You are on page 1of 61

Subscriber access provided by Gothenburg University Library

Pharmaceutical Modeling
In Silico Screening-based Discovery of Novel Inhibitors
of Human Cyclic GMP-AMP Synthase: A Cross-validation
Study of Molecular Docking and Experimental Testing
Wenfeng Zhao, Muya Xiong, Xiaojing Yuan, Minjun Li, Hongbin Sun, and Yechun Xu
J. Chem. Inf. Model., Just Accepted Manuscript • DOI: 10.1021/acs.jcim.0c00171 • Publication Date (Web): 27 May 2020
Downloaded from pubs.acs.org on May 27, 2020

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a service to the research community to expedite the dissemination
of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in
full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully
peer reviewed, but should not be considered the official version of record. They are citable by the
Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore,
the “Just Accepted” Web site may not include all articles that will be published in the journal. After
a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web
site and published as an ASAP article. Note that technical editing may introduce minor changes
to the manuscript text and/or graphics which could affect content, and all legal disclaimers and
ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or
consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W.,


Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 60 Journal of Chemical Information and Modeling

1
2
3
4
5
6
7
8
In Silico Screening-based Discovery of Novel
9
10
11
Inhibitors of Human Cyclic GMP-AMP
12
13 Synthase: A Cross-validation Study of Molecular
14
15
16 Docking and Experimental Testing
17
18
19
20
Wenfeng Zhao†,‡,§,#, Muya Xiong‡,§,#, Xiaojing Yuan‡,§, Minjun Li⊥, Hongbing Sun*,†,
21
22 Yechun Xu*,‡,§
23
24
25
26 †Jiangsu Key Laboratory of Drug Discovery for Metabolic Disease and State Key
27
28
29 Laboratory of Natural Medicines, China Pharmaceutical University, Nanjing
30
31
210009, P.R. China
32
33
34
35 ‡CAS Key Laboratory of Receptor Research, Drug Discovery and Design Center,
36
37
38 Shanghai Institute of Materia Medica, Chinese Academy of Sciences, Shanghai
39
40
41 201203, China
42
43
44 §University of Chinese Academy of Sciences, Beijing 100049, China
45
46
47
48 ⊥Shanghai Synchrotron Radiation Facility, Shanghai Advanced Research
49
50
51 Institute, Chinese Academy of Sciences, Shanghai 201210, China
52
53
54
55
56
57 1
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 2 of 60

1
2
3
4 KEYWORDS: cGAS, crystal structures of complexes, inhibitors, molecular docking,
5
6
7 virtual screening, MD simulations.
8
9
10
11 ABSTRACT: Cyclic GMP-AMP synthase (cGAS) is recently uncovered to be a
12
13
14 promising therapeutic target for immune-associated diseases. Up to now, only a
15
16
17 few inhibitors were identified through high-throughput screening campaigns. Here,
18
19 we reported the discovery of novel inhibitors for the catalytic domain of human
20
21
22 cGAS (h-cGASCD) by virtual screening for the first time. To generate a reliable
23
24
25 docking mode, we first obtained a high-resolution crystal structure of h-cGASCD in
26
27
complex with PF-06928215, a known inhibitor of h-cGAS, followed by MD
28
29
30 simulations on this complex structure. Four fragment hits were identified by the
31
32
33 virtual screening together with a thermal shift assay. Crystal structures of these four
34
35
36 compounds in complex with h-cGASCD were subsequently determined and the
37
38 binding modes of the compounds were similar to those predicted by molecular
39
40
41 docking, supporting the reliability of the docking model. In addition, an enzyme
42
43
44 activity assay identified compound 18 (IC50 = 29.88 ± 3.20 μM) from the compounds
45
46
predicted by the virtual screening. A similarity search of compound 18 followed by
47
48
49 a second virtual screening led to the discovery of compounds S2 (IC50 =13.1 ± 0.09
50
51
52 μM) and S3 (IC50 =4.9 ± 0.26 μM) as h-cGAS inhibitors with improved potency.
53
54
55 Therefore, the present study not only provides the validated hit compounds for
56
57 2
58
59
60 ACS Paragon Plus Environment
Page 3 of 60 Journal of Chemical Information and Modeling

1
2
3
4 further development of h-cGAS inhibitors but also demonstrates a cross-validation
5
6
7 study of virtual screening, in vitro experimental assays and crystal structures
8
9
10 determination.
11
12
13
14 1. INTRODUCTION
15
16
17 The innate immunity senses pathogen-related molecular patterns (PAMPs) and
18
19 damage-associated molecular patterns (DAMPs) through different pattern
20
21
22 recognition receptors (PRRs), and initiates innate immune response.1 Cyclic GMP-
23
24
25 AMP synthase (cGAS) has emerged as the dominant sensor of cytosolic double-
26
27
stranded DNAs (dsDNAs) in a sequence-independent manner,2–5 which are
28
29
30 canonical DAMPs derived from a variety of danger signals and tissue damage, such
31
32
33 as virus invasion, chromosome damage, mitochondrial damage, aging or other
34
35
36 abnormal conditions.6,7 Upon binding with cytoplasmic dsDNAs, cGAS promotes
37
38 cyclization of adenosine 5´-triphosphate (ATP) with guanosine 5´-triphosphate
39
40
41 (GTP) into cyclic [G(2´,5´)pA(3´,5´)p] (cGAMP), which acts as a second messenger
42
43
44 to activate an endoplasmic reticulum membrane-anchored adaptor protein, named
45
46
stimulator of interferon genes (STING).2, 8–11 Activated STING initiates a signaling
47
48
49 cascade by recruiting and activating the TANK-binding kinase 1 (TBK1) to
50
51
52 phosphorylate the transcription factor family of interferon regulatory factors 3
53
54
55
56
57 3
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 4 of 60

1
2
3
4 (IRF3), resulting in gene expression of type-I interferons (IFN-I) to trigger further
5
6
7 immune cascade reactions.8, 12–14
8
9
10 Owing to the high immunogenicity of cytoplasmic dsDNAs, it has been reported
11
12 that abnormal activation of the cGAS pathway is associated with various
13
14
15 autoimmune disorders. Functional loss of three-prime repair exonuclease 1
16
17
18 (TREX1) is a potential pathogenic factor for Aicardi–Goutières syndrome and
19
20
21
systemic lupus erythematosus (SLE), as TREX1 is the primary cellular exonuclease
22
23 in cytoplasms to prevent cells from excessive existence of dsDNAs and cGAS
24
25
26 activation.15–19 Polyarthritis/fetal and neonatal anemia were also identified as a type
27
28
29 I interferonopathy due to loss-of-function mutations in DNase II, another enzyme
30
31 digesting cytoplasmic DNA.20 Activation of cGAS by cytoplasmic escaping
32
33
34 mitochondrial DNAs (mtDNAs) was reported as an essential etiology of age-related
35
36
37 macular degeneration.21 Apart from these autoinflammatory and autoimmune
38
39
40
diseases, acute pancreatitis,22 myocardial infarction23 and other inflammatory
41
42 diseases were also associated with dysfunction of cGAS-STING signaling
43
44
45 pathways.24, 25 Notably, the deletion of cGAS in some autoimmune disease models
46
47
48 showed remissions of symptoms. Together, inhibition of cGAS has been regarded
49
50 as a promising therapeutic strategy for inflammatory and autoimmune diseases.24
51
52
53
54
55
56
57 4
58
59
60 ACS Paragon Plus Environment
Page 5 of 60 Journal of Chemical Information and Modeling

1
2
3
4 The promising therapeutic benefits of cGAS inhibition prompt a considerable
5
6
7 amount of interest in the design of cGAS inhibitors (Figure 1). Through a virtual
8
9
10 screening targeting the site of mouse cGAS (m-cGAS) interacting with DNA (PDB:
11
12 4LEZ), a series of antimalarial drugs (AMDs) were found to disturb the interactions
13
14
15 between cGAS and DNA, reduce the production of cGAMP and inhibit the dsDNA
16
17
18 sensing in THP-1 cells.26, 27 However, these AMDs have been proven to bind to
19
20
21
DNA rather than cGAS. Suramin, an essential drug for the treatment of river
22
23 blindness and African sleeping sickness, was disclosed as a cGAS inhibitor which
24
25
26 blocked the binding between cGAS and DNA.28 RU.521 and G150 were lately
27
28
29 developed on the basis of the hits resulted from high throughput screenings, and
30
31 both compounds occupied the cGAS active site.29,30 By far, RU.521 and G150 are
32
33
34 the most potent inhibitors of m-cGAS and human cGAS (h-cGAS), and their cellular
35
36
37 efficacies have been demonstrated in mouse RAW264.7 cells and human THP-1
38
39
40
cells, respectively. In addition, Hall et al. at Pfizer reported the first h-cGAS selective
41
42 inhibitor (PF-06928215) which possessed a high affinity with the catalytic site and
43
44
45 was identified by a fragment screening via NMR followed by a fragment evolution.
46
47
48 Nevertheless, PF-06928215 displayed no activity in cell-based cGAS assays and
49
50 the underlying mechanism remains unclear.31
51
52
53
54
55
56
57 5
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 6 of 60

1
2
3
4 Virtual screening has proven great advantage in identifying new chemical hits,32
5
6
7 in particular for the case of cGAS to which only a few inhibitors are available. We
8
9
10 thereby utilized the virtual screening campaigns together with in vitro assays and
11
12 crystal structure determination to discover novel inhibitors of h-cGAS. First, we
13
14
15 determined a high-resolution (1.8 Å) crystal structure of the catalytic domain of h-
16
17
18 cGAS (h-cGASCD) in complex with PF-06928215, which was the only validated
19
20
21
inhibitor of h-cGAS when we started this project. Then, a virtual screening campaign
22
23 was subsequently carried out based on this complex structure and several novel-
24
25
26 scaffold hits as h-cGAS inhibitors were successfully identified. Moreover, the crystal
27
28
29 structures of h-cGASCD in complex with these newly discovered hits were
30
31 determined, providing structural basis for further development of novel inhibitors.
32
33
34 As an example, a similarity search upon one of these hits (compound 18) resulted
35
36
37 in the discovery of new inhibitors with improved potency.
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57 6
58
59
60 ACS Paragon Plus Environment
Page 7 of 60 Journal of Chemical Information and Modeling

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25 Figure 1. Chemical structures of reported cGAS inhibitors.
26
27
28 2. METHODS
29
30
31 2.1 h-cGAS Expression and Purification. A gene encoding full-length human
32
33
34 cGAS (Fl-h-cGAS) (amino acids 1-522 of human MB21D1 (Uniprot ID: Q8N884))
35
36 was synthesized at synbio-tech and cloned into pET-15b for expression of Fl-h-
37
38
39 cGAS with an N-terminal 6×His-SUMO2 fusion protein, as described by Zhou, W.
40
41
42 et al. previously.33 Based on this plasmid, a construct for expression of the catalytic
43
44
domain of h-cGAS (amino acids: 157-522, h-cGASCD) was further generated by
45
46
47 molecular cloning. The sequences of two plasmids were confirmed by sequencing.
48
49
50 Rosetta 2 (DE3) (Sigma-Aldrich) was transformed with the above plasmids and
51
52
53 cultured in LB media at 37 °C to reach an OD600 of 0.8. Then the culture temperature
54
55
56
57 7
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 8 of 60

1
2
3
4 was reduced to 16 °C and isopropyl β-D-1-thiogalactopyranoside (IPTG) was added
5
6
7 with a final concentration of 0.5 mM for induction. After 8-10 h, bacteria were
8
9
10 harvested by centrifugation at 4 °C, washed with PBS, flash-frozen in liquid
11
12 nitrogen, and stored at -80 ℃ until lysis for purification.
13
14
15 The protein purification protocol followed the reported procedure33 with
16
17
18 modifications. Cell pellets resulted from 2 L LB culture were lysed in a lysis buffer
19
20
21
(20 mM HEPES, pH 7.5, 400 mM NaCl, 1 mM TCEP) with DNase I using a high-
22
23 pressure homogenizer. The solution was centrifuged at 12,000 g for 30 min, and
24
25
26 the supernatant was slowly passed through a manually packed nickel-bead column
27
28
29 (GE) which was pre-equilibrated with the lysis buffer. The column was washed with
30
31 a wash buffer (20 mM HEPES, pH 7.5, 400 mM NaCl, 50 mM imidazole, 1 mM
32
33
34 TCEP), and then eluted with an elution buffer (20 mM HEPES, pH 7.5, 400 mM
35
36
37 NaCl, 500 mM imidazole, 1 mM TCEP). The eluted fractions were pooled, diluted
38
39
40
with an equal volume buffer without salt (20 mM HEPES, pH 7.5, 1 mM TCEP),
41
42 supplemented with ~250 μg of human SENP2 protease (residues 364–589 with an
43
44
45 M497A mutation), and dialyzed overnight at 4 ℃. Untagged cGAS was further
46
47
48 purified by binding to two 5-mL Heparin HP ion-exchange columns connected in
49
50 tandem (GE Healthcare) and eluted with the buffer (20 mM HEPES, pH 7.5, 1 mM
51
52
53 TCEP) containing a gradient of 300-1000 mM NaCl. The eluted protein was
54
55
56
57 8
58
59
60 ACS Paragon Plus Environment
Page 9 of 60 Journal of Chemical Information and Modeling

1
2
3
4 concentrated and injected into a size-exclusion chromatography column (16/600
5
6
7 Superdex S75, GE Healthcare) equilibrated with the gel filtration buffer (20 mM
8
9
10 HEPES, pH 7.5, 250 mM KCl, 1 mM TCEP, 10% glycerine). The S75 pools from
11
12 each peak were separately collected, concentrated to 8-9 mg/ml, aliquoted and
13
14
15 flash frozen in liquid nitrogen. Fl-h-cGAS was used for the enzyme activity assay,
16
17
18 and h-cGASCD was used for crystallization as well as the thermal shift assay.
19
20
21
2.2 Thermal Shift Assay. Thermal shift assays were performed on a 7500 Fast
22
23 Real-Time PCR System (Thermo Fisher Scientific) with 96-well white plates
24
25
26 (Thermo Fisher Scientific). Each well contained 20 μL 5 μM h-cGASCD and 5×
27
28
29 Protein Thermal Shift dye (Thermo Fisher Scientific) in a buffer (20 mM HEPES, pH
30
31 7.5, 250 mM KCl, 1 mM TCEP, 10% glycerine). Compounds were tested at a final
32
33
34 concentration of 200 μM or 400 μM in 5% (v/v) DMSO. Each plate was sealed with
35
36
37 an optically clear foil and centrifuged for 1 min at 1000 rpm. Then the plates were
38
39
40
heated from 25-95 °C at approximately 1 °C min−1. The fluorescence intensity was
41
42 measured with λex = 480 nm and λem = 580 nm. The melting temperature (Tm) of the
43
44
45 protein with or without the compound was obtained by determining the minimum of
46
47
48 the first derivative curve of the melt curve. The thermal shift (ΔTm) was determined
49
50 by calculating the difference between the Tm of the protein in the presence of and
51
52
53 in the absence of the tested compound.
54
55
56
57 9
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 10 of 60

1
2
3
4 2.3 Pyrophosphatase-coupled cGAS Activity Assay. The enzymatic activity of
5
6
7 cGAS was measured using a pyrophosphatase-coupled assay developed by
8
9
10 Richard and Jungsan34 with some modifications. Briefly, a final concentration of 200
11
12 nM Fl-h-cGAS, 50 nM E.coli pyrophosphatase (Sigma-Aldrich I5907), 1 mM ATP,
13
14
15 0.3 mM GTP, 5 μg/ml HT-DNA (Sigma-Aldrich) in 40 μL reaction buffer (10 mM
16
17
18 HEPES, 140 mM NaCl, 0.01% Tween-20, 5 mM MgCl2, pH 7.5) in the presence or
19
20
21
absence of tested compounds were incubated for 90 min at room temperature.
22
23 Then, 40 μL quenched solutions (50 mM EDTA) were mixed with 20 μL malachite
24
25
26 green solution and incubated for 45 min at room temperature. The absorbance of
27
28
29 each well at ~620 nm was measured and compared to the negative control (without
30
31 HT-DNA) and the positive control (DMSO instead of the compound).
32
33
34 The amount of phosphoric acid represents the amount of pyrophosphoric acid
35
36
37 generated by the reaction, allowing for the determination of the equivalent amount
38
39
40
of cGAMP generated in the reaction. The inhibition ratio of a compound was
41
42 normalized against the positive control and the negative control as follows: inhibition
43
44
45 % = 1- (Abssample−Absaverage negative control)/(Absaverage positive control − Absaverage negative
46
47
48 control).
49
50 2.4 Crystallization and Structure Determination. Crystals of h-cGASCD were
51
52
53 obtained by a hanging-drop vapor diffusion method using 1.5 μL protein (8-9 mg/ml)
54
55
56
57 10
58
59
60 ACS Paragon Plus Environment
Page 11 of 60 Journal of Chemical Information and Modeling

1
2
3
4 plus 1.5 μL reservoir containing PEG 3350 (17.5-19% v/v) and 0.2 M ammonium
5
6
7 citrate (pH 7-7.5) at 4 °C. Crystals of complexes were obtained via soaking apo
8
9
10 crystals of h-cGASCD with compounds at a concentration of 10 mM in the mother
11
12 liquor overnight. Crystals were flash frozen in liquid nitrogen in the presence of the
13
14
15 mother liquor supplemented with 20% glycerol.
16
17
18 The diffraction data of crystals were collected at 100 K on the beamline BL17U1
19
20
21
or BL19U1 at the Shanghai Synchrotron Radiation Facility (SSRF) and were
22
23 processed with the HKL3000 software packages.35 The structures were solved by
24
25
26 molecular replacement, using the program CCP436 with a search model of PDB
27
28
29 code 4LEV.5 The structures were refined using PHENIX.37 With using the program
30
31 Coot,38 compounds and water molecules were fitted into the initial Fo-Fc map. Data
32
33
34 collection and refinement statistics of six determined structures are shown in Table
35
36
37 S1.
38
39
40
2.6 Molecular Dynamics (MD) Simulations. MD simulations were performed to
41
42 observe the key residues interacting with PF-06928215 by using AMBER 14
43
44
45 package,39 which provides a hint for picking compounds after the virtual screening.
46
47
48 Our crystal structure of h-cGASCD in complex with PF-06928215 and the protein
49
50 structure solely (PDB code: 6LRC) were used for setting up the model of MD
51
52
53 simulations. The parameters of PF-06928215 were calculated using the B3LYP
54
55
56
57 11
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 12 of 60

1
2
3
4 method at the 6-31G* level with Gaussian 09.40 Its restricted electrostatic potential
5
6
7 (RESP) was generated by Antechamber41 and assigned the general amber force
8
9
10 field.42 For the protein, we fixed the missing side chains and loops, deleted Zn2+
11
12 ions, and the ff14SB force field was used.43 To mimic the stable interactions with
13
14
15 the Zn2+ ion, four residues (H390, C396, C397, and C404) interacting with Zn2+ were
16
17
18 restricted in all the stages of MD simulations. Next, the protein or the complex were
19
20
21
solvated with an 8-Å TIP3P water box. The whole system was neutralized by adding
22
23 counterions (Cl- or Na+). Two successive energy minimizations were carried out
24
25
26 using the steepest descent and the conjugate gradient algorithm in the Sander
27
28
29 module. First, hydrogen atoms and water molecules were minimized in 2500 steps
30
31 using the steepest descent algorithm combined with the conjugate gradient one.
32
33
34 Second, another 5000 steps of the steepest-descent minimization followed by 5000
35
36
37 cycles of the conjugated gradient one were performed to minimize all atoms of the
38
39
40
systems except a restriction on four residues that coordinate with the Zn2+ ion.
41
42 Then, the system was gradually heated from 0 to 100 K and 100 to 300 K separately
43
44
45 in the NVT ensemble over 5 ps and 100 ps, respectively, with a smaller restriction
46
47
48 on protein and ligand. Then, the system was equilibrated at 300 K and 1 atm without
49
50 any restriction except for the four residues mentioned above. The periodic boundary
51
52
53 conditions were utilized. The SHAKE algorithm44 was applied to constrain the
54
55
56
57 12
58
59
60 ACS Paragon Plus Environment
Page 13 of 60 Journal of Chemical Information and Modeling

1
2
3
4 hydrogen atoms. The long-range electrostatic interactions were calculated by the
5
6
7 particle mesh Ewald (PME) method.45 Finally, a 100-ns productive MD trajectory
8
9
10 was generated for each model at a constant temperature and pressure using the
11
12 PMEMD module46 implemented in the AMBER 14. The coordinates of the protein
13
14
15 or the protein-ligand complex in the MD trajectory were saved every 10 ps.
16
17
18 2.7 Molecular Mechanics Poisson-Boltzmann Surface Area (MM/PBSA)
19
20
21
Calculations. The energy decomposition was calculated based on the 100-ns MD
22
23 trajectories in order to assess the contribution of each residue to the total binding
24
25
26 free energy of PF-06928215 with h-cGASCD. The binding free energy and energy
27
28
29 decomposition calculations were implemented using the MM/PBSA method47 in
30
31 AMBER 14. The binding free energy calculations were performed on 20 snapshots
32
33
34 extracted from the 100 ns trajectories. In MM-PBSA, the binding free energies
35
36
37 (ΔGbind) are calculated by the following equations:
38
39
40
ΔGbind = Gcomplex – (Greceptor + Gligand) (1)
41
42 ΔGbind = ΔEMM + ΔGsolvation – TΔS (2)
43
44
45 where Gcomplex, Greceptor and Gligand are the representations of the free energies of
46
47
48 complex, receptor and ligand, respectively; ΔEMM is the gas phase molecular
49
50 mechanical energy; ΔGsolvation is the desolvation free energy; -TΔS represents the
51
52
53 conformational entropy upon binding of ligand at a temperature of T. The binding
54
55
56
57 13
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 14 of 60

1
2
3
4 free energies (ΔGbind) are calculated using equation 1 and each free energy is
5
6
7 calculated as the sum of the MM energy EMM, the solvation free energy Gsolvation and
8
9
10 the entropy contribution S, respectively (equation 2). The MM energy is
11
12 decomposed into EMM, EMM, and EMM,
complex, rec lig. The solvation free energy
13
14
15 represents the sum of the electrostatic solvation free energy and nonpolar solvation
16
17
18 free energy. The entropy contribution was neglected due to the cost of the
19
20
21
computational process is huge for large protein-ligand systems.
22
23 2.8 Virtual Screening. Our crystal structure of h-cGASCD in complex with PF-
24
25
26 06928215 (PDB code: 6LRC) was used as the receptor structure. Three water
27
28
29 molecules forming more than 3 hydrogen bonds (H-bonds) to ligand or/and residues
30
31 were reserved. The docking grid was centered on the centroid of PF-06928215.
32
33
34 The receptor was then prepared using the Protein Preparation Wizard and Grid
35
36
37 Preparation tools. The OPLS3 force field was used for minimization and grid
38
39
40
generation.48
41
42 We used tools implemented in the Schrödinger suite to screen the ChemDiv
43
44
45 database. We first prepared the 3D structures of ~1,500,000 compounds from the
46
47
48 database with LigPrep.49 The output structures were then supplied for the virtual
49
50 screening workflow. The overall workflow includes filtering the compounds with
51
52
53 Propfilter on QikProp properties and docking them to the receptor with Glide.50 In
54
55
56
57 14
58
59
60 ACS Paragon Plus Environment
Page 15 of 60 Journal of Chemical Information and Modeling

1
2
3
4 the first step, we filtered out compounds that are not compatible with lead-like rules
5
6
7 (250<MW≤350, LogP≤3.5) or Lipinski’s Rule of 5. In the docking step, we carried
8
9
10 out Standard Precision (SP) calculations with default settings. The top 10% of the
11
12 best compounds were reserved for the post-processing analysis. The OPLS3 force
13
14
15 field was used for docking. The survivors were then imported into Canvas for
16
17
18 clustering analysis using the Leader−Follower method.51 Ultimately, the top-ranking
19
20
21
compounds were visually analyzed and 59 compounds with diverse structures were
22
23 purchased for biological testing.
24
25
26 2.9 Similarity Searching. In the similarity searching of compound 18, we used
27
28
29 tools implemented in the Schrödinger suite to screen the ChemDiv database. We
30
31 first prepared the 3D structures of ~1,500,000 compounds from the database with
32
33
34 LigPrep. The workflow of ligand-based virtual screening includes filtering the
35
36
37 compounds with the Shape Screening and docking them to the receptor with
38
39
40
Glide.52 First, the compounds were filtered out with a similarity score below 0.6
41
42 using the structure of compound 18 as a shape query. Second, we carried out
43
44
45 Standard Precision (SP) calculations with default settings to dock the remaining
46
47
48 compounds to the receptor. The top 10% of the best compounds were reserved for
49
50 the post-processing analysis. The OPLS3 force field was used for docking. The
51
52
53 survivors were then imported into Canvas for clustering analysis using the
54
55
56
57 15
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 16 of 60

1
2
3
4 Leader−Follower method. Ultimately, 7 compounds with different structures were
5
6
7 purchased for biological testing.
8
9
10
11
12 3. RESULTS AND DISCUSSION
13
14
15 3.1 A High Resolution Crystal Structure of h-cGASCD in Complex with PF-
16
17
18 06928215. In order to set up the structural biological platform for drug design
19
20
21
targeting cGAS, we first tried to optimize the crystallization conditions of apo h-
22
23 cGASCD to obtain the crystals for inhibitors soaking. And we determined the crystal
24
25
26 structure of h-cGASCD bound with PF-06928215 at a resolution of 1.8 Å (PDB code:
27
28
29 6LRC), which is the highest resolution one among the reported crystal structures of
30
31 h-cGASCD (Figure 2 and Table S1). Similar to the crystal structure of h-cGAS-PF-
32
33
34 06928215 (PDB code: 6NAO) determined previously at a resolution of 3.2 Å,31 the
35
36
37 pyrazolopyrimidine of PF-06928215 is sandwiched between the guanidinium group
38
39
40
of R376 and the aromatic ring of Y436, mimicking the nucleobase. The benzene is
41
42 anchored in a hydrophobic sub-pocket lined by the side chains of F488 and L490.
43
44
45 Nevertheless, we observe that the position of PF-06928215 in our structure is
46
47
48 slightly shifted compared to the one in 6NAO (RMSD =0.73 Å) so that the
49
50 carboxylate headpiece not only interacts with R376 and K36231 but also makes
51
52
53 contact with R302 through two water molecules (Figure 2c). The carbonyl of the
54
55
56
57 16
58
59
60 ACS Paragon Plus Environment
Page 17 of 60 Journal of Chemical Information and Modeling

1
2
3
4 amide forms an H-bond with the hydroxyl of S434 (Figure 2b), which is not found in
5
6
7 the previous structure owing to a larger distance of the acceptor to the donor.
8
9
10 Additionally, owing to the high resolution, more water molecules have been seen in
11
12 our structure and they mediate multiple interactions between PF-06928215 and
13
14
15 residues such as R302, L377, S378, F379, S380, E383, S435, Y436, and N482
16
17
18 (Figure 2c). Together, the crystal structure of h-cGASCD/PF-06928215 determined
19
20
21
at 1.8 Å reveals more distinct protein-ligand interactions and provides a good
22
23 starting point for the design of new inhibitors.
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57 17
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 18 of 60

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38 Figure 2. Crystal structures of h-cGASCD in complex with PF-06928215. (a) The
39
40
41 2Fo–Fc electron density map of the bound PF-06928215 contoured at 1.6 σ level.
42
43
44 (b) The structure of PF-06928215 in the high-resolution crystal structure (PDB code:
45
46 6LRC, yellow) superimposed to that in 6NAO (green). (c) Intermolecular contacts
47
48
49 among PF-06928215, water molecules and amino acids lining the binding pocket
50
51
52 of h-cGASCD.
53
54
55
56
57 18
58
59
60 ACS Paragon Plus Environment
Page 19 of 60 Journal of Chemical Information and Modeling

1
2
3
4 3.2 MD Simulations of h-cGASCD in Complex with PF-06928215. In order to
5
6
7 better understand the binding mode of PF-06928215 with h-cGAS, the high-
8
9
10 resolution crystal structure (PDB code: 6LRC) mentioned above was subjected to
11
12 MD simulations. The C atom RMSD and RMSF of each residues along the 100-
13
14
15 ns MD trajectories show a relative stable structure of h-cGASCD during MD
16
17
18 simulations (Figure S1). The 100-ns MD trajectories were then used to calculate
19
20
21
the binding free energy and energy decomposition utilizing the MM/PBSA method.
22
23 While the entropic term is ignored because of its huge cost of computation, the
24
25
26 calculated binding energy indicates that the binding enthalpy can be used as an
27
28
29 good indicator of the ligand-receptor binding and hit selection.53 The results from
30
31 the MM/PBSA calculations show that the contribution of electrostatic interactions
32
33
34 (EELE= -11.16 kcal/mol) is much smaller compared to the van der Waals interactions
35
36
37 (EVDM= -41.43 kcal/mol), a major contribution to the binding of PF-06928215 with h-
38
39
40
cGAS (Table S2). In accordance with this, the nonpolar interactions (EVDM +
41
42 GNPOLAR) are more significant compared with the polar interaction (EELE + GPB),
43
44
45 indicating that compounds forming more nonpolar interactions with h-cGAS in the
46
47
48 following virtual screening could be selected for biological evaluation.
49
50 To find out which residues count a great deal in the receptor-ligand binding,
51
52
53 energy decomposition calculations were performed. As presented in Figure 3, PF-
54
55
56
57 19
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 20 of 60

1
2
3
4 06928215 forms the most robust Van der Waal interactions with Y436 as well as
5
6
7 R376 and great electrostatic interactions with R376 in the active site. It also shows
8
9
10 that E383 establishes distinct electrostatic interactions with PF-06928215 but with
11
12 a high deviation. However, in the crystal structure, it makes contact with PF-
13
14
15 06928215 through two water molecules (Figure 2c), although the representative
16
17
18 snapshots extracted from the MD trajectory display that the pyrazolopyrimidine
19
20
21
interacts directly with E383 without the aid of water molecules (Figure S2).
22
23 Considering the inconsistency between the crystal structure and the calculation
24
25
26 together with the high deviation, the electrostatic interactions of compounds with
27
28
29 E383 are not taken as a key factor for hit selection. In addition, most of the residues
30
31 in the active-site pocket tend to form nonpolar interactions with the ligand, such as
32
33
34 K362, L377, S378, F379, S434, H437, N482, F488, L490, and L495. Accordingly,
35
36
37 on the basis of energy decomposition calculations on the h-cGASCD-PF-06928215
38
39
40
complex, the compounds establishing more nonpolar or polar interactions with
41
42 residues indicated in Figure 3b would be favored in hit selection.
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57 20
58
59
60 ACS Paragon Plus Environment
Page 21 of 60 Journal of Chemical Information and Modeling

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18 Figure 3. Energy decomposition calculations on the h-cGASCD-PF-06928215
19
20
21
complex. (a) The contributions of van der Waals and electrostatic interactions of
22
23 residues (171-510) to the binding of PF-06928215 with h-cGASCD. (b) Contributions
24
25
26 of these residues positioned less than 5 Å of PF-06928215.
27
28
29
30
3.3 Virtual Screening. The ChemDiv database containing ~1,500,000 compounds
31
32 was applied in our virtual screening with an aim of identifying novel scaffolds of
33
34
35 cGAS inhibitors. As mentioned above, the crystal structure of h-cGASCD in complex
36
37
38 with PF-06928215 (PDB codes: 6LRC and 6NAO) determined at 1.8 and 3.2 Å,
39
40 respectively, were tested for molecular docking. The RMSD of docked conformation
41
42
43 of PF-06928215 compared to it in 6LRC and 6NAO is 1.53 and 1.71 Å (Figure S3),
44
45
46 respectively, suggesting no significant difference was seen between the docking of
47
48
49
two crystal structures to regenerate the binding pose of the ligand. Only our crystal
50
51 structure (PDB code: 6LRC) was thus used for the following virtual screening. In
52
53
54 addition, in the docking model, three water molecules were kept as they established
55
56
57 21
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 22 of 60

1
2
3
4 more than 3 H-bonds to ligand or/and residues (Figure S3a). The RMSD of PF-06928215
5
6
7 between the docked conformation and its crystal structure without water molecules
8
9
10 is 5.16 Å, much larger than the one (1.53 Å) with three water molecules (Figure
11
12 S3a, b). The workflow of the virtual screening and the protocol of molecular docking
13
14
15 are described in Material and Methods. Compared with the docking scores of
16
17
18 ligands from the crystal structures, the compounds with docking scores greater than
19
20
21
-6.4 were retained, resulting in 3000 compounds. To ensure the structural diversity
22
23 of compounds, a clustering analysis based on the Leader−Follower method was
24
25
26 applied to these 3000 compounds so as to decrease the number of compounds with
27
28
29 similar scaffolds. According to the binding free energy and energy decomposition,
30
31 the compounds forming the Van der Waals interactions with Y436 and R376, and
32
33
34 the electrostatic interactions with R376 were retained. In addition, those featuring
35
36
37 more nonpolar interaction with more negative-energy residues listed in Figure 3b
38
39
40
were also selected. In the end, 59 compounds were purchased for biological testing.
41
42 3.4 Evaluation of Virtual Screening Hits via the Thermal Shift Assay (TSA). When
43
44
45 gradually heated, the protein unfolds and its hydrophobic core is exposed, providing
46
47
48 more hydrophobic regions for fluorescence probes binding and resulting in a rise of
49
50 fluorescence intensity. The binding of a ligand to the protein will increase or reduce
51
52
53 the melting temperature (Tm) by stabilizing or destabilizing the protein, respectively.
54
55
56
57 22
58
59
60 ACS Paragon Plus Environment
Page 23 of 60 Journal of Chemical Information and Modeling

1
2
3
4 The difference between the Tm of the protein binding with and without the ligand
5
6
7 indicates the thermal shift (ΔTm).54 The value of ΔTm indicates the binding affinity of
8
9
10 compounds with a targeting protein and such an assay is widely used to screening
11
12 compounds or verify the binding of compounds with a protein.55
13
14
15 We first explored if the TSA was feasible for evaluation of the binding of
16
17
18 compounds with h-cGASCD as such an assay had not been reported for cGAS
19
20
21
inhibitors screening or binding validation. During the process of protein purification
22
23 in our study as well as others, two peaks of h-cGASCD were eluted from the size-
24
25
26 exclusion chromatography, suggesting that h-cGASCD can form dimer without
27
28
29 dsDNA.34 Therefore, we measured the melting temperatures of the purified
30
31 monomer as well as dimer, and the Tm of the dimer is nearly 3 ℃ higher than that
32
33
34 of the monomer (Figure 4a, b). In addition, the high-resolution crystal structure of
35
36
37 the h-cGASCD-PF-06928215 complex was solved by soaking the inhibitor to the
38
39
40
crystals of apo h-cGASCD dimer. Accordingly, the dimeric h-cGASCD was used for
41
42 the following TSA.
43
44
45 Adding of 50 μM PF-06928215 to the dimeric h-cGASCD protein solution results
46
47
48 in 7℃ change in transition temperature (∆Tm) (Figure 4c, d), which means the
49
50 binding of ligands such as PF-06928215 to the active site of h-cGASCD affects the
51
52
53 thermal stability of h-cGASCD. Therefore, we applied the TSA to test the binding of
54
55
56
57 23
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 24 of 60

1
2
3
4 59 compounds resulted from the virtual screening with h-cGASCD. As a result, 4
5
6
7 compounds (3, 17, 23, and 40) at a concentration of 400 μM shifted the melting
8
9
10 temperature of h-cGASCD (ΔTm > 1 °C, Table 1, Figure S4). Interestingly, these four
11
12 compounds all belong to N-tricyclic heterocyclic carboxylic acids and two carboxyl
13
14
15 groups present in compounds 3 and 17 with high ΔTm values (2.50 °C and 2.90 °C
16
17
18 for 3 and 17, respectively).
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 Figure 4. The melting temperature of h-cGASCD with or without compounds
35
36
37 measured by the thermal shift assay. (a, b) The melting curve of monomer (gray)
38
39
40
and dimer (blue) of h-cGASCD. (c, d) The melting curve of the dimeric h-cGASCD in
41
42 the absence (blue) and presence of PF-06928215(orange).
43
44
45
46 Table 1. The melting temperature of h-cGASCD shifted upon the binding of
47
48
49
compounds
50
51
52 Compound Structure MW. ∆Tm (℃) Inhibition radio
53
54
55
56
57 24
58
59
60 ACS Paragon Plus Environment
Page 25 of 60 Journal of Chemical Information and Modeling

1
2
3
4 s (100 μM)
5
6 H
7 O N O

8
3 285 2.50 39.9 %
9
10 HO OO OH

11 O
12 OH

13 N

17 316 2.90 29.9 %


N O
14 N
N

15 HO

16 O

17
18 N
HO

23 314 1.77 21.8 %


H
19 N N N
H
N O
O
20 N

21
22
N
23 O
40 H
295 1.14 16.8 %
24 N N OH

25
26
27 3.5 Crystal Structures of h-cGASCD in Complex with Four Hits. To reveal the
28
29
30 binding features of the four positive hits from the TSA, crystal structures of h-
31
32
33 cGASCD in complex with these compounds were determined (Figures 5, S5 and S6).
34
35 Remarkably, similar to the docking predicted conformations, the heterocyclics of
36
37
38 the four hit compounds are sandwiched between R376 and Y436 (Figure 5), and it
39
40
41 is consistent with the previous finding that this site could accommodate a variety of
42
43
44
2- or 3-ring aromatic systems.56 The isoquinoline-1, 3(2H)-dione head of compound
45
46 3 (PDB code: 6LRE) interacts with residues E383 and S378 through water
47
48
49 molecules, which resembles the interactions of the pyrazolopyrimidine or the
50
51
52 hydroxyl group of PF-06928215 with surrounding residues. While no π-π interaction
53
54 is formed between 3 and F488, the binding of 17, 23 and 40 (PDB code: 6LRI, 6LRJ
55
56
57 25
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 26 of 60

1
2
3
4 and 6LRK, respectively) all result in a side-chain flip of F488 and the π-π
5
6
7 interactions of these compounds with F448 are thus established. In addition, the
8
9
10 carboxyl groups in four compounds form H-bond (s) or/and salt bridge with
11
12 surrounding residues, although their location and orientation are different. For
13
14
15 example, both carboxyl groups of compound 3 form an H-bond to S434, while the
16
17
18 carboxyl group of 17 either contacts with S378 through a direct H-bond or with H437
19
20
21
via a water molecule. A salt bridge is established between the carboxyl group of 23
22
23 and R376, while the carboxyl group of 40 sitting on the chair-like configuration of
24
25
26 the piperidine constructs both the salt bridge and water-mediated H-bonds to two
27
28
29 positive-charged residues R376 and K362. For compounds 17 and 23, their
30
31 carboxyl groups are linked to the N-tricyclic heterocyclic ring through a relatively
32
33
34 long and flexible chain, leading to the partial missing of electron density for the chain
35
36
37 (Figure S5 and S6). Overall, the sandwich interactions of the N-tricyclic heterocyclic
38
39
40
ring with R376 and Y436 act as an anchor for the binding of the four compounds,
41
42 while the accommodation of the diverse location or orientation of these carboxyl
43
44
45 groups by the ligand binding pocket of h-cGAS reveals that this pocket possesses
46
47
48 a relatively open and polar character. Moreover, the docked conformation of the
49
50 ligand is well superimposed to it in the crystal structure for all four cases, supporting
51
52
53 that our docking model as well as the workflow of the virtual screening is reliable to
54
55
56
57 26
58
59
60 ACS Paragon Plus Environment
Page 27 of 60 Journal of Chemical Information and Modeling

1
2
3
4 discover novel h-cGAS inhibitors. These crystal structures also suggest that the
5
6
7 TSA is a reliable assay for the screening of h-cGAS inhibitors, although the
8
9
10 determined ∆Tm values are relatively low. Moreover, we also find the water molecule
11
12 next to R376 can be seen in three crystal structures of complexes except that of
13
14
15 compound 17, although three water molecules including this one were conserved
16
17
18 in the docking model (Figure S7). It is thus suggested that the water molecule next
19
20
21
to R376 may play a pivotal role in h-cGAS-inhibitor binding, providing a better model
22
23 for future virtual screening.
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57 27
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 28 of 60

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38 Figure 5. Superposition of the docked conformation (gray) and bound conformation
39
40
41 of compounds 3 (a), 17 (b), 23 (c), and 40 (d) (PDB codes: 6LRE, 6LRI, 6LRJ, and
42
43
44 6LRK, respectively) in their crystal structures with h-cGASCD. The RMSD of the
45
46 ligand between the docked conformation and the crystal structure is 1.50 Å, 1.26 Å,
47
48
49 0.48 Å, and 1.30 Å, for 3, 17, 23, and 40, respectively.
50
51
52
53
54
55
56
57 28
58
59
60 ACS Paragon Plus Environment
Page 29 of 60 Journal of Chemical Information and Modeling

1
2
3
4 3.6 Evaluation of Virtual Screening Hits via the PPiase-coupled Assay. To
5
6
7 measure inhibitory activities of the 59 compounds against h-cGAS, we applied a
8
9
10 pyrophosphatase (PPiase)-coupled assay34 with some modifications. Briefly, cGAS
11
12 links ATP and GTP to generate cGAMP with two inorganic pyrophosphates (PPi)
13
14
15 which are digested by a pyrophosphatase to generate phosphoric acid. The activity
16
17
18 of cGAS is reflected by the amount of phosphoric acid in the reaction system. With
19
20
21
this assay, we first tested the activity of the reported inhibitors PF-06928215 and
22
23 RU.521 against the full-length human cGAS (FL-h-cGAS). PF-06928215 inhibited
24
25
26 the catalytic activity of FL-h-cGAS with an IC50 of 2.08 ± 0.32 μM, which is close to
27
28
29 the previously reported value (2.0 μM) using a mass spectrometry assay (Figure
30
31 S8a).31 RU.521 inhibited the activity of the enzyme with an IC50 of 5.72 ± 0.32 μM
32
33
34 (Figure S8b), which is also consistent with the data reported lately.30 Subsequently,
35
36
37 the inhibitory activities of the 59 compounds obtained from the virtual screening
38
39
40
were tested at a concentration of 50 μM (poor solubility) or 100 μM, and the results
41
42 are shown in Figure 6a and Table S3. Unexpectedly, although compounds 3, 17,
43
44
45 23, and 40 show a positive result in the TSA and their crystal structures with h-
46
47
48 cGASCD were determined, they do not show an apparent inhibition on FL-h-cGAS
49
50 in this enzymatic activity assay. Instead, compound 18 (Figure 6b), with an IC50 of
51
52
53 29.88 ± 3.20 μM, was identified as the best inhibitor among the 59 compounds. We
54
55
56
57 29
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 30 of 60

1
2
3
4 tried to solve the crystal structure of h-cGASCD in complex with 18 but failed, mainly
5
6
7 due to the low solubility of the compound.
8
9
10 The inconsistent results from two assays may be ascribed to two factors. First,
11
12 different protein samples of h-cGAS were used in two assays, since h-cGASCD and
13
14
15 FL-h-cGAS were used for the TSA and the PPiase-coupled assay, respectively.
16
17
18 Second, the use of HT-DNA in the PPiase-coupled assay might contribute to this
19
20
21
inconsistency. We tried to apply the FL-h-cGAS for the TSA but resulted in an
22
23 uncharacteristic melting curve of the protein, while the curve of h-cGASCD shown in
24
25
26 Figure 4 suggests that h-cGASCD is more suitable than FL-h-cGAS to detect the
27
28
29 binding of compounds with h-cGAS. However, it has been reported that because of
30
31 the cGAS liquid phase separation the catalytic activity of h-cGASCD is weaker than
32
33
34 that of FL-h-cGAS. To be closer to the natural conditions, FL-h-cGAS is more
35
36
37 suitable for the h-cGAS activity assay.57 Nevertheless, the binding of four positive
38
39
40
hits (compounds 3, 17, 23, and 40) with h-cGASCD has been verified by the crystal
41
42 structure determination. Probably, the binding affinity of these four compounds with
43
44
45 h-cGAS is too weak to exert a significant inhibition on the enzyme’s activity. In
46
47
48 addition, because of the low solubility of compound 18, the highest concentration
49
50 used in the TSA is 200 μM and the resulting ΔTm is 0.9 °C.
51
52
53
54
55
56
57 30
58
59
60 ACS Paragon Plus Environment
Page 31 of 60 Journal of Chemical Information and Modeling

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33 Figure 6. Inhibition of the FL-h-cGAS by 59 compounds measured by the PPiase-
34
35
coupled assay. (a) Inhibition of the FL-h-cGAS by the 59 compounds at a
36
37
38 concentration of 100 μM or 50 μM (for compounds with a low solubility). (b)
39
40
41 Chemical structure and dose-dependent inhibition of compound 18 against the FL-
42
43
44 h-cGAS.
45
46
47
48
49 3.7 Similarity Search of Compound 18. A similarity searching of compound 18
50
51
52 was carried out in order to improve the solubility as well as the potency of the
53
54
55 inhibitor. To this end, we first used the structure of 18 as a shape query to search
56
57 31
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 32 of 60

1
2
3
4 new compounds with a similarity score above 0.6, and 1400 compounds from the
5
6
7 ChemDiv database were retained. Given the reliability of our virtual screening
8
9
10 described above, the molecular docking was then applied as the second filtration.
11
12 Afterward, 607 compounds with a docking score above -5.0 were kept while the
13
14
15 docking score of compound 18 was -6.95. These survivors were then imported into
16
17
18 Canvas for clustering analysis using the Leader−Follower method. Finally, 24
19
20
21
compounds passed through a variety of filtering and 21 commercially available ones
22
23 were purchased and tested (Table S4).
24
25
26 We applied the PPiase-coupled assay on these 21 compounds at a concentration
27
28
29 of 62.5 μM (Figure 7a). Two compounds were identified to be more potent than
30
31 compound 18. Compound S2, with a carboxyl group attached to benzene in 18
32
33
34 changed from the ortho to meta position, has IC50 of 13.1 ± 0.09 μM (Figure 7b). In
35
36
37 addition, the IC50 value of compound S3 is 4.9 ± 0.26 μM (Figure 7b), showing a 6-
38
39
40
fold increase in potency compared to compound 18 and a potency similar to that of
41
42 PF-06928215 and RU.521. The similarity scores (Table S5) based on Modified
43
44
45 Tanimoto Distance measure between three novel compounds (compound S2, S3
46
47
48 and 18) identified and two existing ones (PF-06928215 and RU.521) showed that
49
50 there are several varied structural features. These relatively rigid compounds can
51
52
53 be divided into three parts: the aromatic, amide and ring region. In the first part, 2H-
54
55
56
57 32
58
59
60 ACS Paragon Plus Environment
Page 33 of 60 Journal of Chemical Information and Modeling

1
2
3
4 3,4'-bi(1,2,4-triazole) of three compounds replaced 7-hydroxy-5-
5
6
7 phenylpyrazolo[1,5-a]pyrimidine and 4,5-dichloro-1H-benzo[d]imidazole in PF-
8
9
10 06928215 and RU.521, respectively. Regarding RU.521, the difference is slightly
11
12 larger as its 3-methyl-1H-pyrazol-5-ol group plays the same role as the amide does
13
14
15 in other compounds. Especially for the third part, the aromatic benzoic acid in
16
17
18 compound 18 and S2 is distinguished from the aliphatic carboxylic acid in PF-
19
20
21
06928215. Moreover, 4,5,6,7-tetrahydrobenzo[b]thiophene-3-carbonitrile of
22
23 compound S3 shows the largest variety in comparison to RU.521 and PF-06928215.
24
25
26 Together, these results demonstrate the high efficiency of the similarity searching
27
28
29 for the discovery of new and more potent inhibitors of h-cGAS.
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57 33
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 34 of 60

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 Figure 7. Inhibition of the FL-h-cGAS by 21 compounds measured by the PPiase-
35
36
37 coupled assay. (a) Inhibition of the FL-h-cGAS by 21 compounds (at 100 μM)
38
39
40
resulted from the similarity searching of compound 18. (b) Chemical structure and
41
42 dose-dependent inhibition of compounds S2 and S3 against the FL-h-cGAS.
43
44
45
46
47
48 3.8 Binding Modes of Compounds 18, S2 and S3 with h-cGASCD Characterized
49
50
51
by Crystal Structure together with Molecular Docking. After the discovery of
52
53 compounds S2 and S3, we tried again to determine the crystal structures of h-
54
55
56
57 34
58
59
60 ACS Paragon Plus Environment
Page 35 of 60 Journal of Chemical Information and Modeling

1
2
3
4 cGASCD in complex with these two compounds. We failed with compound S3 but
5
6
7 solved the S2 bound structure of h-cGASCD (PDB code: 6LRL). Figure 8a shows the
8
9
10 electron density of the bi(1,2,4-triazole) region of S2 is visible, while the meta-
11
12 benzoic acid moiety is blurred. On the basis of this crystal structure, molecular
13
14
15 docking of compounds 18, S2 and S3 to h-cGAS were performed again to reveal
16
17
18 the binding mode of these compounds (Figure 8b-e). It should be noted that the
19
20
21
electron density of compound S2 in the crystal structure especially the carboxyl is
22
23 not clear, leading the uncertainty in placement of compound S2 in the crystal
24
25
26 structure. Thus, the RMSD of S2 between the crystal structure and the predicted
27
28
29 docking conformation is relatively large (3.02 Å), but it could be 1.57 Å if the
30
31 carboxyl is omitted. The results indicated that three inhibitors with a similar scaffold
32
33
34 formed H-bonds with R376 and N482, a water-bridge H-bond with S380, π-π
35
36
37 interactions with Y436, and cation-π interactions with R376, providing the structural
38
39
40
basis underlying their potency to h-cGAS (Figure 8). Among these three inhibitors,
41
42 compound 18 is less potent than S2 and S3, mainly due to the high possibility of
43
44
45 forming an intra-molecular H-bond between two adjacent groups of the benzene,
46
47
48 the carboxyl and amide groups. Such an intra-molecular H-bond may also account
49
50 for the low solubility of 18. As a comparison, compound S2 with a very similar
51
52
53 structure to 18 possesses a better solubility and potency because such an intra-
54
55
56
57 35
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 36 of 60

1
2
3
4 molecular H-bond can not be established in S2. Nevertheless, the binding of the
5
6
7 most potent S3 slightly moves towards the interior of the binding pocket. The other
8
9
10 change is the hydrophobic interactions formed between the 4,5,6,7-
11
12 tetrahydrobenzo[b]thiophene of S3 and the hydrophobic sub-pocket consisting of
13
14
15 H437, F488, L490, and L495 (Figure 8c). These findings also verify the key
16
17
18 contribution of the nonpolar interactions to the cGAS-inhibitor binding predicted by
19
20
21
MD simulations.
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48 Figure 8. The binding modes of three inhibitors (compounds 18, S2 and S3) with h-
49
50 cGAS revealed by a crystal structure together with molecular docking. (a) The 2Fo–
51
52
53 Fc electron density map of the bound S2 contoured at 1.6 σ level (PDB code: 6LRL).
54
55
56
57 36
58
59
60 ACS Paragon Plus Environment
Page 37 of 60 Journal of Chemical Information and Modeling

1
2
3
4 (b-d) Superimposition of docked conformations of three inhibitors, compounds 18
5
6
7 (c), S2 (d) and S3 (e).
8
9
10
11
12
13 4. CONCLUSION
14
15 Based on the high-resolution crystal structure (1.8 Å, PDB code: 6LRC), the
16
17
18 binding model and the key residues of h-cGAS interacting with PF-06928215 were
19
20
21 studied using MD simulations. On this basis, virtual screening was carried out to
22
23 discover novel scaffolds of h-cGAS inhibitors, and 59 compounds were purchased.
24
25
26 Then we performed the TSA on these compounds and found 4 hits that were further
27
28
29 evaluated via the crystal structure determination. The bound poses of four hits in
30
31
32
the crystal structures are highly consistent with the docking poses. Subsequently,
33
34 a PPiase-coupled assay was used to evaluate the inhibitory activities of the 59
35
36
37 compounds, and compound 18 was found to have a good inhibitory activity. After
38
39
40 that, the similarity search on compound 18 brought compounds S2 and S3 with
41
42 improved inhibitory activities. Compound S3 shows the highest potency among the
43
44
45 compounds resulted from the present study (IC50: 4.9 ± 0.26 μM). Finally, we
46
47
48 determined the crystal structure of compound S2 bound to h-cGASCD and explored
49
50
51
the binding model of this series of compounds by molecular docking.
52
53
54
55
56
57 37
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 38 of 60

1
2
3
4 To sum up, although the open and polar property of the catalytic pocket of h-
5
6
7 cGAS poses a great challenge for virtual screening, several novel-scaffold hits of
8
9
10 h-cGAS inhibitors were successfully obtained through virtual screening. Meanwhile,
11
12 the crystal structures of h-cGASCD bound with these ligands indicate various binding
13
14
15 models of inhibitors with cGAS, providing important hints for further development of
16
17
18 cGAS inhibitors. In addition, our molecular docking and experimental data support
19
20
21
each other, presenting a good example to the discovery of novel ligands of a target
22
23 by such a cross-validation approach.
24
25
26
27
28
29
30
31 ASSOCIATED CONTENT
32
33
Supporting Information
34
35
36 The Supporting Information is available free of charge on the ACS Publications
37
38
39 website at DOI:
40
41
42
Structures of the compounds, experimental assays, molecular docking, and
43
44 MD simulations results.
45
46
47 AUTHOR INFORMATION
48
49
50 Corresponding Authors
51
52
*E-mail: ycxu@simm.ac.cn (Y. X).
53
54
55
56
57 38
58
59
60 ACS Paragon Plus Environment
Page 39 of 60 Journal of Chemical Information and Modeling

1
2
3
4 *E-mail: hongbinsun@cpu.edu.cn (H. S).
5
6
7
8 ORCID
9
10
Yechun Xu: 0000-0002-1581-6155
11
12
13
14 Hongbin Sun: 0000-0002-3452-7674
15
16
17
18 Author Contributions
19
20 #These authors contributed equally to this work.
21
22
23 Notes
24
25
26 The authors declare no competing financial interest.
27
28
29
ACKNOWLEDGMENTS
30
31 We acknowledge the financial support by the National Key R&D Program of China
32
33
34 (No. 2017YFB0202604) and the National Natural Science Foundation of China
35
36
37 (21877122).
38
39 ABBREVIATIONS
40
41
42 h-cGAS, human Cyclic GMP-AMP synthase; h-cGASCD, the catalytic domain of
43
44
45 human cGAS; HEPES, 2-[4-(2-Hydroxyethyl)-1-piperazinyl]ethanesulfonic acid;
46
47
MD, molecular dynamics; MM/PBSA, Molecular Mechanics Poisson-Boltzmann
48
49
50 Surface Area.
51
52
53 REFERENCES
54
55
56
57 39
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 40 of 60

1
2
3
4 (1) Tang, D.; Kang, R.; Coyne, C. B.; Zeh, H. J.; Lotze, M. T. PAMPs and DAMPs:
5
6
7 Signal 0s That Spur Autophagy and Immunity. Immunol Rev 2012, 249 (1), 158–175.
8
9
10 (2) Wu, J.; Sun, L.; Chen, X.; Du, F.; Shi, H.; Chen, C.; Chen, Z. J. Cyclic GMP-AMP
11
12 Is an Endogenous Second Messenger in Innate Immune Signaling by Cytosolic DNA.
13
14
15 Science 2013, 339 (6121), 826–830.
16
17
18 (3) Sun, L.; Wu, J.; Du, F.; Chen, X.; Chen, Z. J. Cyclic GMP-AMP Synthase Is a
19
20
Cytosolic DNA Sensor That Activates the Type I Interferon Pathway. Science 2013, 339
21
22
23 (6121), 786–791.
24
25
26 (4) Gao, P.; Ascano, M.; Wu, Y.; Barchet, W.; Gaffney, B. L.; Zillinger, T.; Serganov,
27
28
29 A. A.; Liu, Y.; Jones, R. A.; Hartmann, G.; Tuschl, T.; Patel, D. J. Cyclic [G(2′,5′)PA(3′,5′)p]
30
31 Is the Metazoan Second Messenger Produced by DNA-Activated Cyclic GMP-AMP
32
33
34 Synthase. Cell 2013, 153 (5), 1094–1107.
35
36
37 (5) Li, X.; Shu, C.; Yi, G.; Chaton, C. T.; Shelton, C. L.; Diao, J.; Zuo, X.; Kao, C. C.;
38
39
Herr, A. B.; Li, P. Cyclic GMP-AMP Synthase Is Activated by Double-Stranded DNA-
40
41
42 Induced Oligomerization. Immunity 2013, 39 (6), 1019–1031.
43
44
45 (6) Barrat, F. J.; Elkon, K. B.; Fitzgerald, K. A. Importance of Nucleic Acid
46
47
48 Recognition in Inflammation and Autoimmunity. Annu. Rev. Med. 2016, 67 (1), 323–336.
49
50 (7) Schlee, M.; Hartmann, G. Discriminating Self from Non-Self in Nucleic Acid
51
52
53 Sensing. Nat Rev Immunol 2016, 16 (9), 566–580.
54
55
56
57 40
58
59
60 ACS Paragon Plus Environment
Page 41 of 60 Journal of Chemical Information and Modeling

1
2
3
4 (8) Ishikawa, H.; Barber, G. N. STING Is an Endoplasmic Reticulum Adaptor That
5
6
7 Facilitates Innate Immune Signalling. Nature 2008, 455 (7213), 674–678.
8
9
10 (9) Ablasser, A.; Goldeck, M.; Cavlar, T.; Deimling, T.; Witte, G.; Röhl, I.; Hopfner,
11
12 K.-P.; Ludwig, J.; Hornung, V. CGAS Produces a 2′-5′-Linked Cyclic Dinucleotide Second
13
14
15 Messenger That Activates STING. Nature 2013, 498 (7454), 380–384.
16
17
18 (10) Zhang, X.; Shi, H.; Wu, J.; Zhang, X.; Sun, L.; Chen, C.; Chen, Z. J. Cyclic GMP-
19
20
AMP Containing Mixed Phosphodiester Linkages Is An Endogenous High-Affinity Ligand
21
22
23 for STING. Mol. Cell 2013, 51 (2), 226–235.
24
25
26 (11) Diner, E. J.; Burdette, D. L.; Wilson, S. C.; Monroe, K. M.; Kellenberger, C. A.;
27
28
29 Hyodo, M.; Hayakawa, Y.; Hammond, M. C.; Vance, R. E. The Innate Immune DNA Sensor
30
31 CGAS Produces a Noncanonical Cyclic Dinucleotide That Activates Human STING. Cell
32
33
34 Rep. 2013, 3 (5), 1355–1361.
35
36
37 (12) Ishikawa, H.; Ma, Z.; Barber, G. N. STING Regulates Intracellular DNA-
38
39
Mediated, Type I Interferon-Dependent Innate Immunity. Nature 2009, 461 (7265), 788–
40
41
42 792.
43
44
45 (13) Shang, G.; Zhang, C.; Chen, Z. J.; Bai, X.; Zhang, X. Cryo-EM Structures of
46
47
48 STING Reveal Its Mechanism of Activation by Cyclic GMP–AMP. Nature 2019, 567
49
50 (7748), 389–393.
51
52
53 (14) Zhang, C.; Shang, G.; Gui, X.; Zhang, X.; Bai, X.; Chen, Z. J. Structural Basis of
54
55
56
57 41
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 42 of 60

1
2
3
4 STING Binding with and Phosphorylation by TBK1. Nature 2019, 567 (7748), 394–398.
5
6
7 (15) Crow, Y. J.; Hayward, B. E.; Parmar, R.; Robins, P.; Leitch, A.; Ali, M.; Black,
8
9
10 D. N.; van Bokhoven, H.; Brunner, H. G.; Hamel, B. C.; Corry, P. C.; Cowan, F. M.; Frints,
11
12 S. G.; Klepper, J.; Livingston, J. H.; Lynch, S. A.; Massey, R. F.; Meritet, J. F.; Michaud, J.
13
14
15 L.; Ponsot, G.; Voit, T.; Lebon, P.; Bonthron, D. T.; Jackson, A. P.; Barnes, D. E.; Lindahl,
16
17
18 T. Mutations in the Gene Encoding the 3’-5’ DNA Exonuclease TREX1 Cause Aicardi-
19
20
Goutieres Syndrome at the AGS1 Locus. Nat Genet 2006, 38 (8), 917–920.
21
22
23 (16) Rice, G. I.; Forte, G. M.; Szynkiewicz, M.; Chase, D. S.; Aeby, A.; Abdel-Hamid,
24
25
26 M. S.; Ackroyd, S.; Allcock, R.; Bailey, K. M.; Balottin, U.; Barnerias, C.; Bernard, G.;
27
28
29 Bodemer, C.; Botella, M. P.; Cereda, C.; Chandler, K. E.; Dabydeen, L.; Dale, R. C.; De
30
31 Laet, C.; De Goede, C. G.; Del, T. M.; Effat, L.; Enamorado, N. N.; Fazzi, E.; Gener, B.;
32
33
34 Haldre, M.; Lin, J. P.; Livingston, J. H.; Lourenco, C. M.; Marques, W. J.; Oades, P.;
35
36
37 Peterson, P.; Rasmussen, M.; Roubertie, A.; Schmidt, J. L.; Shalev, S. A.; Simon, R.;
38
39
Spiegel, R.; Swoboda, K. J.; Temtamy, S. A.; Vassallo, G.; Vilain, C. N.; Vogt, J.;
40
41
42 Wermenbol, V.; Whitehouse, W. P.; Soler, D.; Olivieri, I.; Orcesi, S.; Aglan, M. S.; Zaki,
43
44
45 M. S.; Abdel-Salam, G. M.; Vanderver, A.; Kisand, K.; Rozenberg, F.; Lebon, P.; Crow, Y.
46
47
48 J. Assessment of Interferon-Related Biomarkers in Aicardi-Goutieres Syndrome Associated
49
50 with Mutations in TREX1, RNASEH2A, RNASEH2B, RNASEH2C, SAMHD1, and
51
52
53 ADAR: A Case-Control Study. Lancet Neurol 2013, 12 (12), 1159–1169.
54
55
56
57 42
58
59
60 ACS Paragon Plus Environment
Page 43 of 60 Journal of Chemical Information and Modeling

1
2
3
4 (17) Gray, E. E.; Treuting, P. M.; Woodward, J. J.; Stetson, D. B. Cutting Edge: CGAS
5
6
7 Is Required for Lethal Autoimmune Disease in the Trex1-Deficient Mouse Model of
8
9
10 Aicardi-Goutieres Syndrome. J Immunol 2015, 195 (5), 1939–1943.
11
12 (18) An, J.; Durcan, L.; Karr, R. M.; Briggs, T. A.; Rice, G. I.; Teal, T. H.; Woodward,
13
14
15 J. J.; Elkon, K. B. Expression of Cyclic GMP‐AMP Synthase in Patients With Systemic
16
17
18 Lupus Erythematosus. Arthritis Rheumatol 2017, 69 (4), 800–807.
19
20
(19) Kato, Y.; Park, J.; Takamatsu, H.; Konaka, H.; Aoki, W.; Aburaya, S.; Ueda, M.;
21
22
23 Nishide, M.; Koyama, S.; Hayama, Y.; Kinehara, Y.; Hirano, T.; Shima, Y.; Narazaki, M.;
24
25
26 Kumanogoh, A. Apoptosis-Derived Membrane Vesicles Drive the CGAS–STING Pathway
27
28
29 and Enhance Type I IFN Production in Systemic Lupus Erythematosus. Ann Rheum Dis
30
31 2018, annrheumdis-2018-212988.
32
33
34 (20) Rodero, M. P.; Tesser, A.; Bartok, E.; Rice, G. I.; Della Mina, E.; Depp, M.; Beitz,
35
36
37 B.; Bondet, V.; Cagnard, N.; Duffy, D.; Dussiot, M.; Frémond, M.-L.; Gattorno, M.;
38
39
Guillem, F.; Kitabayashi, N.; Porcheray, F.; Rieux-Laucat, F.; Seabra, L.; Uggenti, C.; Volpi,
40
41
42 S.; Zeef, L. A. H.; Alyanakian, M.-A.; Beltrand, J.; Bianco, A. M.; Boddaert, N.; Brouzes,
43
44
45 C.; Candon, S.; Caorsi, R.; Charbit, M.; Fabre, M.; Faletra, F.; Girard, M.; Harroche, A.;
46
47
48 Hartmann, E.; Lasne, D.; Marcuzzi, A.; Neven, B.; Nitschke, P.; Pascreau, T.; Pastore, S.;
49
50 Picard, C.; Picco, P.; Piscianz, E.; Polak, M.; Quartier, P.; Rabant, M.; Stocco, G.; Taddio,
51
52
53 A.; Uettwiller, F.; Valencic, E.; Vozzi, D.; Hartmann, G.; Barchet, W.; Hermine, O.; Bader-
54
55
56
57 43
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 44 of 60

1
2
3
4 Meunier, B.; Tommasini, A.; Crow, Y. J. Type I Interferon-Mediated Autoinflammation
5
6
7 Due to DNase II Deficiency. Nat Commun 2017, 8 (1), 2176.
8
9
10 (21) Kerur, N.; Fukuda, S.; Banerjee, D.; Kim, Y.; Fu, D.; Apicella, I.; Varshney, A.;
11
12 Yasuma, R.; Fowler, B. J.; Baghdasaryan, E.; Marion, K. M.; Huang, X.; Yasuma, T.;
13
14
15 Hirano, Y.; Serbulea, V.; Ambati, M.; Ambati, V. L.; Kajiwara, Y.; Ambati, K.; Hirahara,
16
17
18 S.; Bastos-Carvalho, A.; Ogura, Y.; Terasaki, H.; Oshika, T.; Kim, K. B.; Hinton, D. R.;
19
20
Leitinger, N.; Cambier, J. C.; Buxbaum, J. D.; Kenney, M. C.; Jazwinski, S. M.; Nagai, H.;
21
22
23 Hara, I.; West, A. P.; Fitzgerald, K. A.; Sadda, S. R.; Gelfand, B. D.; Ambati, J. CGAS
24
25
26 Drives Noncanonical-Inflammasome Activation in Age-Related Macular Degeneration. Nat
27
28
29 Med 2018, 24 (1), 50–61.
30
31 (22) Zhao, Q.; Wei, Y.; Pandol, S. J.; Li, L.; Habtezion, A. STING Signaling Promotes
32
33
34 Inflammation in Experimental Acute Pancreatitis. Gastroenterology 2018, 154 (6), 1822-
35
36
37 1835.e2.
38
39
(23) Cao, D. J.; Schiattarella, G. G.; Villalobos, E.; Jiang, N.; May, H. I.; Li, T.; Chen,
40
41
42 Z. J.; Gillette, T. G.; Hill, J. A. Cytosolic DNA Sensing Promotes Macrophage
43
44
45 Transformation and Governs Myocardial Ischemic Injury. Circulation 2018, 137 (24),
46
47
48 2613–2634.
49
50 (24) Ablasser, A.; Chen, Z. J. CGAS in Action: Expanding Roles in Immunity and
51
52
53 Inflammation. Science 2019, 363 (6431), eaat8657.
54
55
56
57 44
58
59
60 ACS Paragon Plus Environment
Page 45 of 60 Journal of Chemical Information and Modeling

1
2
3
4 (25) Li, T.; Chen, Z. J. The CGAS–CGAMP–STING Pathway Connects DNA Damage
5
6
7 to Inflammation, Senescence, and Cancer. J Exp Med 2018, 215 (5), 1287–1299.
8
9
10 (26) An, J.; Woodward, J. J.; Lai, W.; Minie, M.; Sun, X.; Tanaka, L.; Snyder, J. M.;
11
12 Sasaki, T.; Elkon, K. B. Inhibition of Cyclic GMP-AMP Synthase Using a Novel
13
14
15 Antimalarial Drug Derivative in Trex1 -Deficient Mice. Arthritis Rheumatol 2018, 70 (11),
16
17
18 1807–1819.
19
20
(27) An, J.; Woodward, J. J.; Sasaki, T.; Minie, M.; Elkon, K. B. Cutting Edge:
21
22
23 Antimalarial Drugs Inhibit IFN-β Production through Blockade of Cyclic GMP-AMP
24
25
26 Synthase–DNA Interaction. J.I. 2015, 194 (9), 4089–4093.
27
28
29 (28) Wang, M.; Sooreshjani, M. A.; Mikek, C.; Opoku-Temeng, C.; Sintim, H. O.
30
31 Suramin Potently Inhibits CGAMP Synthase, CGAS, in THP1 Cells to Modulate IFN-β
32
33
34 Levels. Future Med Chem 2018, 10 (11), 1301–1317.
35
36
37 (29) Vincent, J.; Adura, C.; Gao, P.; Luz, A.; Lama, L.; Asano, Y.; Okamoto, R.;
38
39
Imaeda, T.; Aida, J.; Rothamel, K.; Gogakos, T.; Steinberg, J.; Reasoner, S.; Aso, K.; Tuschl,
40
41
42 T.; Patel, D. J.; Glickman, J. F.; Ascano, M. Small Molecule Inhibition of CGAS Reduces
43
44
45 Interferon Expression in Primary Macrophages from Autoimmune Mice. Nat Commun 2017,
46
47
48 8 (1), 750.
49
50 (30) Lama, L.; Adura, C.; Xie, W.; Tomita, D.; Kamei, T.; Kuryavyi, V.; Gogakos, T.;
51
52
53 Steinberg, J. I.; Miller, M.; Ramos-Espiritu, L.; Asano, Y.; Hashizume, S.; Aida, J.; Imaeda,
54
55
56
57 45
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 46 of 60

1
2
3
4 T.; Okamoto, R.; Jennings, A. J.; Michino, M.; Kuroita, T.; Stamford, A.; Gao, P.; Meinke,
5
6
7 P.; Glickman, J. F.; Patel, D. J.; Tuschl, T. Development of Human CGAS-Specific Small-
8
9
10 Molecule Inhibitors for Repression of DsDNA-Triggered Interferon Expression. Nat
11
12 Commun 2019, 10 (1), 2261.
13
14
15 (31) Hall, J.; Brault, A.; Vincent, F.; Weng, S.; Wang, H.; Dumlao, D.; Aulabaugh, A.;
16
17
18 Aivazian, D.; Castro, D.; Chen, M.; Culp, J.; Dower, K.; Gardner, J.; Hawrylik, S.;
19
20
Golenbock, D.; Hepworth, D.; Horn, M.; Jones, L.; Jones, P.; Latz, E.; Li, J.; Lin, L.-L.;
21
22
23 Lin, W.; Lin, D.; Lovering, F.; Niljanskul, N.; Nistler, R.; Pierce, B.; Plotnikova, O.; Schmitt,
24
25
26 D.; Shanker, S.; Smith, J.; Snyder, W.; Subashi, T.; Trujillo, J.; Tyminski, E.; Wang, G.;
27
28
29 Wong, J.; Lefker, B.; Dakin, L.; Leach, K. Discovery of PF-06928215 as a High Affinity
30
31 Inhibitor of CGAS Enabled by a Novel Fluorescence Polarization Assay. PLoS ONE 2017,
32
33
34 12 (9), e0184843.
35
36
37 (32) Shoichet, B. K. Virtual Screening of Chemical Libraries. Nature 2004, 432 (7019),
38
39
862–865.
40
41
42 (33) Zhou, W.; Whiteley, A. T.; de Oliveira Mann, C. C.; Morehouse, B. R.; Nowak,
43
44
45 R. P.; Fischer, E. S.; Gray, N. S.; Mekalanos, J. J.; Kranzusch, P. J. Structure of the Human
46
47
48 CGAS–DNA Complex Reveals Enhanced Control of Immune Surveillance. Cell 2018, 174
49
50 (2), 300-311.e11.
51
52
53 (34) Hooy, R. M.; Sohn, J. The Allosteric Activation of CGAS Underpins Its Dynamic
54
55
56
57 46
58
59
60 ACS Paragon Plus Environment
Page 47 of 60 Journal of Chemical Information and Modeling

1
2
3
4 Signaling Landscape. eLife 2018, 7, e39984.
5
6
7 (35) Minor, W.; Cymborowski, M.; Otwinowski, Z.; Chruszcz, M. HKL-3000: The
8
9
10 Integration of Data Reduction and Structure Solution--from Diffraction Images to an Initial
11
12 Model in Minutes. Acta Crystallogr. D Biol. Crystallogr. 2006, 62 (Pt 8), 859–866.
13
14
15 (36) Collaborative Computational Project, Number 4. The CCP4 Suite: Programs for
16
17
18 Protein Crystallography. Acta Crystallogr. D Biol. Crystallogr. 1994, 50 (Pt 5), 760–763.
19
20
(37) Adams, P. D.; Grosse-Kunstleve, R. W.; Hung, L. W.; Ioerger, T. R.; McCoy, A.
21
22
23 J.; Moriarty, N. W.; Read, R. J.; Sacchettini, J. C.; Sauter, N. K.; Terwilliger, T. C. PHENIX:
24
25
26 Building New Software for Automated Crystallographic Structure Determination. Acta
27
28
29 Crystallogr. D Biol. Crystallogr. 2002, 58 (Pt 11), 1948–1954.
30
31 (38) Emsley, P.; Cowtan, K. Coot: Model-Building Tools for Molecular Graphics. Acta
32
33
34 Crystallogr. D Biol. Crystallogr. 2004, 60 (Pt 12 Pt 1), 2126–2132.
35
36
37 (39) Case, D. A.; Babin, V.; Berryman, J. T.; Betz, R. M.; Cai, Q.; Cerutti, D. S.;
38
39
Cheatham, T. E. I.; Darden, T. A.; Duke, R. E.; Gohlke, H.; Goetz, A. W.; Gusarov, S.;
40
41
42 Homeyer, N.; Janowski, P.; Kaus, J.; Kolossvary, I.; Kovalenko, A.; Lee, T. S.; LeGrand,
43
44
45 S.; Luchko, T.; Luo, R.; Madej, B.; Merz, K. M.; Paesani, F.; Roe, D. R.; Roitberg, A.; Sagui,
46
47
48 C.; Salomon-Ferrer, R.; Seabra, G.; Simmerling, C. L.; Smith, W.; Swails, J.; Walker, R.
49
50 C.; Wang, J.; Wolf, R. M.; Wu, X.; Kollman, P. A. AMBER 14; University of California,
51
52
53 San Francisco: San Francisco, CA, 2014.
54
55
56
57 47
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 48 of 60

1
2
3
4 (40) M. J. Frisch, G. W. T. H. Gaussian 09, Revision A.02; Gaussian, Inc, 2016.
5
6
7 (41) Wang, J.; Wang, W.; Kollman, P. A.; Case, D. A. Automatic Atom Type and Bond
8
9
10 Type Perception in Molecular Mechanical Calculations. J Mol Graph Model 2006, 25 (2),
11
12 247–260.
13
14
15 (42) Wang, J.; Wolf, R. M.; Caldwell, J. W.; Kollman, P. A.; Case, D. A. Development
16
17
18 and Testing of a General Amber Force Field. J Comput Chem 2004, 25 (9), 1157–1174.
19
20
(43) Maier, J. A.; Martinez, C.; Kasavajhala, K.; Wickstrom, L.; Hauser, K. E.;
21
22
23 Simmerling, C. Ff14SB: Improving the Accuracy of Protein Side Chain and Backbone
24
25
26 Parameters from Ff99SB. J Chem Theory Comput 2015, 11 (8), 3696–3713.
27
28
29 (44) Lippert, R. A.; Bowers, K. J.; Dror, R. O.; Eastwood, M. P.; Gregersen, B. A.;
30
31 Klepeis, J. L.; Kolossvary, I.; Shaw, D. E. A Common, Avoidable Source of Error in
32
33
34 Molecular Dynamics Integrators. J Chem Phys 2007, 126 (4), 046101.
35
36
37 (45) Cerutti, D. S.; Duke, R. E.; Darden, T. A.; Lybrand, T. P. Staggered Mesh Ewald:
38
39
An Extension of the Smooth Particle-Mesh Ewald Method Adding Great Versatility. J Chem
40
41
42 Theory Comput 2009, 5 (9), 2322.
43
44
45 (46) Salomon-Ferrer, R.; Gotz, A. W.; Poole, D.; Le Grand, S.; Walker, R. C. Routine
46
47
48 Microsecond Molecular Dynamics Simulations with AMBER on GPUs. 2. Explicit Solvent
49
50 Particle Mesh Ewald. J Chem Theory Comput 2013, 9 (9), 3878–3888.
51
52
53 (47) Miller, B. R.; McGee, T. J.; Swails, J. M.; Homeyer, N.; Gohlke, H.; Roitberg, A.
54
55
56
57 48
58
59
60 ACS Paragon Plus Environment
Page 49 of 60 Journal of Chemical Information and Modeling

1
2
3
4 E. MMPBSA.Py: An Efficient Program for End-State Free Energy Calculations. J Chem
5
6
7 Theory Comput 2012, 8 (9), 3314–3321.
8
9
10 (48) Harder, E.; Damm, W.; Maple, J.; Wu, C.; Reboul, M.; Xiang, J. Y.; Wang, L.;
11
12 Lupyan, D.; Dahlgren, M. K.; Knight, J. L.; Kaus, J. W.; Cerutti, D. S.; Krilov, G.; Jorgensen,
13
14
15 W. L.; Abel, R.; Friesner, R. A. OPLS3: A Force Field Providing Broad Coverage of Drug-
16
17
18 like Small Molecules and Proteins. J Chem Theory Comput 2016, 12 (1), 281–296.
19
20
(49) Schrödinger Release 2015-4: LigPrep; Schrödinger,LLC, 2015.
21
22
23 (50) Halgren, T. A.; Murphy, R. B.; Friesner, R. A.; Beard, H. S.; Frye, L. L.; Pollard,
24
25
26 W. T.; Banks, J. L. Glide: A New Approach for Rapid, Accurate Docking and Scoring. 2.
27
28
29 Enrichment Factors in Database Screening. J Med Chem 2004, 47 (7), 1750–1759.
30
31 (51) Sastry, M.; Lowrie, J. F.; Dixon, S. L.; Sherman, W. Large-Scale Systematic
32
33
34 Analysis of 2D Fingerprint Methods and Parameters to Improve Virtual Screening
35
36
37 Enrichments. J Chem Inf Model 2010, 50 (5), 771–784.
38
39
(52) Sastry, G. M.; Dixon, S. L.; Sherman, W. Rapid Shape-Based Ligand Alignment
40
41
42 and Virtual Screening Method Based on Atom/Feature-Pair Similarities and Volume
43
44
45 Overlap Scoring. J Chem Inf Model 2011, 51 (10), 2455–2466.
46
47
48 (53) Xu, Z.; Zhang, Q.; Shi, J.; Zhu, W. Underestimated Noncovalent Interactions in
49
50 Protein Data Bank. J Chem Inf Model 2019, 59 (8), 3389–3399.
51
52
53 (54) Seetoh, W.-G.; Abell, C. Disrupting the Constitutive, Homodimeric Protein–
54
55
56
57 49
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 50 of 60

1
2
3
4 Protein Interface in CK2β Using a Biophysical Fragment-Based Approach. J. Am. Chem.
5
6
7 Soc. 2016, 138 (43), 14303–14311.
8
9
10 (55) Simeonov, A. Recent Developments in the Use of Differential Scanning
11
12 Fluorometry in Protein and Small Molecule Discovery and Characterization. Expert Opin
13
14
15 Drug Discov 2013, 8 (9), 1071–1082.
16
17
18 (56) Hall, J.; Ralph, E. C.; Shanker, S.; Wang, H.; Byrnes, L. J.; Horst, R.; Wong, J.;
19
20
Brault, A.; Dumlao, D.; Smith, J. F.; Dakin, L. A.; Schmitt, D. C.; Trujillo, J.; Vincent, F.;
21
22
23 Griffor, M.; Aulabaugh, A. E. The Catalytic Mechanism of Cyclic GMP-AMP Synthase
24
25
26 (CGAS) and Implications for Innate Immunity and Inhibition: Catalytic Mechanism of
27
28
29 CGAS. Protein Sci. 2017, 26 (12), 2367–2380.
30
31 (57) Du, M.; Chen, Z. J. DNA-Induced Liquid Phase Condensation of CGAS Activates
32
33
34 Innate Immune Signaling. Science 2018, 361 (6403), 704–709.
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57 50
58
59
60 ACS Paragon Plus Environment
Page 51 of 60 Journal of Chemical Information and Modeling

1
2
3
4
5
6 For Table of Contents Use Only
7
8
9
10
11
12
Inhibitors of Human Cyclic GMP-AMP
13
14
15
Synthase: A Cross-validation Study of Molecular
16
17 Docking and Experimental Testing
18
19
20 Wenfeng Zhao†,‡,§,#, Muya Xiong‡,§,#, Xiaojing Yuan‡,§, Minjun Li⊥, Hongbing Sun*,†,
21
22
Yechun Xu*,‡,§
23
24 †Jiangsu
25 Key Laboratory of Drug Discovery for Metabolic Disease and State
26
27 Key Laboratory of Natural Medicines, China Pharmaceutical University, Nanjing
28
29
30 210009, P.R. China
31
32
33
‡CAS Key Laboratory of Receptor Research, Drug Discovery and Design
34
35
36
37
Center, Shanghai Institute of Materia Medica, Chinese Academy of Sciences,
38
39 Shanghai 201203, China
40
41
42
§University of Chinese Academy of Sciences, Beijing 100049, China
43
44
45
46 ⊥Shanghai
47 Synchrotron Radiation Facility, Shanghai Advanced Research
48
49 Institute, Chinese Academy of Sciences, Shanghai 201210, China
50
51
52
53
54
55
56
57 51
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 52 of 60

1
2
3 *Corresponding Author.
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57 52
58
59
60 ACS Paragon Plus Environment
Page 53 of 60 Journal of Chemical Information and Modeling

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24 Figure 1. Chemical structures of reported cGAS inhibitors.
25
26 158x85mm (300 x 300 DPI)
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 54 of 60

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36 Figure 2. Crystal structures of h-cGASCD in complex with PF-06928215. (a) The 2Fo–Fc electron density
37 map of the bound PF-06928215 contoured at 1.6 σ level. (b) The structure of PF-06928215 in the high-
38 resolution crystal structure (PDB code: 6LRC, yellow) superimposed to that in 6NAO (green). (c)
39 Intermolecular contacts among PF-06928215, water molecules and amino acids lining the binding pocket of
h-cGASCD.
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 55 of 60 Journal of Chemical Information and Modeling

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19 Figure 3. Energy decomposition calculations on the h-cGASCD-PF-06928215 complex. (a) The contributions
20 of van der Waals and electrostatic interactions of residues (171-510) to the binding of PF-06928215 with h-
21 cGASCD. (b) Contributions of these residues positioned less than 5 Å of PF-06928215.
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 56 of 60

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18 Figure 4. The melting temperature of h-cGASCD with or without compounds measured by the thermal shift
19 assay. (a and b) The melting curve of monomer (gray) and dimer (blue) of h-cGASCD. (c and d) The melting
20 curve of the dimeric h-cGASCD in the absence (blue) and presence of PF-06928215(orange).
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 57 of 60 Journal of Chemical Information and Modeling

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38 Figure 5. Superposition of the docking conformation (gray) and bound conformation of (a) compounds 3, (b)
39 17, (c) 23 and, (d) 40 (PDB code: 6LRE, 6LRI, 6LRJ, and 6LRK, respectively) in their crystal structures with
40 h-cGASCD.
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 58 of 60

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31 Figure 6. Inhibition of the FL-h-cGAS by 59 compounds measured by the PPiase-coupled assay. (a)
32 Inhibition of the FL-h-cGAS by the 59 compounds at a concentration of 100 μM or 50 μM (for compounds
with a low solubility). (b) Chemical structure and dose-dependent inhibition of compound 18 against the FL-
33 h-cGAS.
34
35 1112x832mm (96 x 96 DPI)
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 59 of 60 Journal of Chemical Information and Modeling

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35 Figure 7. Inhibition of the FL-h-cGAS by 21 compounds measured by the PPiase-coupled assay. (a)
36 Inhibition of the FL-h-cGAS by 21 compounds (at 100 μM) resulted from the similarity searching of
37 compound 18. (b) Chemical structure and dose-dependent inhibition of compounds S2 and S3 against the
38 FL-h-cGAS.
39
1045x927mm (96 x 96 DPI)
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Journal of Chemical Information and Modeling Page 60 of 60

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28 Figure 8. The binding modes of three inhibitors (compounds 18, S2 and S3) with h-cGAS revealed by a
29 crystal structure together with molecular docking. (a) The 2Fo–Fc electron density map of the bound S2
30 contoured at 1.6 σ level (PDB code: 6LRL). (b-d) Superimposition of docked conformations of three
31 inhibitors, compounds 18 (c), S2 (d) and S3 (e).
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment

You might also like