ArticlePDF Available

Benchmarking the Ability of Common Docking Programs to Correctly Reproduce and Score Binding Modes in SARS-CoV-2 Protease Mpro

Authors:

Abstract and Figures

The coronavirus SARS-CoV-2 main protease, Mpro, is conserved among coronaviruses with no human homolog and has therefore attracted significant attention as an enzyme drug target for COVID-19. The number of studies targeting Mpro for in silico screening has grown rapidly, and it would be of great interest to know in advance how well docking methods can reproduce the correct ligand binding modes and rank these correctly. Clearly, current attempts at designing drugs targeting Mpro with the aid of computational docking would benefit from a priori knowledge of the ability of docking programs to predict correct binding modes and score these correctly. In the current work, we tested the ability of several leading docking programs, namely, Glide, DOCK, AutoDock, AutoDock Vina, FRED, and EnzyDock, to correctly identify and score the binding mode of Mpro ligands in 193 crystal structures. None of the codes were able to correctly identify the crystal structure binding mode (lowest energy pose with root-mean-square deviation < 2 Å) in more than 26% of the cases for noncovalently bound ligands (Glide: top performer), whereas for covalently bound ligands the top score was 45% (EnzyDock). These results suggest that one should perform in silico campaigns of Mpro with care and that more comprehensive strategies including ligand free energy perturbation might be necessary in conjunction with virtual screening and docking.
Content may be subject to copyright.
Benchmarking the Ability of Common Docking Programs to
Correctly Reproduce and Score Binding Modes in SARS-CoV2
Protease Mpro
Shani Zev, Keren Raz, Renana Schwartz, Reem Tarabeh, Prashant Kumar Gupta, and Dan T. Major*
Cite This: J. Chem. Inf. Model. 2021, 61, 29572966
Read Online
ACCESS Metrics & More Article Recommendations *
sıSupporting Information
ABSTRACT: The coronavirus SARS-CoV-2 main protease, Mpro,
is conserved among coronaviruses with no human homolog and
has therefore attracted signicant attention as an enzyme drug
target for COVID-19. The number of studies targeting Mpro for in
silico screening has grown rapidly, and it would be of great interest
to know in advance how well docking methods can reproduce the
correct ligand binding modes and rank these correctly. Clearly,
current attempts at designing drugs targeting Mpro with the aid of
computational docking would benet from a priori knowledge of
the ability of docking programs to predict correct binding modes
and score these correctly. In the current work, we tested the ability
of several leading docking programs, namely, Glide, DOCK,
AutoDock, AutoDock Vina, FRED, and EnzyDock, to correctly identify and score the binding mode of Mpro ligands in 193 crystal
structures. None of the codes were able to correctly identify the crystal structure binding mode (lowest energy pose with root-mean-
square deviation < 2 Å) in more than 26% of the cases for noncovalently bound ligands (Glide: top performer), whereas for
covalently bound ligands the top score was 45% (EnzyDock). These results suggest that one should perform in silico campaigns of
Mpro with care and that more comprehensive strategies including ligand free energy perturbation might be necessary in conjunction
with virtual screening and docking.
INTRODUCTION
Coronaviruses are positive-stranded RNA viruses that infect
humans and animals and cause common and severe respiratory
diseases, including severe acute respiratory syndrome (SARS)
and Middle East respiratory syndrome (MERS).
1,2
These
viruses rely heavily on functional polypeptides that are
generated by proteolytic cleavage of polyproteins that are
translated from viral RNA. The principal coronavirus proteases
responsible for this polypeptide formation are mainly protease
and papain protease. The coronavirus SARS-CoV-2 main
protease, Mpro (henceforth denoted simply as Mpro), has
garnered signicant attention in the past year as an enzyme
drug target due to the COVID-19 pandemic outbreak.
3
Mpro is
a druggable target
4,5
as it is conserved among coronaviruses
and has no human homolog. The rst Mpro structures were
published in early 2020.
3,6,7
The rst crystal structures revealed
that the active form of Mpro is a homodimer containing two
protomers, each composed of three domains (Figure 1A). The
active site in Mpro is located in a cleft between domains I and II
(Figure 1B) and features a noncanonical catalytic CysHis
dyad. The active site is composed of four regions: S1, S1, S2,
and S4, with the catalytic Cys145 located in S1and His41
located in S1and S2 (Figure 1C). To date, well over 200
three-dimensional holostructures have been resolved at a
resolution of 3 Å or better.
3,69
In these structures, ligands
bind to a variety of binding sites, including covalent and
noncovalent binding in the main catalytic site, noncovalent
binding in pockets in between the two proteomers, and weakly
bound at the protein surface (i.e., in between crystallographic
homodimer units). This wealth of structural information makes
Mpro an attractive benchmarking system for testing the ability
of docking programs to correctly identify and rank the correct
poses. This becomes particularly important in light of the
severity of the COVID-19 pandemic
2
and the signicant
number of mutant forms of the virus that are rapidly appearing
and might render current and future vaccines ineective. To
date, Mpro has been studied extensively using ligand docking
and screening tools
1019
and computational enzymology tools,
such as hybrid quantum mechanicsmolecular mechanics
(QM/MM).
2029
Received: March 6, 2021
Published: May 28, 2021
Articlepubs.acs.org/jcim
© 2021 American Chemical Society 2957
https://doi.org/10.1021/acs.jcim.1c00263
J. Chem. Inf. Model. 2021, 61, 29572966
Made available for a limited time for personal research and study only License.
Downloaded via BAR-ILAN UNIV on July 29, 2021 at 07:10:40 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
The number of studies targeting Mpro for in silico screening
has grown,
30
and it would be of great interest to know in
advance how well docking methods can reproduce the correct
ligand binding modes and rank these correctly. Clearly, current
attempts at designing drugs targeting Mpro with the aid of
computational docking are problematic if programs struggle to
predict correct binding modes and score these correctly. This
is true even if common docking programs have undergone
extensive testing since each protein target comes with its own
challenges due to the complexity of binding pocket crevices,
nature of interactions, and solvent exposure. Therefore, critical
evaluation of how well common docking programs perform for
Mpro is important.
Since the development of the rst automated docking
program DOCK,
3135
a multitude of docking software
packages have been developed, with dierent physicochemical
approximations and algorithmic details. Popular docking
programs in addition to DOCK, include Autodock,
3638
Autodock Vina,
39
Glide,
4042
Rosettaligand,
43,44
Gold,
4547
and CDocker.
4851
Widely used search algorithms include
molecular dynamics (MD), Monte Carlo (MC), and genetic
algorithms (GA). Common scoring functions include force
eld-based energy functions, such as CHARMM,
5256
AMBER,
57
and OPLS
58,59
(i.e., molecular mechanics, MM),
and knowledge-based scoring functions, such as DRUG-
SCORE,
60
IT-SCORE,
61
DSX,
62
CHEMSCORE,
63
and
SMoG.
64
Specialized docking programs addressing en-
zymes,
65,66
such as EnzyDock, have also been developed.
67
Current challenges for docking methods include protein
exibility,
68
ligand solvation, and binding-site hydration.
47,69
Thus far, several docking approaches have been employed to
screen Mpro for potential drugs in virtual screening and drug
repurposing campaigns, including Glide,
7,10,13,17,18,7072
Auto-
dock,
11,13,73,74
Autodock Vina,
11,13,14,19,71,73,75,76
Surex,
77
PLANT,
78
DockThor,
76
fast pulling of ligands,
14
deep
docking,
70
algebraic topology and deep learning,
79
and virtual
reality-based docking.
16
However, to the best of our knowl-
edge, no rigorous benchmark study addressing the ability of
such docking tools to reproduce and correctly rank known
ligand binding modes has been published, in spite of the
known inherent challenges in docking.
8083
In the current work, we tested the ability of several leading
docking programs to correctly identify and score the binding
mode of Mpro ligands in 193 crystal structures (Figure S1). We
tested the following docking programs: Glide,
4042
DOCK,
3135
Autodock,
3638
Autodock Vina,
39
FRED,
8487
and EnzyDock.
67
The current results suggest that care should
be taken in applying docking programs to a challenging protein
target such as Mpro.
METHODS
Preparing Mpro Structure Database. The available
crystal structures of ligand-containing SARS-CoV-2 Mpro were
downloaded from the RCSB PDB website (March-December
2020).
88
In total, we collected 193 dierent structures,
including covalent and noncovalent ligands (Table S1 and
Figure S1). All structures were aligned relative to one reference
structure (PDB ID: 5R84) for easy comparison. To perform
docking, we separated each protein and ligand into separate
les, removing crystal waters, ions, and cosolvents. Missing
residues were added using the Modeller homology program.
89
Hydrogens were added using the CHARMM simulation
platform (using HBUILD) for the protein structure and
using Openbabel for the ligands.
90,91
Visual inspection was also
performed. For systems including only one monomer, the
complementary unit was generated using the crystallographic
information included in the PDB le using CHARMM.
Protonation states of His residues were determined based on
hydrogen bonding patterns and knowledge of the chemistry
catalyzed by Mpro (Table S2), and they match the protonation
states of key His residues recently published.
23
All docking
simulations described below commenced with the CHARMM
prepared systems.
Clustering of Ligands and Water Molecules. Chemical
descriptors were calculated for all ligands from the 193 PDB
les using RDKit libraries in Python. Features thought to be
important for ligand binding were chosen. Specically, we
computed the number of rotatable torsions, molecular weight,
number of H-bond donors and acceptors, number of aromatic
rings, the fraction of carbon atoms in sp3hybridization
(relative to all carbon atoms in the molecule), and log Pvalues.
We applied principal component analysis (PCA) using these
ligands descriptors, followed by k-means clustering to cluster
the ligands into groups. We selected the number of clusters by
silhouette analysis of the k-means clustering results. Ligand
clustering was performed using Python 3.7. Density-Based
Spatial Clustering of Applications with Noise (DBSCAN)
clustering was performed to analyze the water molecules from
all crystal structures. These water molecules were not included
in the docking studies.
Subsite-Binding Pocket Binding. To classify the binding
patterns of all the 193 proteinligand complexes, we
categorized the ligands as bound on the surface, at the dimer
interface, or in the active site. The latter was characterized
according to the subsites S1, S1, S2, and S4 (Figure 1C, Table
S3).
7,77
A ligand is considered to occupy a binding pocket if
any ligand atom is within 4.0 Å of any pocket atom and also
within 3.0 Å of the geometric center of the pocket (dened as
geometric center of all pocket atoms). Moreover, a ligand is
considered to be in the proximity of a binding pocket if any
ligand atom is within 4.0 Å of any pocket atom and also within
Figure 1. (A) SARS-CoV-2 Mpro dimer with bound ligand in each
active site (PDB ID: 7BQY). The domains in chain A are colored as
follows: N-nger in bright orange, domain I in pale green, domain II
in light blue, loop L3 in light teal, and domain III in light pink. The
ligand bound to chain A is colored yellow and appears in stick
representation. Both chain B and the ligand bound to it appear in
gray. The same color code applies to (B), where the active site is
shown in greater detail. (C) Four conserved subsites are presented,
along with the Cys 145His 41 dyad (deep teal stick representation).
The covalently bound ligand appears in limon-shade stick
representation. The protein appears as a white surface.
Journal of Chemical Information and Modeling pubs.acs.org/jcim Article
https://doi.org/10.1021/acs.jcim.1c00263
J. Chem. Inf. Model. 2021, 61, 29572966
2958
5.0 Å of the geometric center of the pocket. The criteria were
designed to account for both small and bulky ligands and to
distinguish between binding poses where ligand groups are
well docked inside a subpocket and poses where ligand groups
are located at the periphery of a subpocket. Cutovalues are
suitable for various nonbonded interactions (e.g., ππ
stacking, hydrogen bond, ionic, and hydrophobic interactions).
The nal values were obtained by trial and error and validated
by means of visual inspection. The subsite-binding pocket
occupancy analysis was implemented as a CHARMM
90,91
script.
Docking Protocols. To compare the performance of
selected docking programs for use with Mpro (search algorithm
and scoring function), we performed noncovalent docking
using DOCK,
3135
Autodock Vina,
39
and FRED
8487
and
noncovalent and covalent docking using Glide,
4042
Auto-
dock,
3638
and EnzyDock
67
to the systems described above. In
all docking simulations described below, the ligand was fully
exible, while the protein was xed (except for the covalently
connected complexes, where the appropriate Cys residue was
exible).
Ligand Docking with Glide. Proteins and ligands were
prepared using Schrödingers Maestro (version 11.4, 2017-4
release) Prep Wiz and LigPrep modules, respectively, with
default settings for docking with Glide. All covalent docking
simulations were performed using the CovDock module
available in Glide.
92
For the noncovalent simulations, the
grid was generated using XGlide, which enables creation of
dierent grids in parallel. The grids were centered around the
ligands centroid. The dimensions of the enclosing box and the
bounding box were set to 12 ×12 ×12 Å3and 26 ×26 ×26
Å3, respectively, for all cases. The ligand stereochemistry was
kept during all docking simulations. The number of poses
written per ligand was set to 10,000. The scaling factors of the
vdW radii and the partial atomic charge cutowere set to the
default values 0.80 and 0.15, respectively. Standard precision
(SP) mode was chosen for all ligand docking runs. The
selection of the best-docked ligand structure among the
proposed poses is made based on several model energies
implemented with Glide (docking score, Prime energy and E-
model energy, and cdock anity). Solvent eects were
incorporated using MMGBSA. All reported energies herein
used the docking score function for noncovalent docking and
the cdock anity scoring function for covalent docking as
these performed best [i.e., produced the highest number of
top-ranking structures with root-mean-square deviation (rmsd)
< 2 Å]. In all Glide docking simulations (ligand preparation,
protein preparation, grid generation, covalent, and noncovalent
docking), the OPLS3 force eld
59,93
was used.
Ligand Docking with DOCK (Version 6.9). Proteins and
ligands were prepared for docking simulations using the
DockPrep option of Chimera v.14.
94
The grid was generated
according to the center of mass of the crystal structure ligand
with a grid spacing of 0.4 Å. The maximum and minimum
radius of the sphere generated was set to 4 and 1.4 Å. All the
spheres within 10 Å of the ligand were selected for docking.
The box length surrounding the ligand was set to 6 Å, that is,
the edge of the box from any atom of the ligand was at least 6
Å away, which easily accommodates all the selected spheres.
Ligand Docking with Autodock (Version 4.2). Proteins
and ligand pdbqt les were prepared using standard AutoDock
tools (prepare_exreceptor4.py and prepare_ligand4.py).
These les include Cartesian coordinates and Gasteiger atomic
charges
95
for each atom. AutoDock employs a united atom
method, and thus, no nonpolar hydrogens are present. The
center of mass of the crystallographic ligand was used to
determine the center of the grid. AutoDock uses one grid box
to perform the docking calculations, and the dimensions of this
box were set to 37.5 ×37.5 ×37.5 Å3and the spacing was set
to 0.375 Å for all systems. We performed exible ligand
docking into a rigid protein environment using GA, with
default settings. For covalent docking,
96
each ligand was
prepared with the active Cys residue already present in the
input le using AutoDock tools (prepare_receptor4.py and
prepare_exreceptor4.py). For covalent docking, the ligand
exible torsional angles were presampled using MC simu-
lations with CHARMM prior to docking.
Ligand Docking with Autodock Vina. Protein and
ligand input pdbqt les were prepared in the same way as for
Autodock4.2 (see above). The size of the grid was set to 30.0
Figure 2. Normalized radar plot showing various features of each cluster centroid for SARS-CoV-2 Mpro ligands. Inset: pie plot with the relative
fraction of each cluster among all the ligands studied in this work. The rst number is the cluster name, and the second is the number of ligands in
the cluster.
Journal of Chemical Information and Modeling pubs.acs.org/jcim Article
https://doi.org/10.1021/acs.jcim.1c00263
J. Chem. Inf. Model. 2021, 61, 29572966
2959
×30.0 ×30.0 Å3, and remaining parameters were set to default
values.
Ligand Docking with FRED.
8587
FRED is one of the
docking programs available within the OpenEye scientic
library. For the docking process, proteins and ligands were
prepared using the graphical user interface Make Receptor
provided with OpenEye. FRED creates a potential eld around
the binding site by producing a negative image, which
complements the shape of the protein site. This potential
eld is represented on a contour, which completely surrounds
the ligand. OMEGA, an internal program within OpenEye is
used to generate an ensemble of conformers for each ligand. A
total of 200 dierent conformers were generated for each
ligand for docking, and the 50 lowest energy docked structures
were used to select the best pose in terms of lowest rmsd or
Chemgauss energy scoring relative to the crystal ligand
structure. The proteins and ligands were held xed during
the docking process.
Ligand Docking with EnzyDock. Protein and ligand les
were prepared as described at the beginning of Methods.
CHARMM topology (RTF) and parameters (PRM) les for
the ligands were generated using the CHARMM General Force
Field (CGenFF) program.
53,97
For the proteins, CHARMM 36
was used.
52,54,55
The grid was generated according to the
center of mass of the crystal structure ligand. The grid was
generated with a mesh spacing of 0.25 Å and dimensions of
30.0 Å along each axis. The docking entailed 25 cycles of
simulated MC and MD annealing for 25 dierently rotated and
MC torsion-sampled ligand congurations. Settings for non-
covalent and covalent docking were identical.
RESULTS AND DISCUSSION
Ligand Clustering and the Subsite-Binding Pocket.
To better understand the nature of the 193 Mpro complexes
prior to discussing the docking results, we present analyses of
the ligands and their binding modes. The ligands were
clustered into seven main groups based on their chemical
features by PCA and k-means clustering. The features of each
cluster were normalized, and the average value for each cluster
was calculated (Figure 2). The relative amount of the 193
ligands composing each cluster can be seen in the inserted pie
chart (Figure 2).
For instance, cluster 6 is characterized by 17 ligands of low
molecular weight and high fraction of carbon atoms with sp3
hybridization, while cluster 3 is composed of 78 low-molecular-
weight ligands that are rather rigid and slightly hydrophobic.
Cluster 4 has 15 ligands of high molecular weight, exible
chains with sp3carbons, and several hydrogen bond donors
and acceptors, while cluster 7 has medium-molecular-weight
ligands that are highly hydrophobic with aromatic rings, yet has
some hydrogen bond donors and acceptors.
Next, we analyzed the binding modes of the ligand clusters
in Mpro (Tables 1 and S1). In Table 1 we present the fraction
of each cluster that is bound in a specic subpocket of the
active site, at the surface, or at the interface between the two
monomers. Note that ligands can bind in more than one
pocket, and hence, the fractions do not add up to unity for
each cluster. Inspection of the binding data clearly shows that
ligands of low molecular weight (clusters 1, 3, 5, and 6) tend to
bind at the surface of the protein (i.e., clusters 3, 5, and 6 are
caught in between crystal units) or at the interface between the
homodimer units (cluster 1). Still many low-molecular-weight
ligands occupy pockets S1 and S1as these are covalently
attached to C145. Ligands rich in aromatic rings and
correspondingly high log Pvalues (cluster 7) tend to occupy
all pockets more than average, specically sites S1 and S2. This
is due to favorable ππinteractions with F140, H163, and
H172 located in S1 or H41 and Y54 in S2. Another important
observation is that ligands more likely to bind to S4 (which is
rarely occupied) belong to clusters 2 and 4, which tend to
include large, exible molecules that are rich in H-bond donors
and acceptors.
We also clustered the water molecules in all crystal
structures using DBSCAN clustering. Following clustering,
we removed all waters that overlap any bound ligand (Figure
S2). These waters form a set of active site features that can be
included in docking studies (these waters were not included in
the current docking study).
Docking of Ligands in Mpro.We docked all ligands from
193 crystal Mpro structures into their respective crystal protein
structure (Table S4). These crystal structures include ligands
bound noncovalently to the main binding pocket, surface and
dimer interface, as well as covalently attached ligands. In all
results below, we present the success rate of dierent docking
programs in reproducing the crystal bound poses. For DOCK,
AutoDock Vina, and FRED, the results reect the noncovalent
complexes only.
In Figure 3A we show the overall success of all programs.
Glide and EnzyDock reproduce the correct crystal structure
pose (rmsd < 2 Å) for over 50% of the structures, with success
rates of 64 and 70%, respectively, while for AutoDock, this rate
falls to 40%. However, in many cases, even if the correct pose
is identied, it is not scored as lowest in energy, and the
success rate reduces to 33% (Glide), AutoDock (30%), and
EnzyDock (35%). The overall success rates of Glide,
AutoDock, and EnzyDock are in part due to the covalent
complexes, whose poses are easier to reproduce than the
noncovalent ones. If we analyze the success rates of identifying
only the covalently bound complexes, Glide, AutoDock, and
EnzyDock identify the correct poses 70, 42, and 71% of the
cases, while the correct pose is also scored as the best one in
38, 36, and 45% of the cases (Figure 3B). For the noncovalent
complexes, Glide, DOCK, AutoDock, AutoDock Vina, FRED,
and EnzyDock identify the correct poses in 55, 61, 37, 29, 46,
and 68% of the cases, respectively, while these are ranked as
the lowest energy poses in 26, 20, 24, 14, 14, and 22% of the
cases, respectively (Figure 3C). Finally, if we remove the
complexes with surface-bound ligands (i.e., ligands bound in
between crystal units), all programs perform signicantly better
(Figure 3D). The correct poses are identied (and scored as
best) as follows (%): Glide 74 (39), DOCK 77 (29),
Table 1. Fraction of Ligands in Each Cluster that are
Located in a Specic Active-Site Pocket, on the Surface, or
at the Interface between Monomers in the SARS-CoV-2
Mpro Crystal Structures
S1 S1S2 S4 surface dimer interface
1 0.74 0.32 0.62 0.29 0.06 0.18
2 0.84 0.44 0.74 0.50 0.04 0.00
3 0.42 0.35 0.16 0.05 0.22 0.06
4 0.87 0.67 0.80 0.87 0.00 0.00
5 0.57 0.18 0.25 0.10 0.27 0.00
6 0.24 0.44 0.12 0.00 0.29 0.06
7 0.82 0.50 0.77 0.36 0.09 0.09
Total 0.57 0.38 0.37 0.22 0.17 0.05
Journal of Chemical Information and Modeling pubs.acs.org/jcim Article
https://doi.org/10.1021/acs.jcim.1c00263
J. Chem. Inf. Model. 2021, 61, 29572966
2960
AutoDock 48 (35), AutoDock Vina 40 (23), FRED 61 (18),
and EnzyDock 80 (35), respectively.
Next, we analyze how the dierent programs perform as a
function of binding site locations on Mpro.InFigures S3 and 4,
we present box plots of the best rmsd values and the rmsd
values for the lowest scoring pose for noncovalently bound
ligands, respectively. All methods struggle with ligands bound
at the interface between monomers and on the protein surface,
Figure 3. Docking pose prediction success (%) for selected methods for 193 crystal structures of SARS-CoV-2 Mpro. For each column, the upper
part represents the ability of methods to identify poses with rmsd < 2 Å, while the lower part represents ability of methods to identify a pose with
rmsd < 2 Å as the lowest energy pose. (A) All 193 crystal structures, (B) 108 crystal structures with covalently bound ligands, (C) 85 crystal
structures with noncovalently bound ligands, (D) 51 crystal structures with noncovalently bound ligands excluding surface bound ligands.
Figure 4. Distributions of rmsd values (Å) for the lowest-scoring pose at dierent sites in SARS-CoV-2 Mpro for selected noncovalent docking
methods. The ligands were clustered into groups occupying similar regions, and only clusters with more than seven members are shown.
Figure 5. Distributions of rmsd values (Å) for the lowest-scoring pose at dierent sites in SARS-CoV-2 Mpro for selected covalent docking methods.
The ligands were clustered into groups occupying similar regions, and only clusters with more than 22 members are shown.
Journal of Chemical Information and Modeling pubs.acs.org/jcim Article
https://doi.org/10.1021/acs.jcim.1c00263
J. Chem. Inf. Model. 2021, 61, 29572966
2961
with Glide and FRED producing the best results. Additionally,
most methods perform better for ligands bound to more than a
single pocket (i.e., S1 + S2), and this trend is particularly clear
for Glide and EnzyDock.
Similarly, we analyze how the dierent covalent docking
programs perform as a function of binding site locations on
Mpro.InFigures S4 and 5, we present box plots of the best
rmsd values and the rmsd values for the lowest scoring pose for
covalently bound ligands, respectively. Also here, we observe a
general trend, where the docking methods perform better for
ligands occupying more pockets.
CONCLUSIONS
In wake of the growing threat emerging from the SARS-CoV-2
pandemic, the modeling community has rushed to study a
variety of potential pharmaceutical targets. One of these
targets, Mpro, is particularly attractive as this enzyme has no
human analogue and is conserved among coronaviruses. A
large number of studies have addressed ligand docking and
virtual screening of ligand libraries against Mpro in search of
promising leads. A prerequisite for such studies is the ability of
the docking programs to correctly identify and score ligand
poses. Due to the intense eorts by the scientic community,
there is already a wealth of structural biology information on
Mpro, hence enabling comparative studies of dierent docking
approaches against this target. To date, the available crystal
structures of Mpro include ligands bound covalently and
noncovalently to the main catalytic site, surface, and in
between the two monomers. Here, we studied several leading
docking codes, namely, Glide, DOCK, AutoDock, AutoDock
Vina, FRED, and EnzyDock, and evaluate their ability to
correctly reproduce and score the crystal structure ligand
conguration for 193 Mpro crystal structures. None of the
codes are able to correctly identify and score the crystal
structure in more than 26% of the cases for noncovalently
bound ligands (Glide top performer), whereas for covalently
bound ligands, the top score was 45% (EnzyDock best
performer). Additionally, a general trend, where several of the
docking methods (e.g., Glide and EnzyDock) perform better
for larger, bulkier ligands occupying more than a single pocket,
is observed. All docking methods struggle with prediction of
small ligands. In the original crystal structures, many of the
smaller ligands are surrounded by numerous explicit water
molecules, dimethyl sulfoxide molecules, and Clions that
were removed prior to docking. Thus, these redocking trends
may be ascribed to diculty in accurately scoring docking
poses, where a delicate balance between intra- and
intermolecular terms and solvation terms must be stricken.
In conclusion, the current results suggest that one should
perform docking studies and virtual screening campaigns of
Mpro with care and that more comprehensive strategies might
be necessary. Such strategies might include initial virtual
screening (e.g., using FRED or AutoDock Vina) or docking
(e.g., Glide or EnzyDock), followed by more rigorous ligand
free energy binding calculations
98,99
and in-depth QM/MM
studies.
20,24,26,28,29
Inclusion of conserved water molecules, as
identied in this study, may also be of help in guiding the
docking process. Indeed, MD simulations have pointed to
several water molecules, as important for Mpro.
11,18,20,24,26,29,77
ASSOCIATED CONTENT
*
sıSupporting Information
The Supporting Information is available free of charge at
https://pubs.acs.org/doi/10.1021/acs.jcim.1c00263.
Chemical structures of all docked ligands, conserved
water molecules, additional docking results, all rmsd
data, and active-site pocket occupation (PDF)
AUTHOR INFORMATION
Corresponding Author
Dan T. Major Department of Chemistry and Institute for
Nanotechnology &Advanced Materials, Bar-Ilan University,
Ramat-Gan 52900, Israel; orcid.org/0000-0002-9231-
0676; Email: majort@biu.ac.il
Authors
Shani Zev Department of Chemistry and Institute for
Nanotechnology &Advanced Materials, Bar-Ilan University,
Ramat-Gan 52900, Israel
Keren Raz Department of Chemistry and Institute for
Nanotechnology &Advanced Materials, Bar-Ilan University,
Ramat-Gan 52900, Israel
Renana Schwartz Department of Chemistry and Institute for
Nanotechnology &Advanced Materials, Bar-Ilan University,
Ramat-Gan 52900, Israel
Reem Tarabeh Department of Chemistry and Institute for
Nanotechnology &Advanced Materials, Bar-Ilan University,
Ramat-Gan 52900, Israel
Prashant Kumar Gupta Department of Chemistry and
Institute for Nanotechnology &Advanced Materials, Bar-
Ilan University, Ramat-Gan 52900, Israel; orcid.org/
0000-0002-4792-7538
Complete contact information is available at:
https://pubs.acs.org/10.1021/acs.jcim.1c00263
Author Contributions
S.Z., K.R., and P.K.G. contributed equally. The docking
simulations and system preparations were performed by all
authors. The manuscript was written through contributions of
all authors. All authors have given approval to the nal version
of the manuscript.
Notes
The authors declare no competing nancial interest.
All preprepared 193 Mpro systems and accompanying Python
and CHARMM scripts are available at https://github.com/
shanizev/Benchmarking-SARS-CoV-2.EnzyDockisfreely
available for noncommercial use on https://github.com/
majordt/EnzyDock.
ACKNOWLEDGMENTS
This work was supported by the Israel Ministry of Science,
Technology and Space (grant 3-16310) and Israeli Science
Foundation (grant #1683/18).
ABBREVIATIONS
SARS, severe acute respiratory syndrome; SARS-CoV-2, SARS
coronavirus 2; MERS, Middle East respiratory syndrome; MD,
molecular dynamics; MC, Monte Carlo
Journal of Chemical Information and Modeling pubs.acs.org/jcim Article
https://doi.org/10.1021/acs.jcim.1c00263
J. Chem. Inf. Model. 2021, 61, 29572966
2962
REFERENCES
(1) Anand, K.; Ziebuhr, J.; Wadhwani, P.; Mesters, J. R.; Hilgenfeld,
R. Coronavirus main proteinase (3CLpro) structure: Basis for design
of anti-sars drugs. Science 2003,300, 17631767.
(2) Liu, C.; Zhou, Q.; Li, Y.; Garner, L. V.; Watkins, S. P.; Carter, L.
J.; Smoot, J.; Gregg, A. C.; Daniels, A. D.; Jervey, S.; Albaiu, D.
Research and development on therapeutic agents and vaccines for
COVID-19 and related human coronavirus diseases. ACS Cent. Sci.
2020,6, 315331.
(3) Zhang, L.; Lin, D.; Sun, X.; Curth, U.; Drosten, C.; Sauerhering,
L.; Becker, S.; Rox, K.; Hilgenfeld, R. Crystal structure of SARS-CoV-
2 main protease provides a basis for design of improved α-ketoamide
inhibitors. Science 2020,368, 409412.
(4) Qiao, J.; Li, Y.-S.; Zeng, R.; Liu, F.-L.; Luo, R.-H.; Huang, C.;
Wang, Y.-F.; Zhang, J.; Quan, B.; Shen, C.; Mao, X.; Liu, X.; Sun, W.;
Yang, W.; Ni, X.; Wang, K.; Xu, L.; Duan, Z.-L.; Zou, Q.-C.; Zhang,
H.-L.; Qu, W.; Long, Y.-H. -P.; Li, M.-H.; Yang, R.-C.; Liu, X.; You, J.;
Zhou, Y.; Yao, R.; Li, W.-P.; Liu, J.-M.; Chen, P.; Liu, Y.; Lin, G.-F.;
Yang, X.; Zou, J.; Li, L.; Hu, Y.; Lu, G.-W.; Li, W.-M.; Wei, Y.-Q.;
Zheng, Y.-T.; Lei, J.; Yang, S. SARS-CoV-2 Mpro inhibitors with
antiviral activity in a transgenic mouse model. Science 2021,371,
13741378.
(5) Pelly, S.; Liotta, D. Potent SARS-CoV-2 direct-acting antivirals
provide an important complement to COVID-19 vaccines. ACS Cent.
Sci. 2021,7, 396399.
(6) Dai, W.; Zhang, B.; Jiang, X.-M.; Su, H.; Li, J.; Zhao, Y.; Xie, X.;
Jin, Z.; Peng, J.; Liu, F. Structure-based design of antiviral drug
candidates targeting the SARS-CoV-2 main protease. Science 2020,
368, 13311335.
(7) Jin, Z.; Du, X.; Xu, Y.; Deng, Y.; Liu, M.; Zhao, Y.; Zhang, B.; Li,
X.; Zhang, L.; Peng, C.; Duan, Y.; Yu, J.; Wang, L.; Yang, K.; Liu, F.;
Jiang, R.; Yang, X.; You, T.; Liu, X.; Yang, X.; Bai, F.; Liu, H.; Liu, X.;
Guddat, L. W.; Xu, W.; Xiao, G.; Qin, C.; Shi, Z.; Jiang, H.; Rao, Z.;
Yang, H. Structure of m(pro) from SARS-CoV-2 and discovery of its
inhibitors. Nature 2020,582, 289293.
(8) Kneller, D. W.; Phillips, G.; ONeill, H. M.; Jedrzejczak, R.; Stols,
L.; Langan, P.; Joachimiak, A.; Coates, L.; Kovalevsky, A. Structural
plasticity of SARS-CoV-2 3CL Mpro active site cavity revealed by
room temperature x-ray crystallography. Nat. Commun. 2020,11,
3202.
(9) Günther, S.; Reinke, P. Y. A.; Fernández-García, Y.; Lieske, J.;
Lane, T. J.; Ginn, H. M.; Koua, F. H. M.; Ehrt, C.; Ewert, W.;
Oberthuer, D.; Yefanov, O.; Meier, S.; Lorenzen, K.; Krichel, B.;
Kopicki, J. D.; Gelisio, L.; Brehm, W.; Dunkel, I.; Seychell, B.;
Gieseler, H.; Norton-Baker, B.; Escudero-Pérez, B.; Domaracky, M.;
Saouane, S.; Tolstikova, A.; White, T. A.; Hänle, A.; Groessler, M.;
Fleckenstein, H.; Trost, F.; Galchenkova, M.; Gevorkov, Y.; Li, C.;
Awel, S.; Peck, A.; Barthelmess, M.; Schluenzen, F.; Lourdu Xavier, P.;
Werner, N.; Andaleeb, H.; Ullah, N.; Falke, S.; Srinivasan, V.; França,
B. A.; Schwinzer, M.; Brognaro, H.; Rogers, C.; Melo, D.; Zaitseva-
Doyle, J. J.; Knoska, J.; Peña-Murillo, G. E.; Mashhour, A. R.;
Hennicke, V.; Fischer, P.; Hakanpää, J.; Meyer, J.; Gribbon, P.;
Ellinger, B.; Kuzikov, M.; Wolf, M.; Beccari, A. R.; Bourenkov, G.; von
Stetten, D.; Pompidor, G.; Bento, I.; Panneerselvam, S.; Karpics, I.;
Schneider, T. R.; Garcia-Alai, M. M.; Niebling, S.; Günther, C.;
Schmidt, C.; Schubert, R.; Han, H.; Boger, J.; Monteiro, D. C. F.;
Zhang, L.; Sun, X.; Pletzer-Zelgert, J.; Wollenhaupt, J.; Feiler, C. G.;
Weiss, M. S.; Schulz, E. C.; Mehrabi, P.; Karničar, K.; Usenik, A.;
Loboda, J.; Tidow, H.; Chari, A.; Hilgenfeld, R.; Uetrecht, C.; Cox,
R.; Zaliani, A.; Beck, T.; Rarey, M.; Günther, S.; Turk, D.; Hinrichs,
W.; Chapman, H. N.; Pearson, A. R.; Betzel, C.; Meents, A. X-ray
screening identifies active site and allosteric inhibitors of SARS-CoV-2
main protease. Science 2021,372, 642646.
(10) Wang, J. Fast identification of possible drug treatment of
coronavirus disease-19 (COVID-19) through computational drug
repurposing study. J. Chem. Inf. Model. 2020,60, 32773286.
(11) Acharya, A.; Agarwal, R.; Baker, M. B.; Baudry, J.; Bhowmik,
D.; Boehm, S.; Byler, K. G.; Chen, S. Y.; Coates, L.; Cooper, C. J.;
Demerdash, O.; Daidone, I.; Eblen, J. D.; Ellingson, S.; Forli, S.;
Glaser, J.; Gumbart, J. C.; Gunnels, J.; Hernandez, O.; Irle, S.; Kneller,
D. W.; Kovalevsky, A.; Larkin, J.; Lawrence, T. J.; LeGrand, S.; Liu, S.
H.; Mitchell, J. C.; Park, G.; Parks, J. M.; Pavlova, A.; Petridis, L.;
Poole, D.; Pouchard, L.; Ramanathan, A.; Rogers, D. M.; Santos-
Martins, D.; Scheinberg, A.; Sedova, A.; Shen, Y.; Smith, J. C.; Smith,
M. D.; Soto, C.; Tsaris, A.; Thavappiragasam, M.; Tillack, A. F.;
Vermaas, J. V.; Vuong, V. Q.; Yin, J.; Yoo, S.; Zahran, M.; Zanetti-
Polzi, L. Supercomputer-based ensemble docking drug discovery
pipeline with application to covid-19. J. Chem. Inf. Model. 2020,60,
58325852.
(12) Zhang, L.; Lin, D.; Kusov, Y.; Nian, Y.; Ma, Q.; Wang, J.; von
Brunn, A.; Leyssen, P.; Lanko, K.; Neyts, J.; de Wilde, A.; Snijder, E.
J.; Liu, H.; Hilgenfeld, R. Alpha-ketoamides as broad-spectrum
inhibitors of coronavirus and enterovirus replication: Structure-based
design, synthesis, and activity assessment. J. Med. Chem. 2020,63,
45624578.
(13) Ghahremanpour, M. M.; Tirado-Rives, J.; Deshmukh, M.;
Ippolito, J. A.; Zhang, C.-H.; Cabeza de Vaca, I.; Liosi, M.-E.;
Anderson, K. S.; Jorgensen, W. L. Identification of 14 known drugs as
inhibitors of the main protease of SARS-CoV-2. ACS Med. Chem. Lett.
2020,11, 25262533.
(14) Pham, M. Q.; Vu, K. B.; Pham, T. N. H.; Tran, L. H.; Tung, N.
T.; Vu, V. V.; Nguyen, T. H.; Ngo, S. T. Rapid prediction of possible
inhibitors for SARS-CoV-2 main protease using docking and fpl
simulations. RSC Adv. 2020,10, 3199131996.
(15) Bzówka, M.; Mitusińska, K.; Raczyńska, A.; Samol, A.;
Tuszyński, J. A.; Góra, A. Structural and evolutionary analysis indicate
that the SARS-CoV-2 Mpro is a challenging target for small-molecule
inhibitor design. Int. J. Mol. Sci. 2020,21, 3099.
(16) Deeks, H. M.; Walters, R. K.; Barnoud, J.; Glowacki, D. R.;
Mulholland, A. J. Interactive molecular dynamics in virtual reality is an
effective tool for flexible substrate and inhibitor docking to the SARS-
CoV-2 main protease. J. Chem. Inf. Model. 2020,60, 58035814.
(17) Gahlawat, A.; Kumar, N.; Kumar, R.; Sandhu, H.; Singh, I. P.;
Singh, S.; Sjostedt, A.; Garg, P. Structure-based virtual screening to
discover potential lead molecules for the SARS-CoV-2 main protease.
J. Chem. Inf. Model. 2020,60, 57815793.
(18) Kumar, S.; Sharma, P. P.; Shankar, U.; Kumar, D.; Joshi, S. K.;
Pena, L.; Durvasula, R.; Kumar, A.; Kempaiah, P.; Poonam; Rathi, B.
Discovery of new hydroxyethylamine analogs against 3CL(pro)
protein target of SARS-CoV-2: Molecular docking, molecular
dynamics simulation, and structure-activity relationship studies. J.
Chem. Inf. Model. 2020,60, 57545770.
(19) Ngo, S. T.; Quynh Anh Pham, N.; Thi Le, L.; Pham, D. H.; Vu,
V. V. Computational determination of potential inhibitors of SARS-
CoV-2 main protease. J. Chem. Inf. Model. 2020,60, 57715780.
(20) Świderek, K.; Moliner, V. Revealing the molecular mechanisms
of proteolysis of SARS-CoV-2 Mpro by QM/MM computational
methods. Chem. Sci. 2020,11, 1062610630.
(21) Wang, H.; He, S.; Deng, W.; Zhang, Y.; Li, G.; Sun, J.; Zhao,
W.; Guo, Y.; Yin, Z.; Li, D. Comprehensive insights into the catalytic
mechanism of middle east respiratory syndrome 3c-like protease and
severe acute respiratory syndrome 3c-like protease. ACS Catal. 2020,
10, 58715890.
(22) Khrenova, M. G.; Tsirelson, V. G.; Nemukhin, A. V. Dynamical
properties of enzymesubstrate complexes disclose substrate
specificity of the SARS-CoV-2 main protease as characterized by
the electron density descriptors. Phys. Chem. Chem. Phys. 2020,22,
1906919079.
(23) Verma, N.; Henderson, J. A.; Shen, J. Proton-coupled
conformational activation of sars coronavirus main proteases and
opportunity for designing small-molecule broad-spectrum targeted
covalent inhibitors. J. Am. Chem. Soc. 2020,142, 2188321890.
(24) Ramos-Guzmań, C. A.; Ruiz-Pernía, J. J.; Tuñoń,I.a.
Unraveling the SARS-CoV-2 main protease mechanism using
multiscale methods. ACS Catal. 2020,10, 1254412554.
(25) Mondal, D.; Warshel, A. Exploring the mechanism of covalent
inhibition: Simulating the binding free energy of α-ketoamide
Journal of Chemical Information and Modeling pubs.acs.org/jcim Article
https://doi.org/10.1021/acs.jcim.1c00263
J. Chem. Inf. Model. 2021, 61, 29572966
2963
inhibitors of the main protease of sars-cov-2. Biochemistry 2020,59,
46014608.
(26) Ramos-Guzmán, C. A.; Ruiz-Pernía, J. J.; Tuñón, I. A
microscopic description of SARS-CoV-2 main protease inhibition
with michael acceptors. Strategies for improving inhibitors design.
Chem. Sci. 2021,12, 34893496.
(27) Jaffrelot-Inizan, T.; Célerse, F.; Adjoua, O.; Ahdab, D. E.; Jolly,
L.-H.; Liu, C.; Ren, P.; Montes, M.; Lagarde, N.; Lagardère, L.;
Monmarché, P.; Piquemal, J.-P. High-resolution mining of SARS-
CoV-2 main protease conformational space: Supercomputer-driven
unsupervised adaptive sampling. Chem. Sci. 2021,12, 48894907.
(28) Awoonor-Williams, E.; Abu-Saleh, A. A.-A. A. Covalent and
non-covalent binding free energy calculations for peptidomimetic
inhibitors of SARS-CoV-2 main protease. Phys. Chem. Chem. Phys.
2021,23, 67466757.
(29) Arafet, K.; Serrano-Aparicio, N.; Lodola, A.; Mulholland, A. J.;
González, F. V.; Świderek, K.; Moliner, V. Mechanism of inhibition of
SARS-CoV-2 Mpro by n3 peptidyl michael acceptor explained by QM/
MM simulations and design of new derivatives with tunable chemical
reactivity. Chem. Sci. 2021,12, 14331444.
(30)Chodera,J.;Lee,A.A.;London,N.;vonDelft,F.
Crowdsourcing drug discovery for pandemics. Nat. Chem. 2020,12,
581.
(31) Kuntz, I. D.; Blaney, J. M.; Oatley, S. J.; Langridge, R.; Ferrin,
T. E. A geometric approach to macromolecule-ligand interactions. J.
Mol. Biol. 1982,161, 269288.
(32) Shoichet, B. K.; Kuntz, I. D. Protein docking and
complementarity. J. Mol. Biol. 1991,221, 327346.
(33) Jones, G.; Willett, P. Docking small-molecule ligands into
active-sites. Curr. Opin. Biotechnol. 1995,6, 652656.
(34) Brooijmans, N.; Kuntz, I. D. Molecular recognition and docking
algorithms. Annu. Rev. Biophys. Biomol. Struct. 2003,32, 335373.
(35) London, N.; Miller, R. M.; Krishnan, S.; Uchida, K.; Irwin, J. J.;
Eidam, O.; Gibold, L.; Cimermančič, P.; Bonnet, R.; Shoichet, B. K.;
Taunton, J. Covalent docking of large libraries for the discovery of
chemical probes. Nat. Chem. Biol. 2014,10, 10661072.
(36) Goodsell, D. S.; Olson, A. J. Automated docking of substrates
to proteins by simulated annealing. Proteins 1990,8, 195202.
(37) Morris, G. M.; Goodsell, D. S.; Huey, R.; Olson, A. J.
Distributed automated docking of flexible ligands to proteins: Parallel
applications of autodock 2.4. J. Comput. Aided Mol. Des. 1996,10,
293304.
(38) Morris, G. M.; Huey, R.; Lindstrom, W.; Sanner, M. F.; Belew,
R. K.; Goodsell, D. S.; Olson, A. J. Autodock4 and autodocktools4:
Automated docking with selective receptor flexibility. J. Comput.
Chem. 2009,30, 27852791.
(39) Trott, O.; Olson, A. J. Autodock vina: Improving the speed and
accuracy of docking with a new scoring function, efficient
optimization, and multithreading. J. Comput. Chem. 2010,31, 455
461.
(40) Halgren, T. A.; Murphy, R. B.; Friesner, R. A.; Beard, H. S.;
Frye, L. L.; Pollard, W. T.; Banks, J. L. Glide: A new approach for
rapid, accurate docking and scoring. 2. Enrichment factors in database
screening. J. Med. Chem. 2004,47, 17501759.
(41) Friesner, R. A.; Murphy, R. B.; Repasky, M. P.; Frye, L. L.;
Greenwood, J. R.; Halgren, T. A.; Sanschagrin, P. C.; Mainz, D. T.
Extra precision glide: Docking and scoring incorporating a model of
hydrophobic enclosure for protein-ligand complexes. J. Med. Chem.
2006,49, 61776196.
(42) Repasky, M. P.; Murphy, R. B.; Banks, J. L.; Greenwood, J. R.;
Tubert-Brohman, I.; Bhat, S.; Friesner, R. A. Docking performance of
the glide program as evaluated on the astex and dud datasets: A
complete set of glide sp results and selected results for a new scoring
function integrating watermap and glide. J. Comput. Aided Mol. Des.
2012,26, 787799.
(43) Davis, I. W.; Baker, D. Rosettaligand docking with full ligand
and receptor flexibility. J. Mol. Biol. 2009,385, 381392.
(44) Davis, I. W.; Raha, K.; Head, M. S.; Baker, D. Blind docking of
pharmaceutically relevant compounds using rosettaligand. Protein Sci.
2009,18, 19982002.
(45) Willett, P.; Glen, R. C.; Leach, A. R.; Taylor, R.; Jones, G.
Development and validation of a genetic algorithm for flexible
docking. J. Mol. Biol. 1997,267, 727748.
(46) Verdonk, M. L.; Cole, J. C.; Hartshorn, M. J.; Murray, C. W.;
Taylor, R. D. Improved protein-ligand docking using GOLD. Proteins
2003,52, 609623.
(47) Verdonk, M. L.; Chessari, G.; Cole, J. C.; Hartshorn, M. J.;
Murray, C. W.; Nissink, J. W.; Taylor, R. D.; Taylor, R. Modeling
water molecules in protein-ligand docking using GOLD. J. Med. Chem.
2005,48, 65046515.
(48) Vieth, M.; Hirst, J. D.; Dominy, B. N.; Daigler, H.; Brooks, C.
L., III Assessing search strategies for flexible docking. J. Comput.
Chem. 1998,19, 16231631.
(49) Vieth, M.; Hirst, J. D.; Kolinski, A.; Brooks, C. L., III Assessing
energy functions for flexible docking. J. Comput. Chem. 1998,19,
16121622.
(50) Wu, G.; Robertson, D. H.; Brooks, C. L., III; Vieth, M. Detailed
analysis of grid-based molecular docking: A case study of CDOCKER-
a CHARMm-based md docking algorithm. J. Comput. Chem. 2003,24,
15491562.
(51) Gagnon, J. K.; Law, S. M.; Brooks, C. L. III Flexible
CDOCKER: Development and application of a pseudo-explicit
structure-based docking method within CHARMM. J. Comput.
Chem. 2016,37, 753762.
(52) MacKerell, A. D.; Bashford, D.; Bellot, M.; Dunbrack, J. R. L.;
Evanseck, J. D.; Field, M. J.; Fischer, S.; Gao, J.; Guo, H.; Ha, S.;
Joseph-McCarthy, D.; Kuchnir, L.; Kuczera, K.; Lau, F. T. K.; Mattos,
C.; Michnick, S.; Ngo, T.; Nguyen, D. T.; Prodhom, B.; Reiher, W. E.,
III; Roux, B.; Schlenkrich, M.; Smith, J. C.; Stote, R.; Straub, J.;
Watanabe, M.; Kuczera, J.; Yin, D.; Karplus, M. All-atom empirical
potential for molecular modeling and dynamics studies of proteins. J.
Phys. Chem. B 1998,102, 35863616.
(53) Vanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.;
Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I.;
Mackerell, A. D., Jr. CHARMM general force field: A force field for
drug-like molecules compatible with the CHARMM all-atom additive
biological force fields. J. Comput. Chem. 2010,31, 671690.
(54) Huang, J.; MacKerell, A. D., Jr. CHARMM36 all-atom additive
protein force field: Validation based on comparison to nmr data. J.
Comput. Chem. 2013,34, 21352145.
(55) Huang, J.; Rauscher, S.; Nawrocki, G.; Ran, T.; Feig, M.; Groot,
B. L. d.; Grubmüller, H.; Alexander D. MacKerell, J. Charmm36m: An
improved force field for folded and intrinsically disordered proteins.
Nat. Methods 2017,14,7173.
(56) Lin, F.-Y.; Alexander D. MacKerell, J. Improved modeling of
halogenated ligandprotein interactions using the drude polarizable
and CHARMM additive empirical force fields. J. Chem. Inf. Model.
2019,59, 215228.
(57) Cornell, W. D.; Cieplak, P.; Bayly, C. I.; Gould, I. R.; Merz, K.
M.; Ferguson, D. M.; Spellmeyer, D. C.; Fox, T.; Caldwell, J. W.;
Kollman, P. A. A second generation force field for the simulation of
proteins, nucleic acids, and organic molecules. J. Am. Chem. Soc. 1995,
117, 51795197.
(58) Jorgensen, W. L.; Tirado-Rives, J. The OPLS force field for
proteins. Energy minimizations for crystals of cyclic peptides and
crambin. J. Am. Chem. Soc. 1988,110, 16571666.
(59) Jorgensen, W. L.; Maxwell, D. S.; Tirado-Rives, J. Development
and testing of the OPLS all-atom force field on conformational
energetics and properties of organic liquids. J. Am. Chem. Soc. 1996,
118, 1122511236.
(60) Velec, H. F.; Gohlke, H.; Klebe, G. Drugscore(CSD)-
knowledge-based scoring function derived from small molecule crystal
data with superior recognition rate of near-native ligand poses and
better affinity prediction. J. Med. Chem. 2005,48, 62966303.
(61) Grinter, S. Z.; Yan, C.; Huang, S. Y.; Jiang, L.; Zou, X.
Automated large-scale file preparation, docking, and scoring:
Journal of Chemical Information and Modeling pubs.acs.org/jcim Article
https://doi.org/10.1021/acs.jcim.1c00263
J. Chem. Inf. Model. 2021, 61, 29572966
2964
Evaluation of itscore and stscore using the 2012 community structure-
activity resource benchmark. J. Chem. Inf. Model. 2013,53, 1905
1914.
(62) Neudert, G.; Klebe, G. DSX: A knowledge-based scoring
function for the assessment of protein-ligand complexes. J. Chem. Inf.
Model. 2011,51, 27312745.
(63) Moitessier, N.; Englebienne, P.; Lee, D.; Lawandi, J.; Corbeil,
C. R. Towards the development of universal, fast and highly accurate
docking/scoring methods: A long way to go. Br. J. Pharmacol. 2008,
153,S7S26.
(64) Ishchenko, A. V.; Shakhnovich, E. I. SMall Molecule Growth
2001 (SMoG2001): An improved knowledge-based scoring function
for protein-ligand interactions. J. Med. Chem. 2002,45, 27702780.
(65) OBrien, T. E.; Bertolani, S. J.; Tantillo, D. J.; Siegel, J. B.
Mechanistically informed predictions of binding modes for
carbocation intermediates of a sesquiterpene synthase reaction.
Chem. Sci. 2016,7, 40094015.
(66) OBrien, T. E.; Bertolani, S. J.; Zhang, Y.; Siegel, J. B.; Tantillo,
D. J. Predicting productive binding modes for substrates and
carbocation intermediates in terpene synthases-bornyl diphosphate
synthase as a representative case. ACS Catal. 2018,8, 33223330.
(67) Das, S.; Shimshi, M.; Raz, K.; Nitoker, N.; Mhashal, A. R.;
Ansbacher, T.; Major, D. T. EnzyDock: Protein-ligand docking of
multiple reactive states along a reaction coordinate in enzymes. J.
Chem. Theory Comput. 2019,15, 51165134.
(68) Kokh, D. B.; Wade, R. C.; Wenzel, W. Receptor flexibility in
small-molecule docking calculations. Wiley Interdiscip. Rev.: Comput.
Mol. Sci. 2011,1, 298314.
(69) Osterberg, F.; Morris, G. M.; Sanner, M. F.; Olson, A. J.;
Goodsell, D. S. Automated docking to multiple target structures:
Incorporation of protein mobility and structural water heterogeneity
in autodock. Proteins 2002,46,3440.
(70) Ton, A. T.; Gentile, F.; Hsing, M.; Ban, F.; Cherkasov, A. Rapid
identification of potential inhibitors of SARS-CoV-2 main protease by
deep docking of 1.3 billion compounds. Mol. Inf. 2020,39,
No. e2000028.
(71) Fischer, A.; Sellner, M.; Neranjan, S.; Smiesko, M.; Lill, M. A.
Potential inhibitors for novel coronavirus protease identified by virtual
screening of 606 million compounds. Int. J. Mol. Sci. 2020,21, 3626.
(72) Achilonu, I.; Iwuchukwu, E. A.; Achilonu, O. J.; Fernandes, M.
A.; Sayed, Y. Targeting the SARS-CoV-2 main protease using FDA-
approved isavuconazonium, a p2-p3 alpha-ketoamide derivative and
pentagastrin: An in-silico drug discovery approach. J. Mol. Graphics
Modell. 2020,101, 107730.
(73) Gentile, D.; Patamia, V.; Scala, A.; Sciortino, M. T.; Piperno,
A.; Rescifina, A. Putative inhibitors of SARS-CoV-2 main protease
from a library of marine natural products: A virtual screening and
molecular modeling study. Mar. Drugs 2020,18, 225.
(74) Shivanika, C.; Deepak Kumar, S.; Ragunathan, V.; Tiwari, P.;
Sumitha, A.; Brindha Devi, P. Molecular docking, validation, dynamics
simulations, and pharmacokinetic prediction of natural compounds
against the SARS-CoV-2 main-protease. J. Biomol. Struct. Dyn. 2020,
127.
(75) Keretsu, S.; Bhujbal, S. P.; Cho, S. J. Rational approach toward
COVID-19 main protease inhibitors via molecular docking, molecular
dynamics simulation and free energy calculation. Sci. Rep. 2020,10,
17716.
(76) Ferraz, W. R.; Gomes, R. A.; Novaes, A. L. S.; Trossini, G. H.
G. Ligand and structure-based virtual screening applied to the SARS-
CoV-2 main protease: An in silico repurposing study. Future Med.
Chem. 2020,12, 18151828.
(77) Stoddard, S. V.; Stoddard, S. D.; Oelkers, B. K.; Fitts, K.;
Whalum, K.; Whalum, K.; Hemphill, A. D.; Manikonda, J.; Martinez,
L. M.; Riley, E. G.; Roof, C. M.; Sarwar, N.; Thomas, D. M.; Ulmer,
E.; Wallace, F. E.; Pandey, P.; Roy, S. Optimization rules for SARS-
CoV-2 Mpro antivirals: Ensemble docking and exploration of the
coronavirus protease active site. Viruses 2020,12, 942.
(78) Maffucci, I.; Contini, A. In silico drug repurposing for SARS-
CoV-2 main proteinase and spike proteins. J. Proteome Res. 2020,19,
46374648.
(79) Nguyen, D. D.; Gao, K.; Chen, J.; Wang, R.; Wei, G.-W.
Unveiling the molecular mechanism of SARS-CoV-2 main protease
inhibition from 137 crystal structures using algebraic topology and
deep learning. Chem. Sci. 2020,11, 1203612046.
(80) Mortenson, P. N. Diverse, high-quality test set for the
validation of protein-ligand native and cross-docking performance. J.
Med. Chem. 2007,50, 726741.
(81) Plewczynski, D.; Łaźniewski, M.; Augustyniak, R.; Ginalski, K.
Can we trust docking results? Evaluation of seven commonly used
programs on PDBbind database. J. Comput. Chem. 2011,32, 742
755.
(82) Chen, Y. C. Beware of docking! Trends Pharmacol. Sci 2015,36,
7895.
(83) Fu, D. Y.; Meiler, J. Predictive power of different types of
experimental restraints in small molecule docking: A review. J. Chem.
Inf. Model. 2018,58, 225233.
(84) Hawkins, P. C. D.; Skillman, A. G.; Nicholls, A. Comparison of
shape-matching and docking as virtual screening tools. J. Med. Chem.
2007,50,7482.
(85) McGann, M. FRED pose prediction and virtual screening
accuracy. J. Chem. Inf. Model. 2011,51, 578596.
(86) McGann, M. FRED and hybrid docking performance on
standardized datasets. J. Comput. Aided Mol. Des. 2012,26, 897906.
(87) Kelley, B. P.; Brown, S. P.; Warren, G. L.; Muchmore, S. W.
POSIT: Flexible shape-guided docking for pose prediction. J. Chem.
Inf. Model. 2015,55, 17711780.
(88) Berman, H. M.; Henrick, K.; Nakamura, H. Announcing the
worldwide protein data bank. Nat. Struct. Biol. 2003,10, 980.
(89) Šali, A.; Blundell, T. L. Comparative protein modelling by
satisfaction of spatial restraints. J. Mol. Biol. 1993,234, 779815.
(90) Brooks, B. R.; Bruccoleri, R. E.; Olafson, B. D.; States, D. J.;
Swaminathan, S.; Karplus, M. CHARMM - a program for macro-
molecular energy, minimization, and dynamics calculations. J. Comput.
Chem. 1983,4, 187217.
(91) Brooks, B. R.; Brooks, C. L., III; Mackerell, A. D., Jr.; Nilsson,
L.; Petrella, R. J.; Roux, B.; Won, Y.; Archontis, G.; Bartels, C.;
Boresch, S.; Caflisch, A.; Caves, L.; Cui, Q.; Dinner, A. R.; Feig, M.;
Fischer, S.; Gao, J.; Hodoscek, M.; Im, W.; Kuczera, K.; Lazaridis, T.;
Ma, J.; Ovchinnikov, V.; Paci, E.; Pastor, R. W.; Post, C. B.; Pu, J. Z.;
Schaefer, M.; Tidor, B.; Venable, R. M.; Woodcock, H. L.; Wu, X.;
Yang, W.; York, D. M.; Karplus, M. CHARMM: The biomolecular
simulation program. J. Comput. Chem. 2009,30, 15451614.
(92) Zhu, K.; Borrelli, K. W.; Greenwood, J. R.; Day, T.; Abel, R.;
Farid, R. S.; Harder, E. Docking covalent inhibitors: A parameter free
approach to pose prediction and scoring. J. Chem. Inf. Model. 2014,
54, 19321940.
(93) Harder, E.; Damm, W.; Maple, J.; Wu, C.; Reboul, M.; Xiang, J.
Y.; Wang, L.; Lupyan, D.; Dahlgren, M. K.; Knight, J. L.; Kaus, J. W.;
Cerutti, D. S.; Krilov, G.; Jorgensen, W. L.; Abel, R.; Friesner, R. A.
OPLS3: A force field providing broad coverage of drug-like small
molecules and proteins. J. Chem. Theory Comput. 2016,12, 281296.
(94) Pettersen, E. F.; Goddard, T. D.; Huang, C. C.; Couch, G. S.;
Greenblatt, D. M.; Meng, E. C.; Ferrin, T. E. UCSF Chimeraa
visualization system for exploratory research and analysis. J. Comput.
Chem. 2004,25, 16051612.
(95) Gasteiger, J.; Marsili, M. Iterative partial equalization of orbital
electronegativitya rapid access to atomic charges. Tetrahedron
1980,36, 32193228.
(96) Bianco, G.; Forli, S.; Goodsell, D. S.; Olson, A. J. Covalent
docking using autodock: Two-point attractor and flexible side chain
methods. Protein Sci. 2016,25, 295301.
(97) Yu, W.; He, X.; Vanommeslaeghe, K.; Alexander D. MacKerell,
J. Extension of the CHARMM general force field to sulfonyl-
containing compounds and its utility in biomolecular simulations. J.
Comput. Chem. 2012,33, 24512468.
Journal of Chemical Information and Modeling pubs.acs.org/jcim Article
https://doi.org/10.1021/acs.jcim.1c00263
J. Chem. Inf. Model. 2021, 61, 29572966
2965
(98) Wang, L.; Berne, B. J.; Friesner, R. A. On achieving high
accuracy and reliability in the calculation of relative proteinligand
binding affinities. Proc. Natl. Acad. Sci. U.S.A. 2012,109, 19371942.
(99) Zhang, C.-H.; Stone, E. A.; Deshmukh, M.; Ippolito, J. A.;
Ghahremanpour, M. M.; Tirado-Rives, J.; Spasov, K. A.; Zhang, S.;
Takeo, Y.; Kudalkar, S. N.; Liang, Z.; Isaacs, F.; Lindenbach, B.;
Miller, S. J.; Anderson, K. S.; Jorgensen, W. L. Potent noncovalent
inhibitors of the main protease of SARS-CoV-2 from molecular
sculpting of the drug perampanel guided by free energy perturbation
calculations. ACS Cent. Sci. 2021,7, 467475.
Journal of Chemical Information and Modeling pubs.acs.org/jcim Article
https://doi.org/10.1021/acs.jcim.1c00263
J. Chem. Inf. Model. 2021, 61, 29572966
2966
... This can be caused by the inherent limitations of docking methods, including approximations in representing isolated systems and the omission of factors like solvation [75,76], combined with the complexity of the biological target. In agreement, multiple docking protocols for Mpro have shown unsatisfactory performance [39,78]. Thus, clarification of the binding modes of compounds 13-13c will require the generation of experimental structures in future work. ...
... The discrepancy between enzymatic inhibition and antiviral efficacy could be caused by a reduced stability of the compound in cell culture, permeability issues, increased efflux transport in Vero cells, or buffering by host or viral off-target proteins. Compound 13a showed the best antiviral effect among the candidates with an EC 50 [39,78]. Thus, clarification of the binding modes of compounds 13-13c will require the generation of experimental structures in future work. ...
Article
Full-text available
SARS-CoV-2 Main Protease (Mpro) is an enzyme that cleaves viral polyproteins translated from the viral genome, which is critical for viral replication. Mpro is a target for anti-SARS-CoV-2 drug development. Herein, we performed a large-scale virtual screening by comparing multiple structural descriptors of reference molecules with reported anti-coronavirus activity against a library with >17 million compounds. Further filtering, performed by applying two machine learning algorithms, identified eighteen computational hits as anti-SARS-CoV-2 compounds with high structural diversity and drug-like properties. The activities of twelve compounds on Mpro's enzymatic activity were evaluated by fluorescence resonance energy transfer (FRET) assays. Compound 13 (ZINC13878776) significantly inhibited SARS-CoV-2 Mpro activity and was employed as a reference for an experimentally hit expansion. The structural analogues 13a (ZINC4248385), 13b (ZNC13523222), and 13c (ZINC4248365) were tested as Mpro inhibitors, reducing the enzymatic activity of recombinant Mpro with potency as follows: 13c > 13 > 13b > 13a. Then, their anti-SARS-CoV-2 activities were evaluated in plaque reduction assays using Vero CCL81 cells. Subtoxic concentrations of compounds 13a, 13c, and 13b displayed in vitro antiviral activity with IC 50 in the mid micromolar range. Compounds 13a-c could become lead compounds for the development of new Mpro inhibitors with improved activity against anti-SARS-CoV-2.
... Molecular docking simulations were performed using three different software: AutoDock Vina and RDock, which are open-source programs, as well as Glide, which is a commercially available program. Different molecular docking software exhibited varying performances when applied to the same PDB structures, which is consistent with previously published results 26 (Fig. 4). When evaluating virtual screening, we primarily used three widely studied metrics: Enrichment Factors (EF), Area Under the Receiver Operating Characteristic Curve (ROC-AUC), and Boltzmann-Enhanced Discrimination of ROC (BEDROC). ...
Article
Full-text available
The coronavirus disease 19 pandemic, caused by severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), has led to a global health crisis with millions of confirmed cases and related deaths. The main protease (Mpro) of SARS-CoV-2 is crucial for viral replication and presents an attractive target for drug development. Despite the approval of some drugs, the search for effective treatments continues. In this study, we systematically evaluated 342 holo-crystal structures of Mpro to identify optimal conformations for structure-based virtual screening (SBVS). Our analysis revealed limited structural flexibility among the structures. Three docking programs, AutoDock Vina, rDock, and Glide were employed to assess the efficiency of virtual screening, revealing diverse performances across selected Mpro structures. We found that the structures 5RHE, 7DDC, and 7DPU (PDB Ids) consistently displayed the lowest EF, AUC, and BEDROCK scores. Furthermore, these structures demonstrated the worst pose prediction results in all docking programs. Two structural differences contribute to variations in docking performance: the absence of the S1 subsite in 7DDC and 7DPU, and the presence of a subpocket in the S2 subsite of 7DDC, 7DPU, and 5RHE. These findings underscore the importance of selecting appropriate Mpro conformations for SBVS, providing valuable insights for advancing drug discovery efforts.
... In comparison, molecular docking, particularly when targeting M pro , has shown limited success in accurately replicating crystallized conformations, with success rates rarely exceeding 26%. 75,76 This comparison highlights the superior performance of our consensus pharmacophore model in predicting ligand poses, confirming its potential utility in the discovery and optimization of novel inhibitors for SARS-CoV-2 M pro . ...
Article
Full-text available
The severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) main Protease (Mpro) is an enzyme that cleaves viral polyproteins translated from the viral genome and is critical for viral replication. Mpro is a target for anti-SARS-CoV-2 drug development, and multiple Mpro crystals complexed with competitive inhibitors have been reported. In this study, we aimed to develop an Mpro consensus pharmacophore as a tool to expand the search for inhibitors. We generated a consensus model by aligning and summarizing pharmacophoric points from 152 bioactive conformers of SARS-CoV-2 Mpro inhibitors. Validation against a library of conformers from a subset of ligands showed that our model retrieved poses that reproduced the crystal-binding mode in 77% of the cases. Using models derived from a consensus pharmacophore, we screened >340 million compounds. Pharmacophore-matching and chemoinformatics analyses identified new potential Mpro inhibitors. The candidate compounds were chemically dissimilar to the reference set, and among them, demonstrating the relevance of our model. We evaluated the effect of 16 candidates on Mpro enzymatic activity finding that seven have inhibitory activity. Three compounds (1, 4, and 5) had IC50 values in the midmicromolar range. The Mpro consensus pharmacophore reported herein can be used to identify compounds with improved activity and novel chemical scaffolds against Mpro. The method developed for its generation is provided as an open-access code (https://github.com/AngelRuizMoreno/ConcensusPharmacophore) and can be applied to other pharmacological targets.
Article
Full-text available
This review highlights methods for studying structure activity relationships of natural products and proposes that these methods are complementary and could be used to build an iterative computational-experimental workflow.
Article
Full-text available
Alchemical binding free energy calculations are one of the most accurate methods for estimating ligand-binding affinity. Assessing the accuracy of the approach over protein targets is one of the most interesting issues. The free energy difference of binding between a protein and a ligand was calculated via the alchemical approach. The alchemical approach exhibits satisfactory accuracy over four targets, including AmpC beta-lactamase (AmpC); glutamate receptor, ionotropic kainate 1 (GluK1); heat shock protein 90 (Hsp90); and severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) main protease (Mpro). In particular, the correlation coefficients between calculated binding free energies and the respective experiments over four targets range from 0.56 to 0.86. The affinity computed via free energy perturbation (FEP) simulations is overestimated over the experimental value. Particularly, the electrostatic interaction free energy rules the binding process of ligands to AmpC and GluK1. However, the van der Waals (vdW) interaction free energy plays an important role in the ligand-binding processes of HSP90 and SARS-CoV-2 Mpro. The obtained results associate with the hydrophilic or hydrophobic properties of the ligands. This observation may enhance computer-aided drug design.
Article
Full-text available
Molecular docking has become an increasingly important tool for drug discovery. In this review, we present a brief introduction of the available molecular docking methods, and their development and applications in drug discovery. Molecular docking has been widely employed as a fast and inexpensive technique in the past decades, both in academic and industrial settings. Although this discipline has now had enough time to consolidate, many aspects remain challenging and there is still not a straightforward and accurate route to readily pinpoint true ligands among a set of molecules, nor to identify with precision the correct ligand conformation within the binding pocket of a given target molecule. Nevertheless, new approaches continue to be developed and the volume of published works grows at a rapid pace. These recent developments incrementally contribute to an increase in accuracy and are expected, given time, and together with advances in computing power and hardware capability, to eventually accomplish the full potential of this area. Applications of molecular docking in drug development have evolved significantly since it was first created to aid in the study of molecular recognition processes between small and large compounds.
Chapter
Molecular docking is a popular computational tool in drug discovery. Leveraging structural information, docking software predicts binding poses of small molecules to cavities on the surfaces of proteins. Virtual screening for ligand discovery is a useful application of docking software. In this chapter, using the enigmatic KRAS protein as an example system, we endeavor to teach the reader about best practices for performing molecular docking with UCSF DOCK. We discuss methods for virtual screening and docking molecules on KRAS. We present the following six points to optimize our docking setup for prosecuting a virtual screen: protein structure choice, pocket selection, optimization of the scoring function, modification of sampling spheres and sampling procedures, choosing an appropriate portion of chemical space to dock, and the choice of which top scoring molecules to pick for purchase.
Article
Full-text available
A large-scale screen to target SARS-CoV-2 The severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) genome is initially expressed as two large polyproteins. Its main protease, M pro , is essential to yield functional viral proteins, making it a key drug target. Günther et al. used x-ray crystallography to screen more than 5000 compounds that are either approved drugs or drugs in clinical trials. The screen identified 37 compounds that bind to M pro . High-resolution structures showed that most compounds bind at the active site but also revealed two allosteric sites where binding of a drug causes conformational changes that affect the active site. In cell-based assays, seven compounds had antiviral activity without toxicity. The most potent, calpeptin, binds covalently in the active site, whereas the second most potent, pelitinib, binds at an allosteric site. Science , this issue p. 642
Article
Full-text available
The design of novel compounds targeting the SARS-CoV-2 Mpro that are potent, drug-like, and work synergistically with other direct-acting antivirals is of vital importance.
Article
Full-text available
Starting from our previous finding of 14 known drugs as inhibitors of the main protease (Mpro) of SARS-CoV-2, the virus responsible for COVID-19, we have redesigned the weak hit perampanel to yield multiple noncovalent, nonpeptidic inhibitors with ca. 20 nM IC50 values in a kinetic assay. Free-energy perturbation (FEP) calculations for Mpro-ligand complexes provided valuable guidance on beneficial modifications that rapidly delivered the potent analogues. The design efforts were confirmed and augmented by determination of high-resolution X-ray crystal structures for five analogues bound to Mpro. Results of cell-based antiviral assays further demonstrated the potential of the compounds for treatment of COVID-19. In addition to the possible therapeutic significance, the work clearly demonstrates the power of computational chemistry for drug discovery, especially FEP-guided lead optimization.
Article
Full-text available
Targeting the SARS-CoV-2 main protease Vaccines are an important tool in the fight against COVID-19, but developing antiviral drugs is also a high priority, especially with the rise of variants that may partially evade vaccines. The viral protein main protease is required for cleaving precursor polyproteins into functional viral proteins. This essential function makes it a key drug target. Qiao et al. designed 32 inhibitors based on either boceprevir or telaprevir, both of which are protease inhibitors approved to treat hepatitis C virus. Six compounds protected cells from viral infection with high potency, and two of these were selected for in vivo studies based on pharmokinetic experiments. Both showed strong antiviral activity in a mouse model. Science , this issue p. 1374
Article
Full-text available
We provide an unsupervised adaptive sampling strategy capable of producing μs-timescale molecular dynamics (MD) simulations of large biosystems using many-body polarizable force fields (PFF). The global exploration problem is decomposed into a set of separate MD trajectories that can be restarted within a selective process to achieve sufficient phase-space sampling. Accurate statistical properties can be obtained through reweighting. Within this highly parallel setup, the Tinker--HP package can be powered by an arbitrary large number of GPUs on supercomputers, reducing exploration time from years to days. This approach is used to tackle the urgent modeling problem of the SARS--CoV--2 Main Protease (Mpro) producing more than 38 μs of all-atom simulations of its apo, ligand-free, dimer using the high-resolution AMOEBA PFF. A first 15.14 μs simulation (physiological pH) is compared to available non--PFF long-timescale simulation data. A detailed clustering analysis exhibits striking differences between FFs, AMOEBA showing a richer conformational space. Focusing on key structural markers related to the oxyanion hole stability, we observe an asymmetry between protomers. One of them appears less structured resembling the experimentally inactive monomer for which a 6 s simulation was performed as a basis of comparison. Results highlight the plasticity of Mpro active site. The C--terminal end of its less structured protomer is shown to oscillate between several states, being able to interact with the other protomer, potentially modulating its activity. Active and distal sites volumes are found to be larger in the most active protomer within our AMOEBA simulations compared to non-PFFs as additional cryptic pockets are uncovered. A second 17 μs AMOEBA simulation is performed with protonated His172 residues mimicking lower pH. Data show the protonation impact on the destructuring of the oxyanion loop. We finally analyze the solvation patterns around key histidine residues. The confined AMOEBA polarizable water molecules are able to explore a wide range of dipole moments, going beyond bulk values, leading to a water molecule counts consistent with experiment. Results suggest that the use of PFFs could be critical in drug discovery to accurately model the complexity of the molecular interactions structuring Mpro.
Article
Full-text available
The irreversible inhibition of the main protease of SARS-CoV-2 by a Michael acceptor known as N3 has been investigated using multiscale methods. The noncovalent enzyme-inhibitor complex was simulated using classical Molecular Dynamics techniques and the pose of the inhibitor in the active site was compared to that of the natural substrate, a peptide containing the Gln-Ser scissile bond. The formation of the covalent enzyme-inhibitor complex was then simulated using hybrid QM/MM free energy methods. After binding, the reaction mechanism was found to be composed of two steps: i) the activation of the catalytic dyad (Cys145 and His41) to form an ion pair and ii) a Michael addition where the attack of the Sγ atom of Cys145 to the Cβ atom of the inhibitor precedes the water-mediated proton transfer from His41 to the Cα atom. The microscopic description of protease inhibition by N3 obtained from our simulations is strongly supported by the excellent agreement between the estimated activation free energy and the value derived from kinetic experiments. Comparison with the acylation reaction of a peptide substrate suggests that that N3-based inhibitors could be improved by adding chemical modifications that could facilitate the formation of the catalytic dyad ion pair.
Article
Full-text available
We present a supercomputer-driven pipeline for in silico drug discovery using enhanced sampling molecular dynamics (MD) and ensemble docking. Ensemble docking makes use of MD results by docking compound databases into representative protein binding-site conformations, thus taking into account the dynamic properties of the binding sites. We also describe preliminary results obtained for 24 systems involving eight proteins of the proteome of SARS-CoV-2. The MD involves temperature replica exchange enhanced sampling, making use of massively parallel supercomputing to quickly sample the configurational space of protein drug targets. Using the Summit supercomputer at the Oak Ridge National Laboratory, more than 1 ms of enhanced sampling MD can be generated per day. We have ensemble docked repurposing databases to 10 configurations of each of the 24 SARS-CoV-2 systems using AutoDock Vina. Comparison to experiment demonstrates remarkably high hit rates for the top scoring tranches of compounds. We also demonstrate that, using Autodock-GPU on Summit, it is possible to perform exhaustive docking of one billion compounds in under 24 h. Finally, we discuss preliminary results and planned improvements to the pipeline, including the use of quantum mechanical (QM), machine learning, and artificial intelligence (AI) methods to cluster MD trajectories and rescore docking poses.
Article
A good docking algorithm requires an energy function that is selective, in that it clearly differentiates correctly docked structures from misdocked ones, and that is efficient, meaning that a correctly docked structure can be identified quickly. We assess the selectivity and efficiency of a broad spectrum of energy functions, derived from systematic modifications of the CHARMM param19/toph19 energy function. In particular, we examine the effects of the dielectric constant, the solvation model, the scaling of surface charges, reduction of van der Waals repulsion, and nonbonded cutoffs. Based on an assessment of the energy functions for the docking of five different ligand–receptor complexes, we find that selective energy functions include a variety of distance-dependent dielectric models together with truncation of the nonbonded interactions at 8 Å. We evaluate the docking efficiency, the mean number of docked structures per unit of time, of the more selective energy functions, using a simulated annealing molecular dynamics protocol. The largest improvements in efficiency come from a reduction of van der Waals repulsion and a reduction of surface charges. We note that the most selective potential is quite inefficient, although a hierarchical approach can be employed to take advantage of both selective and efficient energy functions. © 1998 John Wiley & Sons, Inc. J Comput Chem 19: 1612–1622, 1998
Article
COVID-19, the disease caused by the newly discovered coronavirus — SARS-CoV-2, has created global health, social, and economic crisis. As of mid-January 2021, there are over 90 million confirmed cases...
Article
The SARS coronavirus 2 (SARS-CoV-2) main protease (Mpro) is an attractive broad-spectrum antiviral drug target. Despite the enormous progress in structure elucidation, the Mpro’s structure–function relationship remains poorly understood. Recently, a peptidomimetic inhibitor has entered clinical trial; however, small-molecule orally available antiviral drugs have yet to be developed. Intrigued by a long-standing controversy regarding the existence of an inactive state, we explored the proton-coupled dynamics of the Mpros of SARS-CoV-2 and the closely related SARS-CoV using a newly developed continuous constant pH molecular dynamics (MD) method and microsecond fixed-charge all-atom MD simulations. Our data supports a general base mechanism for Mpro’s proteolytic function. The simulations revealed that protonation of His172 alters a conserved interaction network that upholds the oxyanion loop, leading to a partial collapse of the conserved S1 pocket, consistent with the first and controversial crystal structure of SARS-CoV Mpro determined at pH 6. Interestingly, a natural flavonoid binds SARS-CoV-2 Mpro in the close proximity to a conserved cysteine (Cys44), which is hyper-reactive according to the CpHMD titration. This finding offers an exciting new opportunity for small-molecule targeted covalent inhibitor design. Our work represents a first step toward the mechanistic understanding of the proton-coupled structure–dynamics–function relationship of CoV Mpros; the proposed strategy of designing small-molecule covalent inhibitors may help accelerate the development of orally available broad-spectrum antiviral drugs to stop the current pandemic and prevent future outbreaks.