key: cord-0315803-oeiz7zbo authors: Evans, Daniel J.; Yovanno, Remy A.; Rahman, Sanim; Cao, David W.; Beckett, Morgan Q.; Patel, Milan H.; Bandak, Afif F.; Lau, Albert Y. title: Finding Druggable Sites in Proteins using TACTICS date: 2021-02-21 journal: bioRxiv DOI: 10.1101/2021.02.21.432120 sha: ccfb186427302586f3290611f9236532bc373c15 doc_id: 315803 cord_uid: oeiz7zbo Structure-based drug discovery efforts require knowledge of where drug-binding sites are located on target proteins. To address the challenge of finding druggable sites, we developed a machine-learning algorithm called TACTICS (Trajectory-based Analysis of Conformations To Identify Cryptic Sites), which uses an ensemble of molecular structures (such as molecular dynamics simulation data) as input. First, TACTICS uses k-means clustering to select a small number of conformations that represent the overall conformational heterogeneity of the data. Then, TACTICS uses a random forest model to identify potentially bindable residues in each selected conformation, based on protein motion and geometry. Lastly, residues in possible binding pockets are scored using fragment docking. As proof-of-principle, TACTICS was applied to the analysis of simulations of the SARS-CoV-2 main protease and methyltransferase and the Yersinia pestis aryl carrier protein. Our approach recapitulates known small-molecule binding sites and predicts the locations of sites not previously observed in experimentally determined structures. The TACTICS code is available at https://github.com/Albert-Lau-Lab/tactics_protein_analysis. Designing allosteric modulators requires a knowledge of which parts of the protein can bind ligands with high affinity. This is not a trivial task, especially in cases where the binding site can be classified as "cryptic". A cryptic site can be broadly defined as a site that is apparent when a ligand is bound to it but difficult to detect when no ligand is present (1) . One proposed mechanism for cryptic site formation is the "induced fit" mechanism, in which the ligand causes the protein to adopt a conformation that it would otherwise not adopt. Another proposed mechanism is "conformational selection", in which ligands bind and stabilize pocket-containing conformations that are rare (but not nonexistent) in unbound proteins. It is likely that many cryptic sites undergo a combination of these mechanisms (2) , (3) , (4) . The exact definition of a cryptic pocket is subject to debate; a particularly difficult question involves the classification of sites that are visible in some unbound structures but are hidden in other structures. Some classify these sites as cryptic (5) , but this is not universally accepted (1) . Finding a cryptic site on a protein does not guarantee that the protein is druggable. For a cryptic site to be of interest for drug design, it must also be allosteric: the binding site's conformation must influence the protein's function (6) . Despite this caveat, the ability to find cryptic pockets opens up an additional promising avenue for drug design. A variety of algorithms have been developed to study binding sites (both cryptic and non-cryptic). Some algorithms characterize the behavior of known pockets. Many of these algorithms analyze molecular dynamics (MD) simulations, including EPOCK(7), POVME (8) , (9) , (10) , and TRAPP (11) , (12) , (13) . These algorithms are useful for characterizing the behavior of known pockets, but their reliance on prior knowledge of binding pocket locations limits their use on poorly understood proteins. Other algorithms predict the locations of previously unknown pockets. Some take a static structure (such as a PDB file) as input, such as LIGSITE (14) , KVFinder (15) , COACH (16) , COFACTOR (17) , ConCavity (18) , FPocket (19) , GalaxySite (20) , ghecom (21) , DoGSite (22) , and FTMAP (23) . This approach can sometimes provide druggability scores of predicted pockets, as seen in PRANK (24) and DoGSiteScorer (25) . Using a single structure reduces computational time but makes it difficult to find cryptic sites that are not present in all conformations. Using these programs on MD trajectories would require users to either run them on each frame (which is computationally expensive) or select individual frames to be analyzed (which increases the time and expertise required). Other algorithms use MD simulations; these algorithms consider multiple protein conformations and thus are sometimes able to find cryptic sites. One variety of MD procedure is "mixed-solvent MD", in which proteins are simulated in a system containing both water and smallmolecule probes (26) , (27) , (28) , (29) , (30) , (31) . The probes may interact with the binding sites on the protein, revealing the sites' locations. Probes can open pockets that operate by the induced fit mechanism. The SWISH procedure (2) , (32) combines the use of probes with enhanced sampling. A procedure somewhat similar to mixed-solvent MD "wraps" the protein in a coating of ligand before MD simulation (33) . While these procedures are extremely powerful, they require specialized simulations. This prevents their use on precomputed trajectories. Other algorithms that find pockets or binding sites take the results of precomputed MD simulations that may not include probes. EPOS BP (34), trj_cavity (35) , MDPocket (36) , D3Pockets (37) , and NanoShaper (38) , (39) begin with geometry-based pocket finders that neglect physicochemical properties; then they analyze the newly found pockets. Pocketron (40) begins with NanoShaper, then performs detailed analyses with an emphasis on characterizing allosteric regulation. FTDyn (41) docks small probes onto frames from precomputed MD trajectories in order to find bindable regions. The exposon procedure (42) uses correlated changes in solvent-accessible surface area to find cryptic pockets. Each of these procedures can be quite useful, but all have limitations. Some do not consider physicochemical properties (such as those examined in fragment docking), leaving users uncertain about the extent to which geometrically reasonable pockets will bind ligands. Other programs run on each frame of the trajectory, making it time consuming to run the software and interpret the output for long simulations. Users can mitigate this by selecting individual frames for analysis (e.g., through clustering), but this increases the workflow's complexity. By incorporating the abilities to both select individual frames for analysis and consider physicochemsitry into a single software package, TACTICS fulfils a unique niche. CryptoSite(5) takes a different approach by calculating its own MD trajectories without probes within the algorithm. The JEDI procedure(43) also includes probe-free MD simulations, with the JEDI score used as a collective variable to enhance sampling. However, the procedure requires users to define a region of interest, limiting the applicability to proteins lacking any information about cryptic site locations. Each of these strategies has advantages and disadvantages. Mixed-solvent MD probes can open cryptic pockets that operate by the induced-fit mechanism. However, mixed-solvent MD simulations must avoid undesirable interactions between probe molecules. Strategies for this include adding repulsion between probes(30), choosing probes that do not aggregate (26) , and simulating each probe type separately (27) . MD simulations without probes avoid the issue of probe-probe interactions, but they risk missing cryptic sites that require induced fit. Because many cryptic sites' opening mechanisms include conformational selection (in addition to induced fit), it is possible that probe-free simulations will expose part (if not all) of the existing cryptic sites. Developers of pocket-finding software must also consider whether their approach should analyze existing MD simulations or run MD within the algorithm. The latter strategy allows software users who are unfamiliar with MD to explore conformational space without the effort of learning to set up their own simulation. However, automatic system generation may prevent users from studying proteins that require specialized or nonstandard conditions. Software that takes precomputed MD trajectories assumes that users have the knowledge necessary to set up and run an MD simulation. However, it gives users more flexibility in the types of simulations that can be studied. Software that analyzes precomputed trajectories also permits trajectories calculated for other studies to be repurposed as TACTICS input for pocket-finding research. The ability of machine learning (ML) to find trends in complex data makes it a promising technique for the nontrivial task of predicting small molecule binding sites. A few algorithms that find binding sites use machine learning(16),(5), but most do not. Machine learning has the potential to substantially improve binding site prediction, including prediction of "cryptic pockets" that are not present in crystal structures (1) . Thus efforts to apply ML and other cutting-edge computational techniques to the prediction of binding sites can have substantial impacts on the process of designing drugs. We designed and implemented a machine learning algorithm that finds druggable pockets within MD trajectories. The algorithm is named TACTICS ("Trajectory-based Analysis of Conformations To Identify Cryptic Sites"). As described below, TACTICS correctly identifies several known or predicted ligand binding sites in SARS-CoV-2 proteins. It also predicts several novel binding sites in the SARS-CoV-2 2'-O RNA methyltransferase and the Yersinia pestis aryl carrier protein. These predicted sites provide new opportunities for drug development. The TACTICS software is freely available so that it can be used to find druggable sites in other proteins. The ML model used in TACTICS was trained using a database of crystal structures. Each protein in the database has a single "holo" structure with a ligand bound in the cryptic site and up to 50 "apo" structures without a ligand in the site. These ensembles of structures are designed to imitate MD trajectories in which a few frames have the cryptic site that is absent in most frames. Database construction began with a reconstruction of the Cryptosite Database (5) . Each cryptic site in this database has an "apo" structure with the site hidden and a "holo" structure with a ligand in the cryptic site. The original database contains the AMPAR LBD twice. Both entries were removed from the reconstructed database. After the Cryptosite Database was reconstructed, it was extended using a procedure similar to that used in Beglov et al.(3) . For each database protein, the PDB's 95% sequence identity cluster was downloaded. This provided additional apo structures to be compared with each holo structure. This is important because the database is designed to mimic MD trajectories in which many frames may have the pocket region in conformations different from the bound conformation. While the percentage of time that a pocket is open in an MD trajectory is likely to be systemdependent, it is reasonable to suppose that ensembles of conformations generated by MD would frequently have high ratios of apo-like structures to holo-like structures. The ML training database was designed to mimic this. MODELLER (44) was used to match the holo structure's residue numbers with those of each structure in the cluster. PyMOL (45) was used to select residues in the holo structure within 9 Å of the co-crystallized ligand; these residues were used to align the holo structure with each extended-database structure. All extended-site structures with a ligand other than water or a metal within 5 Å of the holo ligand were removed from the database. This was done because no apo structures should have the cryptic site bound. Our procedure assumes that in the Cryptosite database, cryptic sites are closed (or otherwise unbindable) in all unbound ("apo") structures. However, some An additional shortcoming is presented by the variability in the number of apo structures per holo structure. Some proteins have been crystallized more often than others, resulting in a biased distribution of crystal structures. To reduce the severity of this issue, a maximum of 50 apo structures per holo structure was used. This does not completely eliminate the issue, as proteins with fewer than 50 structures will be undersampled. However, it reduces the issue's severity. A machine learning model was constructed to find cryptic sites. The model predicts the likelihood that a given residue in a given structure is part of an accessible cryptic site. The model was designed to return False for residues that are not part of a cryptic site, and for residues that are part of a cryptic site that is not accessible in the input structure. It is only intended to return True when the input structure contains an accessible cryptic site. The model uses several features. One feature is the ConCavity score for each residue (18) . ConCavity is a pocket detection algorithm that considers the geometry of the input structure and (if allowed by the user) the sequence conservation of each residue. It has been used in other multi-step pocket finding algorithms(5), (16) . ConCavity has the option of only using the structural information (ignoring sequence conservation); TACTICS uses this option. Including sequence conservation in ConCavity may improve TACTICS's results; this might be added to future versions of the software. ConCavity's structure-based approach starts by using Ligsite (14) to determine which points near the protein are part of geometric pockets. ConCavity then clusters these points into binding sites and uses these sites to assign scores to each residue. It considers only a single structure. Another feature in TACTICS's ML model is the change in ConCavity score between the current structure and a reference structure. This feature can only be calculated from conformational ensembles such as MD trajectories, so it sets TACTICS apart from algorithms that work on a single structure. The feature was included to increase TACTICS's ability to find cryptic sites. It is important because flexibility has been observed to be associated with cryptic site locations(3). The reference structure must be provided by the software user. It is expected that the reference structure will usually be the first frame of the MD trajectory, but another structure can be chosen (as long as it is aligned to the MD trajectory and has identical residue numbering). A druggable cryptic pocket should include more than one residue. The model considers this by including as a feature the average ConCavity score of the residues to either side (in the sequence) of the residue in question. It might be useful to consider spatially close residues that may not be close on the sequence; future versions of TACTICS may add this capability. The model also considers the change in sequence neighbors' ConCavity scores between the current structure and the reference structure. The model includes the change in α-carbon position between the current structure and the reference structure. It also includes the neighbors' average change in α-carbon position. The values of these features were calculated for each residue of each structure in the database. The input trajectory must be aligned so that the protein's center of mass does not move over the course of the simulation. If this is not done, then the entire protein's motion will incorrectly be included when calculating changes in α-carbon positions. Alignment must be done by the user prior to running TACTICS. 20% of the data was randomly selected and set aside as a test set for evaluating the model's performance. Testing data was not used to train the model. Model performance was carried out using the area under the receiver operating characteristic (abbreviated AUROC or ROC-AUC). The statistics behind the AUROC merit a brief discussion. As the confidence threshold required for the model to classify an item as True is varied, the number of true and false positives also varies. The receiver operating characteristic (ROC) is a graph of true positive rate vs. false positive rate as the confidence threshold is varied. The area under the ROC curve (AUROC) is the fraction of the time that the model will assign higher confidence to a randomly chosen positive sample than to a randomly chosen negative sample (46) . The advantage of the AUROC over the (conceptually simpler) percent accuracy metric is that the latter is biased if the dataset has unequal numbers of positive and negative samples. The AUROC has been used to quantify the performance of other pocket-finding algorithms(5), (16) ; it was selected to quantify the performance of TACTICS's ML model. Scikit-learn(47) was used to train a random forest model on the 80% of the data that was not set aside. Three-fold cross-validation was used to select hyperparameter values.After training the model on the training dataset, the AUROC score was calculated on the test dataset. A result of 0.856 was obtained. This AUROC was calculated on a dataset of crystal structures, while the ultimate application of the model is to MD trajectories. Our use of this score assumes that the conformations in the crystal structure dataset are representative of what would be seen in MD simulation data. The extent to which this assumption is true impacts the extent to which our AUROC can be assumed to represent performance on MD data. Nevertheless, the score is useful as a rough estimate of whether the model is likely to be useful. After calculating the AUROC, the model was re-trained on the entire database (composed of both the training and testing sets). As useful as the ML model is, it was observed to predict an unrealistic number pf This is done because cryptic pockets are expected to open and close; a residue that always has a high score is unlikely to be part of a cryptic pocket. The above procedure generates a set of residues that arepredicted to be part of cryptic pockets. These residues are clustered based on their proximity to each other; the result is a set of predicted binding regions. To investigate which predicted sites are likely to be druggable, fragment docking is performed on each site. AutoDock Vina (50) is used for the docking; the fragments are the 16 molecules used by FTMap (41) and Cryptosite (5): acetaldehyde, acetamide, acetone, acetonitrile, benzaldehyde, benzene, cyclohexane, dimethyl ether, ethane, ethanol, isobutanol, isopropanol, methylamine, phenol, urea, and N,N dimethylformamide. Docking is performed in a box surrounding the predicted binding site; the default is to add 8 Å in each dimension. As in Cryptosite, each residue's "dock score" is calculated by counting the number of times that a docked fragment ligand is within 3.5 Å of the residue. The dock score is used here to verify and expand upon the ML predictions. It is expected that predictions are more likely to be accurate when the residues with high ML scores are in the same region as residues with high dock scores. Substantial differences between docking and ML results may indicate inaccurate ML predictions. They may also represent cases where docking has found additional residues that are important for binding but were missed by the ML model. Figure 1A shows a flowchart of the TACTICS algorithm. To facilitate the process of interpreting TACTICS's results, a procedure was developed for visualizing the results in PyMOL (45) . Residues that are highlighted by the ML algorithm are shown as sticks. Dock scores are shown as b-factor putty. The "search spaces" where docking ligands are allowed to be are drawn as boxes. Each time that the code runs, it creates a file with PyMOL commands to display the results. Figure 1B shows typical TACTICS output. Boxes indicate the edges of regions that fragments are allowed to dock in. Residue coloring and thickness are determined by fragment dock scores. Black sticks indicate residues with high ML scores. The SARS-CoV-2 main protease (nsp5) model is based on PDB ID 6Y2E(51). Hydrogens were added using CHARMM (52) . The system was neutralized and solvated with ~150 mM KCl for a total system size of 126,736 atoms. A weak center of mass restraint was applied to N, CA, and C atoms of residues 19-21, 36-38, and 80-82 during equilibration. No restraints were used during production simulations. The initial atomic model of the SARS-CoV-2 2'-O RNA methyltransferase (nsp10/nsp16 heterodimer) is based on PDB ID 6W61. It was prepared in CHARMM-GUI (53) . Missing amino acids in nsp10 were added using the Modloop server (54) . All crystallographic waters and coordinated Zn 2+ ions were included. The system was neutralized and solvated with ~150 mM KCl for a total system size of 65,958 atoms. Zn 2+ coordinating cysteines were deprotonated but retain a neutral charge using a custom force field topology. This was done to mimic the geometry of the coordination bond without introducing a nonexistent negative charge on the sulfur atoms. To prevent the dislodging of the Zn 2+ ion from its binding cavity, a 10 kcal/mol/Å 2 harmonic restraint was applied between the Zn 2+ ion and each coordinating cysteine residue. A 1 kcal/mol/Å 2 harmonic potential position restraint is applied to Cα atoms in core regions of the nsp10/nsp16 dimer (residues 70-72, 86-88, and 171-173 in nsp16) to keep the protein centered within the box. These restraints were used during both equilibration and production. All REST2 simulations (55) were carried out in NAMD2.13 (56) . The CHARMM36 force field was used for protein and ions(57) , (58) , (59) while the TIP3P model was used for water (60) . The nsp10/nsp16 dimer and the protease were subjected to energy minimization and a 1 ns equilibration under NPT conditions at 1 atm and 300 K. All REST2 simulations used an NPT ensemble at 1 atm, with the effective temperature ranging from 300 to 350 K, spanning across 12 replicas for each system. Solute tempering is applied to protein atoms only and exchange attempt frequency is set to every 2 ps. A 1 ns pre-production REST2 simulation was carried out to initialize exchange probabilities. Each replica was simulated for 50 ns to obtain exchange probabilities between 20% and 30%. For equilibration, pre-production, and production simulations, a timestep of 2 fs was used. Periodic boundary conditions were used. All electrostatic interactions were computed using the particle-mesh Ewald (PME) algorithm, and short-range, non-bonded interactions were truncated at 12 Å. The Aryl Carrier Protein (ArCP) simulation systems are based on PDB structures 5TTB (apo)(61), 2N6Y (holo) (62) , and 2N6Z (loaded) (62) . The loaded and holo structures each have a phosphopantetheine (PP) "arm" attached to Ser52 and the loaded form harbors a salicylate at the end of this arm; these moieties were parameterized using CGenFF (63) , (64) , (65) . Each protein (apo, holo, and loaded) was minimized in a vacuum before being solvated in a water box whose sides are 60Å wide. 17 sodium atoms and 18 chlorine atoms were added, neutralizing the system and providing a concentration of 150 mM NaCl. Each system underwent three rounds of 60,000 ps equilibration at 300K using Langevin dynamics. Equilibration and production simulations used an NPT ensemble. For each system, the simulation was run in NAMD using the CHARMM36 force field. The simulations were carried out at 300 K and 1 atm with periodic boundary conditions. A timestep of 2 femtoseconds was used, and a trajectory snapshot was written every 500 timesteps (1 ps). A total of 200 ns of simulation data was obtained, and the latter 100 ns was used for TACTICS analysis. While default values exist for TACTICS parameters, these values may not be optimal for all systems. Poor parameter choices may lead to TACTICS not finding few (if any) pockets or finding unreasonably many pockets. Finding the best parameter values may require running TACTICS with several values and evaluating which run has the most reasonable results. This introduces an unfortunate subjectivity into the TACTICS workflow. But eliminating all user-defined parameters would risk compromising the software's ability to generate useful results across a diverse range of input protein systems. The tunable parameters are as follows. num_clusters is the number of frames from the trajectory that will be selected for TACTICS analysis. Frame selection uses k-means clustering. The number of frames displayed in the TACTICS output may sometimes be less than num_clusters; this occurs if there are frames with no pockets. Values of roughly 3-15 are expected to be reasonable for most situations; it is recommended to start with a value near 8. ml_score_thresh is the minimum ML score needed to be considered when clustering residues into binding sites. The default value is 0.8; substantial variation may be needed but values of 0.5-0.9 should work for most systems. ml_std_thresh is the maximum standard deviation between frames that a residues ML score must have for the residue to be considered when clustering residues into binding sites. It may also benefit from substantial variation; the default value is 0.25 and values from roughly 0.05 to 0.4 are suggested. It should be emphasized that the trajectories should be aligned before running TACTICS. This removes the motion of the entire protein that would affect the ML model's change-in-position features. More information on how to run TACTICS can be found in the software's README file. For the main protease, we used num_clusters=8, ml_score_thresh=0.6, and ml_std_thresh=0.09. For the methyltransferase, we used num_clusters=8, ml_score_thresh=0.6, and ml_std_thresh=0.1875. For the ArCP, we used num_clusters=13, ml_score_thresh=0.6, and ml_std_thresh=0.15. Each trajectory was aligned using MDAnalysis (49) , (48) . Using TACTICS to analyze large datasets is computationally intensive. Thus we divided the data into chunks and ran TACTICS on each chunk. For each SARS-CoV-2 protein, we ran TACTICS on the first 20 ns, the next 20 ns, and the last 10 ns. Analyzing the ArCP simulations was less computationally demanding, so these simulations were not divided into chunks. Our SARS-CoV-2 methyltransferase simulations included 2 zinc ions. Our ArCP simulations included a covalently attached phosphopantetheine (PP) arm. These atoms can be expected to impact ligand binding to any nearby sites by, for example, steric hindrance. Ignoring these non-standard components during TACTICS analysis might introduce inaccuracies. Including them in TACTICS analysis required special consideration; the procedure is expected to be useful for other proteins with nonstandard components. TACTICS requires each residue name be one of the standard amino acids. Additionally, each residue must have an alpha carbon (atom name CA). TACTICS deletes atoms that do not meet these conditions. A relatively straightforward way of getting TACTICS to recognize coordinated zinc atoms is to change their residue names from Zn to an amino acid and change their atom names to CA. Based on how TACTICS works (including ConCavity and AutoDock Vina, which it runs), we believe that approximating coordinated zincs as alpha carbons is reasonable. Because both methyltransferase zinc atoms coordinate with multiple residues in the protein, we expect the zincs' electrostatic effects to be minor enough that ignoring them does not dramatically impact our results. (Atoms with strong hydrophobicity or hydrogen-bond formation favorability would impact fragment docking.) The alphacarbon approximation acknowledges that zinc atoms create excluded volume; thus, making the approximation is preferable to letting TACTICS delete the zincs. In the case of the ArCP, the loaded and holo forms have a PP arm covalently attached to a serine residue. The modified residue already has an alpha carbon, but it must be renamed to an amino acid (e.g. Ser) before TACTICS analysis. We also had to rename an atom with atom name "CS"; the atom is supposed to be carbon but TACTICS attempted to parse the "CS" as cesium. To demonstrate the capabilities of TACTICS, we applied the algorithm to three proteins. Two of them are SARS-CoV-2 proteins (main protease and methyltransferase). TACTICS finds previously reported binding sites and novel sites in these proteins. The third protein (aryl carrier protein) is relatively small and has significant posttranslational modifications; it illustrates the versatility of TACTICS for unusual systems. The SARS-CoV-2 proteins were simulated using REST2(55), while the aryl carrier protein was simulated using equilibrium MD. This demonstrates TACTICS's ability to analyze conformational ensembles generated through a variety of methods. The SARS-CoV-2 main protease (M-pro; nsp5), is responsible for cleaving the polyprotein produced from translation of the viral RNA genome. Specifically, M-pro cleaves the peptide bond after glutamine residues most frequently occurring in the recognition sequence Leu-Gln-(Ser, Ala, Gly)(51) (66) . Peptide cleavage by M-pro occurs by a catalytic dyad of cysteine and histidine residues in the active site (67) . While cleavage after glutamine residues has been observed for the main proteases of other coronaviruses, including SARS-CoV(68) and MERS(69), it has not been frequently observed in human proteases (66) . From a therapeutic perspective, this reduces the probability of off-target effects and makes M-pro a promising target for drug design. Most notably, the binding of peptidomimetic inhibitors to the proteolytic active site has been explored as a means to inactivate the protease (70) . By mimicking the geometry of the substrate, these inhibitors become "trapped" in Figure S1 . To the left of the crystallized ligand in the figure, TACTICS identifies additional residues as druggable. Many of these residues were also highlighted in the Folding@Home results (74) , underscoring the likelihood that these residues are bindable. Because TACTICS was designed with an emphasis on finding cryptic sites and minimizing false positives, it uses ml_std_thresh to filter out residues whose scores don't change. Thus it is not obvious that TACTICS would be able to find active sites (which might be expected to remain visible throughout the simulation). However, enzymes can be highly flexible, particularly when unbound (82) . Additionally, ml_std_thresh was set to a low value for the main protease. Thus TACTICS's ability to find the active site is not entirely unexpected. This success is an example of the software's versatility. TACTICS also identifies a known allosteric site in the main protease. A ligand is crystallized at this site in PDB ID 5RGJ(77); while the crystal structure has a ligand bound to each protomer of the dimer, the two ligands are extremely close to each other. This ligand has been shown to decrease the protease's activity (73) TACTICS output for a frame in which a known allosteric site is found. This site binds a ligand called x0425 or Z1401276297. The x0425-binding allosteric site is found in the teal and green residues at the top of the image. In this image, the dimer interface is also high-scoring (red residues in center). D. The same as (C), with the known inhibitor called x0425 or Z1401276297 added from PDB ID 5RGJ. E. TACTICS output for a frame in which a known binding site at the dimer interface is found. F. The same as (E), with the known inhibitor x1187 or Z2643472210 added from PDB ID 5RFA. from 5RGJ. Thus it is plausible that larger ligands could bind to this site. The dimer interface is another potential drug target identified by TACTICS. Dimerization is expected to be important to protease activity (83) , so ligands that bind at the dimer interface and disrupt dimerization are likely candidates for drugs. PDB ID 5RFA has two molecules of the compound x1187 bound at the dimer interface (77) . Each molecule interacts with both protomers. Experimental results suggest that x1187 inhibits dimerization and protease activity (73) , confirming the importance of the dimer interface. Results on x1187 show the complexity of the dimer interface's behavior; x1187 binds the dimer form in the crystal structure (77) but inhibits dimerization in experimental tests (73) . TACTICS correctly predicts that the region where x1187 binds is druggable. This is shown in Figure 2 E, F and Supplemental Figure S3 .. The MD simulation contains C-terminal residues that are missing from the crystal structure; some of these residues would sterically clash with the ligand in its crystallized pose. It is plausible that these residues could move to accommodate the ligand in an induced-fit mechanism. Despite the uncertainty regarding the conformational and ligand-binding characteristics of this site, the fact that TACTICS identifies the site as druggable demonstrates the algorithm's ability to find experimentally validated binding sites. The x1187-binding site is close to where AT7519 has been crystallized (78) . In some frames where TACTICS identifies the x1187 site, the predicted site location also includes some residues involved in binding AT7519. TACTICS is less successful at identifying AT7519-binding residues than x1187-binding residues. AT7519 shows relatively small antiviral activity (78) . The weak antiviral activity and incomplete identification by TACTICS raise the possibility that the site could be less favorable to ligand binding than other sites. During the COVID-19 pandemic, one of the many promising druggable targets is the In the apo-ArCP, TACTICS predicts two binding sites. One predicted site involves helices a1 and a2; it is shown in Figure 5 A, B and Supplemental Figure S9 . Potentially bindable residues on a1 include Q23, R27, and E31 while key residues on a2 include R54, R57, and W61. As seen in the slice image, the site is divided into Figure S10 ), the site includes a1 residues H17, D20, and Y21 along with a4 residues M87, L88, and E93. In another conformation (Figure 5 E, F and Supplemental Figure S11 ), the site includes a1 residues R16, H17, and D20 along with a4 residues S91, P92 and loop2 (between a2 and a3) residue Y67. In a third conformation (Figure 5 G, H and Supplemental Figure S12 ), the site includes a1 residues D14, H17, A18, and Y21 along with a4 residues N84, Q85, and L88. In holo-ArCP, TACTICS predicts a binding site at the N and C termini. This site is shown in Figure 6 A, B and Supplemental Figure S13 . The site has the same location as the "Apo 2" site discussed above; dock scores and conformational flexibility are very similar between the apo and holo results. TACTICS also predicts a binding site located primarily on loop1 (between a1 and a2). It includes residues L39, H40, S43, and A48. In the loaded-ArCP, TACTICS predicts a binding site that includes helix a4 and loop2 (between a2 and a3). This site is shown in Figure 6 C, D and Supplemental Figure S14 . Potentially important residues on helix a4 include L86, M87, and R90. Potentially important residues on loop2 include Y67 and R68. This site provides an interesting contrast to the "Apo 2" site (seen in the apo and holo results). Helix a1 is a key part of the "Apo 2" site, but it plays a very minor role in the pocket shown in the loaded figure. The differences between the two pockets are caused by movement of the protein's C terminus, as shown in Supplemental Figure S14 A. Figure S15 . The predicted site includes residues in loop1 (between a1 and a2), a2, and a3. Potentially important residues on loop1 and a2 include I46, L50, the S52-PP-salicylate arm, and L55. Y75 is a potentially important residue on a3. We created the TACTICS algorithm that analyzes MD simulations to find druggable sites in proteins. TACTICS first clusters the trajectory to select several frames for analysis. It then applies a random forest ML model to each selected frame. The model takes a protein conformation and scores each residue's druggability in that conformation. The model was trained on a database of crystal structures; to capture conformational diversity, multiple structures were included for each protein. Sites identified by the ML algorithm undergo fragment docking to further characterize their bindability. To our knowledge, TACTICS is the only freely available program that both uses an ML model specifically designed for MD trajectories and accepts already-generated MD data. The TACTICS software is available at https://github.com/Albert-Lau-Lab/tactics_protein_analysis . We applied the TACTICS algorithm to the SARS-CoV-2 main protease (nsp5) and 2'-O RNA methyltransferase (nsp10/16). In the main protease, TACTICS found the active site and multiple known allosteric sites; all of these sites have been observed experimentally. In the methyltransferase, TACTICS found the RNA and SAM binding sites and predicted the locations of sites not observed in experimental results. We also applied TACTICS to the Yersinia pestis ArCP. The absence or presence of a large phosphopantetheine (PP) arm in apo or loaded and holo forms makes it a unique test case. TACTICS identifies several potential drug-binding sites that merit further study. Site predictions differ between the protein's apo, loaded, and holo forms, suggesting possible differences between the forms' bindability. The TACTICS results for this protein emphasize the software's usefulness for proteins with unusual characteristics. Cryptic binding sites on proteins: definition, detection, and druggability Understanding Cryptic Pocket Formation in Protein Targets by Enhanced Sampling Simulations Exploring the structural origins of cryptic sites on proteins Equilibrium fluctuations of a single folded protein reveal a multitude of potential cryptic allosteric sites Expanding the Druggable Proteome by Characterization and Prediction of Cryptic Binding Sites Investigating Cryptic Binding Sites by Molecular Dynamics Simulations Epock: rapid analysis of protein pocket dynamics POVME: An algorithm for measuring binding-pocket volumes POVME 2.0: An Enhanced Tool for Determining Pocket Shape and Volume Characteristics POVME 3.0: Software for Mapping Binding Pocket Flexibility TRAPP: A Tool for Analysis of Transient Binding Pockets in Proteins TRAPP webserver: predicting protein binding site flexibility and detecting transient binding pockets Druggability Assessment in TRAPP Using Machine Learning Approaches LIGSITE: automatic and efficient detection of potential small molecule-binding sites in proteins KVFinder: steered identification of protein cavities as a PyMOL plugin Protein-ligand binding site recognition using complementary binding-specific substructure comparison and sequence profile alignment COFACTOR: improved protein function prediction by combining structure, sequence and protein-protein interaction information Predicting Protein Ligand Binding Sites by Combining Evolutionary Sequence Conservation and 3D Structure Fpocket: An open source platform for ligand pocket detection GalaxySite: ligand-binding-site prediction by using molecular docking Detection of multiscale pockets on protein surfaces using mathematical morphology Analyzing the Topology of Active Sites: On the Prediction of Pockets and Subpockets Fragment-based identification of druggable 'hot spots' of proteins using Fourier domain correlation techniques Improving protein-ligand binding site prediction accuracy by classification of inner pocket points using local features Combining Global and Local Measures for Structure-Based Druggability Predictions Allosteric Proteins by Dynamics Simulations in the Presence of Probe Molecules Deciphering Cryptic Binding Sites on Proteins by Mixed-Solvent Molecular Dynamics Moving Beyond Active-Site Detection: MixMD Applied to Allosteric Systems MixMD Probeview: Robust Binding Site Prediction from Cosolvent Simulations Computational Fragment-Based Binding Site Identification by Ligand Competitive Saturation Identification and characterization of fragment binding sites for allosteric ligand design using the site identification by ligand competitive saturation hotspots approach Exploring Cryptic Pockets Formation in Targets of Pharmaceutical Interest with SWISH Systematic exploration of multiple drug binding sites Transient Pockets on Protein Surfaces Involved in Protein−Protein Interaction Efficient Characterization of Protein Cavities within Molecular Simulation Trajectories: trj_cavity MDpocket: open-source cavity detection and characterization on molecular dynamics trajectories D3Pockets: A Method and Web Server for Systematic Analysis of Protein Pocket Dynamics A general and Robust Ray-Casting-Based Algorithm for Triangulating Surfaces at the Nanoscale NanoShaper-VMD interface: computing and visualizing surfaces, pockets and channels in molecular systems Allosteric Communication Networks in Proteins Revealed through Pocket Crosstalk Analysis The FTMap family of web servers for determining and characterizing ligand-binding hot spots of proteins Cooperative Changes in Solvent Exposure Identify Cryptic Pockets, Switches, and Allosteric Coupling A Collective Variable for the Rapid Exploration of Protein Druggability Comparative Protein Structure Modeling Using MODELLER The PyMOL Molecular Graphics System An introduction to ROC analysis Scikit-Learn: Machine Learning in Python MDAnalysis: A Python Package for the Rapid Analysis of Molecular Dynamics Simulations MDAnalysis: A toolkit for the analysis of molecular dynamics simulations AutoDock Vina: Improving the speed and accuracy of docking with a new scoring function, efficient optimization, and multithreading Crystal structure of SARS-CoV-2 main protease provides a basis for design of improved α-ketoamide inhibitors CHARMM: the biomolecular simulation program A web-based graphical user interface for CHARMM ModLoop: automated modeling of loops in protein structures Replica exchange with solute scaling: A more efficient version of replica exchange with solute tempering (REST2) Scalable molecular dynamics with NAMD All-Atom Empirical Potential for Molecular Modeling and Dynamics Studies of Proteins Extending the treatment of backbone energetics in protein force fields: limitations of gas-phase quantum mechanics in reproducing protein conformational distributions in molecular dynamics simulations Optimization of the Additive CHARMM All-Atom Protein Force Field Targeting Improved Sampling of the Backbone Φ, ψ and Side-Chain χ1 and χ2 Dihedral Angles Comparison of simple potential functions for simulating liquid water Molecular impact of covalent modifications on nonribosomal peptide synthetase carrier protein communication Solution Structure of a Nonribosomal Peptide Synthetase Carrier Protein Loaded with Its Substrate Reveals Transient, Well-Defined Contacts CHARMM general force field: A force field for drug-like molecules compatible with the CHARMM all-atom additive biological force fields Automation of the CHARMM General Force Field (CGenFF) I: Bond Perception and Atom Typing Automation of the CHARMM General Force Field (CGenFF) II: Assignment of Bonded Parameters and Partial Atomic Charges SARS-CoV-2 M pro inhibitors and activity-based probes for patient-sample imaging Targeting the Dimerization of the Main Protease of Coronaviruses: A Potential Broad-Spectrum Therapeutic Strategy The substrate specificity of SARS coronavirus 3C-like proteinase Prediction and biochemical analysis of putative cleavage sites of the 3C-like protease of Middle East respiratory syndrome coronavirus The development of Coronavirus 3C-Like protease (3CLpro) inhibitors from 2010 to 2020 Only one protomer is active in the dimer of SARS 3C-like proteinase The N-terminal octapeptide acts as a dimerization inhibitor of SARS coronavirus 3C-like proteinase Allosteric Inhibition of the SARS-CoV-2 Main Protease: Insights from Mass Spectrometry Based Assays SARS-CoV-2 Simulations Go Exascale to Capture Spike Opening and Reveal Cryptic Pockets Across the Proteome. bioRxiv D3Targets-2019-nCoV: a webserver for predicting drug targets and for multi-target and multi-site based virtual screening against COVID-19 Elucidation of cryptic and allosteric pockets within the SARS-CoV-2 protease. bioRxiv Crystallographic and electrophilic fragment screening of the SARS-CoV-2 main protease Massive X-ray screening reveals two allosteric drug binding sites of SARS-CoV-2 main protease. bioRxiv Structural basis of SARS-CoV-2 main protease inhibition by a broad-spectrum anti-coronaviral drug Structure of Mpro from SARS-CoV-2 and discovery of its inhibitors SARS-CoV-2 Main Protease: A Molecular Dynamics Study Protein Flexibility and Stiffness Enable Efficient Enzymatic Catalysis Targeting the Dimerization of the Main Protease of Coronaviruses: A Potential Broad-Spectrum Therapeutic Strategy Structural Basis of RNA Cap Modification by SARS-CoV-2 Structural analysis of the SARS-CoV-2 methyltransferase complex involved in coronaviral Epigenetic protein families: A new frontier for drug discovery SAM/SAH Analogs as Versatile Tools for SAM-Dependent Methyltransferases A chemical biology toolbox to study protein methyltransferases and epigenetic signaling Genotype and phenotype of COVID-19: Their roles in pathogenesis Crystal structure and functional analysis of the SARS-coronavirus RNA cap 2′-omethyltransferase nsp10/nsp16 complex Binding of the Methyl Donor S-Adenosyl-l-Methionine to Middle East Respiratory Syndrome Coronavirus 2'-O-Methyltransferase nsp16 Promotes Recruitment of the Allosteric Activator nsp10 Structural analysis of the SARS-CoV-2 methyltransferase complex involved in RNA cap creation bound to sinefungin High-resolution structures of the SARS-CoV-2 2′-O-methyltransferase reveal strategies for structure-based inhibitor design Predicting protein interaction sites: binding hotspots in protein-protein and protein-ligand interfaces We thank Dominique P. Frueh for helpful discussions. MD simulation data and TACTICS output is available from the Zenodo website; it can be found using the DOI 10.5281/zenodo.4538911. The TACTICS code is available at https://github.com/Albert-Lau-Lab/tactics_protein_analysis.