metastable states. First, we want to make sure that the metastable states detected analyzing the m1 and m2 trajectories separately are the same as if we run the algorithm on the merged 20 000 configurations. We check it in the case of the mobile contacts metric. We find that all the clusters involve only frames from the first monomer or only from the second. There is no relevant cluster that shares structures from both monomers, meaning that in terms of the contact map the configurations of m1 are different from the configurations of m2. Due to their chemical identity, in an ergodic simulation the configurations explored by the two monomers should be nearly identical. Therefore, the first important result of our analysis is that 100 μs of MD simulation is not sufficient to explore ergodically all the configuration space, as recently claimed also by Cocina et al. 35 This is also visible by looking at Figure 1a : most states are visited only two or three times. Consequently, the mean residence time cannot be meaningfully estimated. We instead compute the maximum residence time, considering it a proxy of the metastability of each state. These times are shown in the upper panel of Figure 1b and range from 0.20 to 16.07 μs. To quantify the accessibility to the catalytic site, we estimate the average solvent-accessible surface area (SASA) of the dyad and what we call the pocket doorway area (PDA), which quantifies the opening of the catalytic pocket from the position of four selected Cα carbons (see caption of Figure 1b) . The two quantities, presented in the middle and lower panels of Figure 1b , are in general quite correlated, although not in all the states. Indeed, contrary to PDA, SASA is sensitive to what happens in the direct proximity of the catalytic residues, while neglecting more macroscopic rearrangements of the catalytic pocket. Lastly, we characterize the local differences in the states by analyzing in detail their contact structure and their backbone arrangement. In the case of the mobile contacts, we analyze the intramonomer contacts which change significantly between at least 2 of the 18 states; furthermore, we also track the behavior of a few intermonomer contacts that might reflect some changes in the metastable states' catalytic activity. 6, 36 The contact structure of the selected states is summarized by the table in Figure 2a . As for the backbone, we analyze the ψ dihedral angles in the loops closing the cavity and a few other dihedrals which change significantly in the various states (see Supporting Information). As mentioned above, the catalytic dyad His 41 -Cys 145 is located in the pocket between the protein domains I and II. The access to this cavity is controlled by the flexible loop structures highlighted in Figure 2b Last, the β-sheet loop from Met 162 to Gly 170 delimits the cavity from the right in Figure 2b (we call it right loop); it is the least flexible, but it interacts with the N-finger of the other monomer and is crucial for shaping the substrate binding pocket. 38 All m2 states except m2:5 have the upper flap not tilted down and retracted with respect to the pocket, with contact Tyr 118 -Asn 142 almost always formed and contact Gly 138 -His 172 almost never formed. These two contacts are almost always mutually exclusive, with the exception of states m1:6 and m2:5, in which both contacts are formed at the same time. Another important difference among states, not related with the loops, is that dihedrals from Leu 227 to Asn 238 (bottom right in Figure 2b ) in all states of m1 are arranged in α configuration, so that an α-helix is Figure 2 . Selected intramonomer contacts and intermonomer contacts (marked with a star (*)). In the case of intermonomer contacts, the residue of the monomer which is excluded by the metric that defines a state is marked with a star (*). For each contact (columns), the average over the configurations of a given state is reported in the corresponding row. Such contacts are divided into two subgroups by a double vertical line: on the left those between residues belonging to the flexible loops which control the access to the binding pocket and on the right the other contacts. For readability, the entries take only three possible labels: 0 when the average over the configurations belonging to a state is <0.3, namely the contact is not formed; 1 when the average is >0.7, namely the contact is formed; n in all other case. Contacts whose label does not vary in any of the states of a given monomer are reported in light gray color. The picture shows a VMD 57 representation of monomeric M pro in state m1:1: on the left-hand side is the enzyme binding pocket, which encloses the catalytic dyad (in red); all other highlighted residue couples refer to the contact with the corresponding color in the table. The Journal of Physical Chemistry Letters pubs.acs.org/JPCL Letter formed and contact Tyr 239 -Leu 287 is always formed; in m2 such an α-helix structure is often defective. As for the contact between the N-finger and domain III (contact Gly 2 -Asn 214 ), in m2 it is often formed, while it is broken in most m1 states. We describe all the states in detail in the Supporting Information. Hereby, we focus on the most stable, the most open, and the most closed according to the SASA and PDA observables. Among the most closed states we mention m1:7, m1:9, m2:3, and m2:5. State m1:9 is very similar to m1:10 in its contact and backbone structure, with the exception of the left flap, which is more open in state m1:10. State m1:9 is also structurally similar to m1:7the only difference among the contacts is Pro 132 -Thr 196 , which is formed in m1:7 and not in m1:9, allowing the lower loop to be more flexible. In both, the upper flap is tilted downward, but the left flap backbone is open. In m1:9 the side chains of the residues in the loops surrounding the binding pocket are oriented toward the catalytic dyad, causing such state to rank among the lowest in SASA. State m1:7 ranks among the lowest in PDA and as the lowest in SASA; the reason lies in the side chains of the lower and left flaps, in particular of Thr 45 and Gln 189 , which form a contact and effectively close the access to the reactive site. State m2:3 ranks as the third lowest in both SASA and PDA. Cys 145 is not well covered, but on the other hand His 41 is less accessible than in most other states. As most m2 states, m2:3 has the upper flap bent upward and contact Gly 138 -His 172 not formed. The linker loop is not stretched, 41 . The upper flap, rather than being flat or tilted down, is oriented upward, causing a deformation in domain II which allows cont. Gly 138 -His 172 to be formed. Remarkably, like m1:9, state m2:5 is one of the few states with cont. Ala 285 -Ala 285 * not tightly formed. Our analysis shows that the accessibility to the catalytic dyad is reflected in the forming and breaking of few relevant contacts around the reactive cavity. For example, cont. Glu 47 -Leu 57 is not formed when the left flap is closed, a condition common to most states in which the catalytic dyad is not accessible. Similarly, the catalytic site (in particular Cys 145 ) is less exposed when the upper flap it tilted downward, i.e., when cont. Tyr 118 -Asn 142 is not formed. The druggability analysis software PockDrug 40 finds one pocket in correspondence of the residues of each of the two contacts (respectively called left pocket and upper pocket) and assigns to them a druggability probability of 0.68 ± 0.08 and 0.95 ± 0.03. Targeting these two regions with drug-like compounds, blocking the formation of the mentioned contacts, might prove a successful strategy for the inhibition of the catalytic activity. The distribution of SASA over all configurations in which contact Tyr 118 -Asn 142 is not formed is significantly shifted toward lower SASA values than in the cases in which the contact is formed (see Figure 3b ). Our analysis on the relevant contacts also unveils the presence of another interesting pocket far from the catalytic site, in the interface region between domains II and III (right-hand side of the table in Figure 2a) . The five relevant contacts in this region are Arg 131 -Thr 199 , Arg 131 -Asp 289 , Pro 132 -Thr 196 , Asp 197 , and Thr 198 -Asn 238 , Tyr 239 -Leu 287 . This region, which we call the distal pocket, has been previously identified and screened for docking and has been predicted as a potential druggable target. 41, 42 It has also been suggested as a target for allosteric inhibition of the catalytic activity. 43, 44 Coherently, the predicted druggability score is 0.65 ± 0.08. Experimental confirmation of the viability of the distal pocket as a target comes from crystallographic fragment screening. 42 Among the hits that were identified, three are particularly interesting. Fragment Mpro-x0390, classified as "high confidence", is in contact with atoms from five different residues, among which four are involved in the relevant contacts mentioned above. Fragment Mpro-x0464, also classified as "high confidence", is in contact with 11 residues, among which six are involved in the relevant contacts. Fragment Mpro-x1163, classified as "correct ligand but with weak density", is in contact with nine residues, among which five are involved in the relevant contacts. With a completely different approach, the database Pocketome 45 identifies for the coronavirus M pro a bindable pocket in the distal region, with two possible ligands (entry R1AB_SARS2_P6); this pocket includes residues Pro 132 , Thr 196 , Thr 198 , Asn 238 , and Tyr 239 , all involved in the five relevant distal pocket contacts. Alternatively, many other algorithms have been developed for the detection and scoring of druggable pockets. 46−52 We decided to further benchmark our findings by running the pocket detection software fpocket. 53 While for most structures the analysis does not detect any pocket in the distal region, the structures in the core set of state m1:9 display two pockets in contact with various residues in the distal region, even if with low druggability. Finally, we analyze the whole trajectory with the software MDpocket, 54 which quantifies in terms of a frequency grid the points involved in accessible pockets: the frequency value ranges from 0 if a point is never found along the trajectory in an open pocket to 1 if it is always found. The software assigns low values to the distal pockets: this suggests that the distal pocket is observed as a transient site, which makes its detection nontrivial. With the aim of verifying the presence of allosteric effects involving the distal pocket, we focus on the above-mentioned contacts in this region. We compute, e.g., the distribution of the PDA and of the SASA restricted to the frames in which the contact patterns are those of states m2:4 and m2:5 in the table in Figure 2a . Despite all considered residues being far from the binding pocket, the distributions of the PDA and of the SASA are sizably different in the two conditions. This suggests that if these five contacts could be forced to be formed or broken according to the desired pattern, e.g., by a drug-like compound, one could influence the PDA and the SASA, controlling indirectly the access to the reactive site. Comparing the table in Figure 2a and Figure 1b , a good candidate for allosteric drugging seems to be the contact pattern of state m1:9: (0,0,0,1,1). Interestingly, the PDA and SASA distributions obtained by selecting only the first three of the five contacts, namely (0,0,0), do not differ significantly from those with all five contacts involved (see e.g. Figure 3b ). We finally analyze the conservation of the residues involved in all the proposed contact patterns in the sequences of proteins belonging to the same family as M pro . We perform a multiple sequence alignment of our sequence (from PDB 6Y84 39 ) with all the sequences in the Pfam 55 seed of the corresponding family, Coronavirus endopeptidase C30 (Pfam entry PF05409). Similarly to ref 56, we find that many of the residues involved in the proposed target sites are conserved in all or most of the sequences (see Supporting Information) and furthermore all of them are conserved in the sequence of Human SARS coronavirus (SARS-CoV). In conclusion, our data analysis approach allowed us to identify 18 putative metastable states of the M pro of SARS-CoV-2. We characterized these states in terms of their structural differences, identifying some contacts which are selectively formed or broken in the different states. We believe that this analysis brings insight on the molecule's conformational changes which might prove useful for the design of pharmaceutical inhibitors. Our analysis approach is useful especially for understanding (and eventually controlling) the global dynamics of a protein, since treats the region of the catalytic cavity and any other part of the protein within the same framework. We stress that the same kind of analysis can easily be applied to any other candidate target proteins, due to its extreme generality. Based on this analysis we propose some possible target sites for the design of drug-like molecules, some of which directly in contact with the flaps regulating the access to the enzyme's active site, some located in the distal pocket at the interface between domains II and III of the monomers. We provide evidence of allosteric effects connected to such pocket, and we propose as drug target simply three contacts whose inhibition is correlated to a reduction in the access to the catalytic site; a more refined drug design could yield even stronger catalytic inhibition. We show that all three proposed target sites are comprised in pockets with high druggability score according to the software PockDrug. We find that all residues involved in the proposed target sites are conserved between the M pro of Human The Journal of Physical Chemistry Letters pubs.acs.org/JPCL Letter SARS-CoV and Human SARS-CoV-2 and that many of them are conserved in most sequences in the seed of the Pfam family to which they both belong. We interpret this as a comforting indication for the validity of our proposed targets. Moreover, the conservation of all such residues might suggest that mutations are unlikely, thus hopefully the displayed allosteric mechanisms are resistant to possible future mutations. A further possible interesting way to validate the viability of the predicted pockets as potential drug targets, especially of the distal pocket, would be analyzing the effect of mutations in that region on the catalytic activity. Finally, a dynamical docking simulation would be the next step to assess our findings from a more accurate biochemical standpoint. To summarize, the added value provided by our analysis is two-fold. First, and most importantly, we provide the structure of the state which should be targeted for drug design. This structure does not coincide with the crystallographic structure, and not even with the most likely configuration observed in the MD simulationindeed some crucial tertiary contacts which are formed in the crystal are not formed in the structure we propose, and these contacts form and break dynamically along the trajectory. Available bioinformatic tools for searching druggable cavities do not normally provide hints on the structural rearrangement which should be induced by the drug to modify the properties of the catalytic cavity, as we are instead able to do. In the second nontrivial insight provided by our analysis, it unveils high mobility in the distal pocket region, excluding the presence of relevant conformational changes coupled with the accessibility of the catalytic dyad in other sites. Even if we cannot exclude that allosteric effects may arise even from other pockets, our findings suggest prioritizing these targets among the wealth of putative binding sites found by automatic scanning. The structures of the putative metastable states described in this work are available in Supporting Information for independent structural analysis and for targeted drug design which, we hope, will be performed by groups with the appropriate competencies. The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpclett.0c03182. Definition of PDA; mobile and relevant dihedrals analysis; detailed structural description of the metastable states; mutation and conservation analysis of relevant residues within the same protein family (PDF) Structures of the putative metastable states described in this work (ZIP) A pneumonia outbreak associated with a new coronavirus of probable bat origin A new coronavirus associated with human respiratory disease in China Structure of Mpro from SARS-CoV-2 and discovery of its inhibitors Crystal structure of SARS-CoV-2 main protease provides a basis for design of improved a-ketoamide inhibitors An overview of severe acute respiratory syndrome-coronavirus (SARS-CoV) 3CL protease inhibitors: Peptidomimetics and small molecule chemotherapy Structure of coronavirus main proteinase reveals combination of a chymotrypsin fold with an extra α-helical domain The crystal structures of severe acute respiratory syndrome virus main protease and its complex with an inhibitor Peptide-like and small-molecule inhibitors against Covid-19 Identification of potential molecules against COVID-19 main protease through structure-guided virtual screening approach Virtual screening of approved drugs as potential SARS-CoV-2 main protease inhibitors Prediction of the SARS-CoV-2 (2019-nCoV) 3C-like protease (3CL pro) structure: virtual screening reveals velpatasvir, ledipasvir, and other drug repurposing candidates Putative inhibitors of SARS-CoV-2 main protease from a library of marine natural products: A virtual screening and molecular modeling study Virtual screening and repurposing of FDA approved drugs against COVID-19 main protease Allostery without conformational change -A plausible model Dynamic allostery: Linkers are not merely flexible Allosteric proteins and cellular control systems Allostery: an illustrated definition for the second secret of life Introduction to Protein Ensembles and Allostery Emerging computational methods for the rational discovery of allosteric drugs Conformational selection in protein binding and function Allostery in disease and in drug discovery Trends in GPCR drug discovery: new agents, targets and indications Allosteric modulators of GPCRs: a novel approach for the treatment of CNS disorders Inhibition of p38 MAP kinase by utilizing a novel allosteric binding site Allosteric small-molecule kinase inhibitors Allosteric targeting of receptor tyrosine kinases Molecular Dynamics Simulations Related to SARS-CoV-2, D. E. Shaw Research Technical Data Computing the Free Energy without Collective Variables Explicit Characterization of the Free-Energy Landscape of a Protein in the Space of All Its Cα Carbons Which similarity measure is better for analyzing protein structures in a molecular dynamics trajectory? Automatic topography of high-dimensional data sets by non-parametric Density Peak clustering Estimating the intrinsic dimension of datasets by a minimal neighborhood information Clustering by fast search and find of density peaks Sapphire-Based Clustering Identification of Novel Inhibitors of the SARS Coronavirus Main Protease 3CLpro Structural and evolutionary analysis indicate that the sars-COV-2 mpro is a challenging target for small-molecule inhibitor design Computational Studies of SARS-CoV-2 3CLpro: Insights from MD Simulations RCSB PDB -6Y84: SARS-CoV-2 main protease with unliganded active site (2019-nCoV PockDrug-Server: a new web server for predicting pocket druggability on holo and apo proteins Elucidation of cryptic and allosteric pockets within the SARS-CoV-2 protease Computational Analysis of Dynamic Allostery and Control in the SARS-CoV-2 Main Protease SARS-CoV-2 Simulations Go Exascale to Capture Spike Opening and Reveal Cryptic Pockets Across the Proteome Pocketome via Comprehensive Identification and Classification of Ligand Binding Envelopes A computational procedure for determining energetically favorable binding sites on biologically important macromolecules POCKET: a computer graphics method for identifying and displaying protein cavities and their surrounding amino acids POVME 3.0: Software for Mapping Binding Pocket Flexibility Efficient Characterization of Protein Cavities within Molecular Simulation Trajectories: trj_cavity A Tool for Analysis of Transient Binding Pockets in Proteins Transient pockets on protein surfaces involved in protein-protein interaction Allosteric Communication Networks in Proteins Revealed through Pocket Crosstalk Analysis MDpocket: open-source cavity detection and characterization on molecular dynamics trajectories The Pfam protein families database in 2019 Structural basis of SARS-CoV-2 3CLpro and anti-COVID-19 drug discovery from medicinal plants VMD: Visual molecular dynamics The Journal of Physical Chemistry Letters pubs.acs.org/JPCL Letter