Next Article in Journal
A Small Molecule That Promotes Cellular Senescence Prevents Fibrogenesis and Tumorigenesis
Next Article in Special Issue
Development of Novel Ecto-Nucleotide Pyrophosphatase/Phosphodiesterase 1 (ENPP1) Inhibitors for Tumor Immunotherapy
Previous Article in Journal
Age and Chronodisruption in Mouse Heart: Effect of the NLRP3 Inflammasome and Melatonin Therapy
Previous Article in Special Issue
The Role of DAMPS in Burns and Hemorrhagic Shock Immune Response: Pathophysiology and Clinical Issues. Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Control of STING Agonistic/Antagonistic Activity Using Amine-Skeleton-Based c-di-GMP Analogues

1
National Institute of Health Sciences, 3-25-26 Tonomachi, Kawasaki 210-9501, Japan
2
Graduate School of Medical Life Science, Yokohama City University, Yokohama 230-0045, Japan
3
Graduate School of Medicine, Dentistry and Pharmaceutical Sciences, Division of Pharmaceutical Science of Okayama University, 1-1-1 Tsushimanaka, Kita 700-8530, Japan
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(12), 6847; https://doi.org/10.3390/ijms23126847
Submission received: 2 June 2022 / Revised: 15 June 2022 / Accepted: 18 June 2022 / Published: 20 June 2022
(This article belongs to the Special Issue Small Molecules, Influence of Molecular Pathways 2.0)

Abstract

:
Stimulator of Interferon Genes (STING) is a type of endoplasmic reticulum (ER)-membrane receptor. STING is activated by a ligand binding, which leads to an enhancement of the immune-system response. Therefore, a STING ligand can be used to regulate the immune system in therapeutic strategies. However, the natural (or native) STING ligand, cyclic-di-nucleotide (CDN), is unsuitable for pharmaceutical use because of its susceptibility to degradation by enzymes and its low cell-membrane permeability. In this study, we designed and synthesized CDN derivatives by replacing the sugar-phosphodiester moiety, which is responsible for various problems of natural CDNs, with an amine skeleton. As a result, we identified novel STING ligands that activate or inhibit STING. The cyclic ligand 7, with a cyclic amine structure containing two guanines, was found to have agonistic activity, whereas the linear ligand 12 showed antagonistic activity. In addition, these synthetic ligands were more chemically stable than the natural ligands.

1. Introduction

Innate immunity is the first defense mechanism against pathogens. Unlike acquired immunity, innate immunity functions independently of the type of foreign substance, thereby leading to a rapid defense response against pathogens. The recognition of pathogens in innate immunity is mediated by pattern-recognition receptors (PRRs), which detect characteristic molecular patterns of bacteria and viruses (pathogen-associated molecular patterns, PAMP) and signals called damage/danger-associated molecular patterns (DAMP), which are released from injured cells [1]. Recently, cyclic guanosine monophosphate-adenosine monophosphate synthase (cGAS), which is a PRR that recognizes double-stranded DNA (dsDNA) in the cytoplasm, was discovered in mammalian cells [2,3,4,5,6]. cGAS detects cytoplasmic DNA and activates downstream-signaling cascades to stimulate the production of type I interferon and to induce innate immune responses by the cGAS-stimulator of the interferon genes’ (STING)-tank-binding kinase 1 (TBK1) pathway [7]. Thus, cGAS plays an important role in enhancing cytokine production [8].
cGAS recognizes dsDNA in the cytoplasm that is either self-derived (particularly in senescent cells and cancer cells) or exogenous (from bacteria, DNA viruses, and RNA viruses) [2] and produces 2′,3′-cyclic guanosine monophosphate-adenosine monophosphate (2′,3′-cGAMP) [9,10,11,12]. 2′,3′-cGAMP acts as a second messenger that binds to and activates a type of homodimeric endoplasmic reticulum (ER)-membrane receptor, STING [13,14]. In addition, cyclic dinucleotides (CDNs), such as bis-(3′-5′)-cyclic dimeric guanosine monophosphate (c-di-GMP) [15] and bis-(3′-5′)-cyclic dimeric adenosine monophosphate (c-di-AMP), are also unique second messengers produced by bacteria that promote STING-mediated immune activation (Figure 1) [16,17]. Ligand-bound STING migrates from ER to the Golgi apparatus [18] and then undergoes palmitoylation of Cys88/91, resulting in oligomerization [19,20]. The oligomerized STING binds to TBK1 via the carbon-terminal tail (CTT), causing phosphorylation (activation) of TBK1 and activated TBK1 in STING phosphorylates interferon regulatory-factor 3 (IRF3). This series of events promote the dimerization and nuclear translocation of IRF3 [21,22]. In addition, activated STING promotes nuclear-factor kappa B (NF-κB) activation and nuclear translocation via I kappa B kinase (IKK) activation, promoting cytokine production [23]. STING is finally transported to the lysosome for degradation, and STING-mediated signaling converges [24]. STING activation enhances the immune-system response; therefore, it is expected to be utilized in cancer immunotherapy [25,26]. For example, tumor formation was effectively suppressed by combined treatment with STING agonists [27,28,29,30,31,32,33] and adjuvant agents [34,35,36,37,38]. STING activation has also been reported to inhibit viral infections. Since this is an immune-enhancing effect rather than targeting a specific pathogen, it is effective against a wide range of targets, such as human immunodeficiency virus (HIV) and severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) [39,40,41]. In addition, homeostatic activation of STING causes abnormal immune activation and cytokine production. Thus, STING is a causative protein for autoimmune diseases, such as amyotrophic lateral sclerosis (ALS) and Sjögren’s syndrome, for which there is still no cure [42,43,44]. These observations suggest that STING antagonists [45,46,47] may be potential therapeutic agents for treating diseases related to the homeostatic activation of STING. Therefore, STING has attracted attention as an important drug-discovery target, potentially involved in various diseases. In fact, many studies on STING ligands have been reported in recent years [27,28,29,30,31,32,33,45,46,47]. However, most of the reported agonists were developed to target cancers, and there are few examples of their application in the treatment of infectious diseases [25,26,34,35,36,37,38,39,40,41]. Most antagonists developed are covalent ligands, which raises concerns about specificity, and there are few reports of non-covalent antagonists, many of which have low affinity [45,46,47]. Therefore, there is still a need to develop ligands that target STING because of its diverse roles.
CDN molecules, the natural ligands of STING, have un-druglike properties such as low cell-membrane permeability and are susceptible to inactivation by phosphodiesterase (PDE). These drawbacks must be overcome for CDNs to be used as medicinal drugs. In addition, CDN molecules are difficult to derivatize because of their structure, sugar-phosphate backbones, which are highly polar and reactive moieties, making their chemical synthesis challenging. Most of the currently reported CDN derivatives are limited to those with a phosphodiester skeleton, which are prone to degradation by PDE [48,49,50,51,52,53]. Although several research groups have reported CDN derivatives where the phosphodiester-sugar backbone is replaced with a different linker structure [54,55,56], there are no examples of CDN-based STING ligands with substitution of the entire sugar-phosphodiester backbone. In this study, we designed new CDN mimetics with the phosphodiester skeleton replaced by an amine skeleton, which is easy to synthesize and derivatize. We identified that the amine-skeleton-based molecules activate or inhibit STING. That is, the cyclic ligand with a cyclic amine structure containing two guanines was found to have agonistic activity, whereas the linear ligand showed antagonistic activity.

2. Results and Discussion

2.1. Design and Synthesis of Compounds

X-ray analysis of c-di-GMP and STING showed that the two guanine moieties of c-di-GMP interact strongly with Tyr164 of STING, via hydrogen bonds and π–π stacking interactions [57]. In contrast, the contribution of the sugar-phosphodiester moiety of c-di-GMP to the interaction with STING is smaller than those of the guanine moieties. Thus, we hypothesized that CDN derivatives with an improved membrane permeability and chemical stability against PDE can be developed, by replacing the sugar-phosphodiester moiety of c-di-GMP with other suitable skeletons.
In the new molecular design, cyclic amines were selected to construct the c-di-GMP mimics. Polyamines are readily available to construct a variety of cyclic structures. In addition to the cyclic analogues, linear-amine derivatives were also synthesized. Each derivative was prepared by introducing the guanine moieties into the cyclic and linear-amine backbones via a condensation reaction (Scheme 1 and Scheme 2) [58].

2.2. IFN-β Induction/Inhibition Activity of CDN Derivatives with an Amine Skeleton

The interaction of the synthesized CDN derivatives with STING was evaluated by a dual-luciferase reporter-gene assay. HEK293T cells, which are reported not to express STING, were employed for this assay [2]. Initially, we evaluated whether a STING-dependent increase in IFN-β gene expression is observed by transient transfection with the gene encoding STING (Figure S1).
An increase in fluorescence intensity was observed by adding cGAMP, confirming that our evaluation system can assess STING-dependent interferon induction. The result showed that compound 7 with a cyclic skeleton exhibited STING-dependent interferon-induction activity, although this was weaker than with cGAMP (Figure 2 and Figure S2). Interestingly, the results also suggested that compound 12 with a linear skeleton displayed antagonistic activity (Figure 3).
The observed reversal of activity between cyclic compound 7 and linear compound 12 might be attributed to the flexibility of the linker structure. STING is known to change its conformation into several structures upon ligand binding [59], and those that form the same conformational oligomerization to induce subsequent signaling [2]. Moreover, the activity of STING ligands has been reported to be higher in compounds with high rigidity [60], suggesting that the binding of rigid ligands contributes to the activation of STING by fixing the conformation of the protein. For further investigation, we synthesized several linear compounds and evaluated their IFN-β gene-induction activities. That results also showed that linear-amine-based compounds tend to exhibit antagonist activity (Table S1).
In this assay, cGAMP requires digitonin treatment to permeabilize the cell membrane, whereas amine-skeleton-based compound 7 showed relatively higher activity than cGAMP, even without digitonin treatment. When cells were not permeabilized with digitonin, the activity of cGAMP was markedly reduced to about 1/178; however, the decrease in the activity of compound 7 with the amine skeleton was relatively small, at about 1/27 (Figure S3, Table S2). Thus, replacing the amine skeleton appears to eliminate the negative charge and improve the cell-membrane permeability, when compared with CDNs with the sugar-phosphodiester moiety.

2.3. Stability of CDN Derivatives against Nuclease

The phosphodiester bond of CDNs has been reported to be digested by hydrolytic enzymes in vivo. Thus, to evaluate the chemical stability of the synthesized compounds, CDN derivatives were treated with nuclease P1 (NP1), which cleaves phosphodiester bonds to yield linear 5′-monophosphate nucleotides. The digested products in the reaction mixture were confirmed by HPLC analysis. In the presence of NP1, rapid digestion of CDNs was observed (Figure S4A,B). In contrast, no digestion was observed for the amine-skeleton cyclic compound 7 without the sugar-phosphate backbone (Figure S4C). This result indicates that it is possible to replace the sugar-phosphate moiety with an amine moiety, to obtain stability against at least NP1.

2.4. Docking Study between STING and the Binding Ligands

Finally, we performed docking simulations to analyze the binding mode of compounds 7 and 12, which showed agonist and antagonist activities toward STING, respectively. Docking simulations were performed for compounds 7 and 12 using the crystal structures of the active (PDB ID: 4LOH) and inactive (PDB ID: 6MXE) forms of STING, respectively. Since it has been reported that two molecules of cpd. 18, an existing STING inhibitor, bind to the ligand-binding domain (LBD) and inhibit the activity of STING [46], we simulated compound 12 with two molecules in addition to the simulation with a single molecule. The results suggest that compounds 7 and 12 bind to the ligand-binding site of STING in a structure similar to that of the ligand in their respective crystal structures (Figure 4 and Figure S5). In particular, the results indicated that compound 7 maintains π–π stacking interactions between the guanine base and Tyr164, which is thought to be important for the activation of STING (Figure 4A and Figure S5A). Docking simulations with two molecules of compound 12 suggest that compound 12 interacts with the broader surface of STING via a hydrogen bond with Pro264, as in cpd. 18, indicating that two molecules of compound 12 can bind to STING (Figure 4C and Figure S5C).

3. Materials and Methods

All reagents were obtained from commercial suppliers and were used as received, unless otherwise noted.

3.1. Synthesis and Characterization of Compounds

The synthesis procedure and analytical data for the synthesized compounds are provided in the Supporting Information.

3.2. Dual Luciferase-Reporter-Based Biological Evaluation in HEK293T Cells

HEK293T cells [61] were maintained in DMEM (Sigma-Aldrich, St. Louis, MO, USA) and were supplemented with 10% FBS (Sigma-Aldrich) and penicillin-streptomycin (Nacalai Tesque, Kyoto, Japan). The expression plasmid for human STING (Flag-STING) was constructed by inserting a fragment coding for STING into the p3×Flag-CMV10 (Flag-Empty; Sigma-Aldrich). The fragment was generated from a PCR product. HEK293T cells were plated in 24-well plates at 1 × 105 cells/well. The next day, the cells were transfected together with Flag-STING or Flag-Empty as well as IFN-β-firefly luciferase (#102597, Addgene, Watertown, MA, USA) [62] and SV40-Renilla luciferase (#E223A, Promega, Madison, WI, USA)-reporter constructs using Lipofectamine LTX (Thermo Fisher Scientific, Tokyo, Japan). Compounds were mixed with a 100× volume of digitonin permeabilization solution (50 mM HEPES, pH 7.0, 100 mM KCl, 3 mM MgCl2, 0.1 mM DTT, 85 mM sucrose, 0.2% BSA, 1 mM ATP, 0.1 mM GTP, 10 μg/mL digitonin) or DMEM medium. The medium was aspirated from the cells and replaced with 200 μL for each sample mixture. Cells were incubated for 30 min at 37 °C. Wells were again aspirated, and fresh DMEM medium (FBS: +, P/S: –, 500 μL/well) was added. Following stimulation for 6 h with the compounds, the cells were lysed in passive lysis buffer (Promega) for 15 min at 37 °C. The cell lysates were incubated with firefly luciferase substrate (Promega) and the Renilla luciferase-substrate coelenterazine (Promega), and the luminescence was measured using a Wallac ARVO Sx 1420 Multilabel Counter (PerkinElmer Japan, Chiba, Japan). The relative IFN-β expression was calculated as firefly luminescence relative to Renilla luminescence.

3.3. Stability Test against an Exonuclease

cGAMP (2′3′-cGAMP; InvivoGene, San Diego, CA, USA) and CDN derivatives (1 µg) were incubated in a solution (20 μL) containing the enzyme (New England Biolabs, Ipswich, MA, USA; 2.5-mU NP1 in 50 mM acetate buffer containing 1.25 mM Tris-HCl, 2.5 mM NaCl, 50 nM ZnCl2 (pH 5.5)) or 50 mM acetate buffer (negative control) for 15 min at 37 °C (water bath). The reaction was terminated by heat inactivation at 75 °C for 10 min. Ten microliters of each aliquot were injected directly into the HPLC (column: CAPCELL PAK MG-II [C18, 4.6 × 250 mm, 5 μm] (Osaka soda, Osaka, Japan); flow rate 0.6 mL/min; detection at 254 nm) for analysis. TEAA buffer (5 mM, pH 7.0) (solvent A) and acetonitrile (solvent B) were used as the mobile phase. The gradient was set as follows: 0–30 min: 0% B to 20% B; 30–35 min: 100% B; 35–40 min: 0% B (for c-di-GMP); or 0–30 min: 5% B to 50% B; 30–35 min: 100% B; 35–40 min: 5% B (for cGAMP and CDN derivatives).

3.4. Molecular Docking

Possible interactions between amine-skeleton CDN derivatives and STING (PDB IDs: 4LOH, 6MXE) were examined by performing docking studies using the Molecular Operating Environment (MOE) 2020.0901. The binding site of the protein was defined as the native ligand in the X-ray structure. Docking simulations of amine-skeleton CDN derivatives bound to STING were carried out using the standard protocols of general docking. The docking workflow followed an “induced fit” protocol that allowed the side chains of the receptor pocket to move according to the conformation of the ligand, and the positions were constrained. The weight used to lock the side-chain atoms in their original positions was 10. All docking poses were first ranked by the London dG score; then, a force field refinement was performed on the top 100 poses, followed by s rescoring of GBVI/WSA dG. The Amber14:EHT [63] force field was used for calculating the conformations.

4. Conclusions

In this study, we synthesized CDN analogues, in which the phosphate backbone is replaced with the amine moiety, as new STING ligands and found them to be stable against nucleases. In particular, cyclic-amine compound 7 and linear-amine compound 9 were found to have agonistic and antagonistic activities, respectively. Furthermore, docking simulations suggest that the agonistic (for 7) and antagonistic (for 9) actions observed in those molecules are due to differences in their interactions with STING. Although the activities of the amine-skeleton-based CDN molecules found in this study are still lower than that of cGAMP, it is possible to develop more active ligands by improving the physical properties, such as cell-membrane permeability and chemical stability. The derivatization of further CDN ligands that target STING and detailed studies of their mechanism of activity are currently underway.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/ijms23126847/s1.

Author Contributions

Conceptualization, G.T. and Y.D.; methodology, G.T. and N.S.; investigation, Y.Y., N.S., and M.N.; resources, G.T. and Y.D.; writing—original draft preparation, Y.Y.; writing—review and editing, M.N., N.S., G.T., and Y.D.; supervision, G.T.; project administration, Y.D.; funding acquisition, G.T. and Y.D. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported in part by grants from AMED under grant numbers JP21mk0101197 (to Y.D.) and JP21ak0101073 (to Y.D.), the Japan Society for the Promotion of Science (KAKENHI, grants JP19K16333 to G.T.; JP21K05320 to Y.D.; and JP18H05502 to Y.D.), the Sumitomo Foundation (to Y.D.), the Japan Foundation of Applied Enzymology (to Y.D.), the Kobayashi Foundation for Cancer Research (to Y.D.), the Foundation for Promotion of Cancer Research in Japan (to Y.D.), and the Tokushukai Scholarship Foundation (to Y.Y.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Palm, N.W.; Medzhitov, R. Pattern recognition receptors and control of adaptive immunity. Immunol. Rev. 2009, 227, 221–233. [Google Scholar] [CrossRef] [PubMed]
  2. Sun, L.; Wu, J.; Du, F.; Chen, X.; Chen, Z.J. Cyclic GMP-AMP synthase is a cytosolic DNA sensor that activates the type I interferon pathway. Science 2013, 339, 789–791. [Google Scholar] [CrossRef] [Green Version]
  3. Civril, F.; Deimling, T.; de Oliveira Mann, C.C.; Ablasser, A.; Moldt, M.; Witte, G.; Hornung, V.; Hopfner, K.P. Structural mechanism of cytosolic DNA sensing by cGAS. Nature 2013, 498, 332–337. [Google Scholar] [CrossRef] [Green Version]
  4. Kranzusch, P.J.; Lee, A.S.; Berger, J.M.; Doudna, J.A. Structure of human cGAS reveals a conserved family of second-messenger enzymes in innate immunity. Cell Rep. 2013, 3, 1362–1368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Li, X.; Shu, C.; Yi, G.; Chaton, C.T.; Shelton, C.L.; Diao, J.; Zuo, X.; Kao, C.C.; Herr, A.B.; Li, P. Cyclic GMP-AMP synthase is activated by double-stranded DNA-induced oligomerization. Immunity 2013, 39, 1019–1031. [Google Scholar] [CrossRef] [Green Version]
  6. Zhang, X.; Wu, J.; Du, F.; Xu, H.; Sun, L.; Chen, Z.; Brautigam, C.A.; Zhang, X.; Chen, Z.J. The Cytosolic DNA Sensor cGAS Forms an Oligomeric Complex with DNA and Undergoes Switch-like Conformational Changes in the Activation Loop. Cell Rep. 2014, 6, 421–430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Ishikawa, H.; Barber, G.N. The STING pathway and regulation of innate immune signaling in response to DNA pathogens. Cell. Mol. Life Sci. 2011, 68, 1157–1165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Ishikawa, H.; Barber, G.N. STING is an endoplasmic reticulum adaptor that facilitates innate immune signalling. Nature 2008, 455, 674–678. [Google Scholar] [CrossRef]
  9. Cai, X.; Chiu, Y.H.; Chen, Z.J. The cGAS-cGAMP-STING pathway of cytosolic DNA sensing and signaling. Mol. Cell 2014, 54, 289–296. [Google Scholar] [CrossRef] [Green Version]
  10. Ablasse, A.; Goldeck, M.; Cavlar, T.; Deimling, T.; Witte, G.; Röhl, I.; Hopfner, K.P.; Ludwig, J.; Hornung, V. cGAS produces a 2′-5′-linked cyclic dinucleotide second messenger that activates STING. Nature 2013, 498, 380–384. [Google Scholar] [CrossRef] [Green Version]
  11. Gao, P.; Ascano, M.; Wu, Y.; Barchet, W.; Gaffney, B.L.; Zillinger, T.; Serganov, A.A.; Liu, Y.; Jones, R.A.; Hartmann, G.; et al. Cyclic [G(2′,5′)pA(3′,5′)p] is the metazoan second messenger produced by DNA-activated cyclic GMP-AMP synthase. Cell 2013, 153, 1094–1107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Zhang, X.; Shi, H.; Wu, J.; Zhang, X.; Sun, L.; Chen, C.; Chen, Z.J. Cyclic GMP-AMP containing mixed phosphodiester linkages is an endogenous high-affinity ligand for STING. Mol. Cell 2013, 51, 226–235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Wu, J.; Sun, L.; Chen, X.; Du, F.; Shi, H.; Chen, C.; Chen, Z.J. Cyclic GMP-AMP Is an Endogenous Second Messenger in Innate Immune Signaling by Cytosolic DNA. Science 2013, 339, 826–830. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Shang, G.; Zhang, C.; Chen, Z.J.; Bai, X.C.; Zhang, X. Cryo-EM structures of STING reveal its mechanism of activation by cyclic GMP-AMP. Nature 2019, 567, 389–393. [Google Scholar] [CrossRef]
  15. Römling, U.; Galperin, M.Y.; Gomelsky, M. Cyclic di-GMP: The first 25 years of a universal bacterial second messenger. Microbiol. Mol. Biol. Rev. 2013, 77, 1–52. [Google Scholar] [CrossRef] [Green Version]
  16. Krasteva, P.V.; Sondermann, H. Versatile modes of cellular regulation via cyclic dinucleotides. Nat. Chem. Biol. 2017, 13, 350–359. [Google Scholar] [CrossRef]
  17. Burdette, D.L.; Monroe, K.M.; Sotelo-Troha, K.; Iwig, J.S.; Eckert, B.; Hyodo, M.; Hayakawa, Y.; Vance, R.E. STING is a direct innate immune sensor of cyclic di-GMP. Nature 2011, 478, 515–518. [Google Scholar] [CrossRef]
  18. Taguchi, T.; Mukai, K. Innate immunity signalling and membrane trafficking. Curr. Opin. Cell Biol. 2019, 59, 1–7. [Google Scholar] [CrossRef]
  19. Mukai, K.; Konno, H.; Akiba, T.; Uemura, T.; Waguri, S.; Kobayashi, T.; Barber, G.N.; Arai, H.; Taguchi, T. Activation of STING requires palmitoylation at the Golgi. Nat. Commun. 2016, 7, 11932. [Google Scholar] [CrossRef]
  20. Ergun, S.L.; Femandez, D.; Weiss, T.M.; Li, L. STING Polymer Structure Reveals Mechanisms for Activation, Hyperactivation, and Inhibition. Cell 2019, 178, 290–301. [Google Scholar] [CrossRef]
  21. Zhang, C.; Shang, G.; Gui, X.; Zhang, X.; Bai, X.C.; Chen, Z.J. Structural basis of STING binding with and phosphorylation by TBK1. Nature 2019, 567, 394–398. [Google Scholar] [CrossRef] [PubMed]
  22. Zhong, B.; Yang, Y.; Li, S.; Wang, Y.Y.; Li, Y.; Diao, F.; Lei, C.; He, X.; Zhang, L.; Tien, P.; et al. The adaptor protein MITA links virus-sensing receptors to IRF3 transcription factor activation. Immunity 2008, 29, 538–550. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. De Oliveira Man, C.C.; Orzalli, M.H.; King, D.S.; Kagan, J.C.; Lee, A.S.Y.; Kranzusch, P.J. Modular Architecture of the STING C-Terminal Tail Allows Interferon and NF-κB Signaling Adaptation. Cell Rep. 2019, 27, 1165–1175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Gonugunta, V.K.; Sakai, T.; Pokatayev, V.; Yang, K.; Wu, J.; Dobbs, N.; Yan, N. Trafficking-Mediated STING Degradation Requires Sorting to Acidified Endolysosomes and Can Be Targeted to Enhance Anti-tumor Response. Cell Rep. 2017, 21, 3234–3242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Barber, G.N. STING: Infection, inflammation and cancer. Nat. Rev. Immunol. 2015, 15, 760–770. [Google Scholar] [CrossRef] [Green Version]
  26. Yum, S.; Li, M.; Frankel, A.E.; Chen, Z.J. Roles of the cGAS-STING pathway in cancer immunosurveillance and immunotherapy. Annu. Rev. Cancer Biol. 2019, 3, 323–344. [Google Scholar] [CrossRef]
  27. Zhang, Y.; Sun, Z.; Pei, J.; Luo, Q.; Zeng, X.; Li, Q.; Yang, Z.; Quan, J. Identification of α-Mangostin as an Agonist of Human STING. ChemMedChem 2018, 13, 2057–2064. [Google Scholar] [CrossRef]
  28. Ramanjulu, J.M.; Pesiridis, G.S.; Yang, J.; Concha, N.; Singhaus, R.; Zhang, S.Y.; Tran, J.L.; Moore, P.; Lehmann, S.; Eberl, H.C.; et al. Design of amidobenzimidazole STING receptor agonists with systemic activity. Nature 2018, 564, 439–443. [Google Scholar] [CrossRef]
  29. Chin, E.N.; Yu, C.; Vartabedian, V.F.; Jia, Y.; Kuma, M.; Gam, A.M.; Vernier, W.; Ali, S.H.; Kissai, M.; Lazar, D.C.; et al. Antitumor activity of a systemic STING-activating non-nucleotide cGAMP mimetic. Science 2020, 369, 993–999. [Google Scholar] [CrossRef]
  30. Pa, B.S.; Perera, S.A.; Piesvaux, J.A.; Presland, J.P.; Schroeder, G.K.; Cumming, J.N.; Trotte, B.W.; Altman, M.D.; Buevich, A.V.; Cash, B.; et al. An orally available non-nucleotide STING agonist with antitumor activity. Science 2020, 369, eaba6098. [Google Scholar]
  31. Pryde, D.C.; Middya, S.; Banerjee, M.; Shrivastava, R.; Basu, S.; Ghosh, R.; Yadav, D.B.; Surya, A. The discovery of potent small molecule activators of human STING. Eur. J. Med. Chem. 2021, 209, 112869. [Google Scholar] [CrossRef]
  32. Kim, D.S.; Endo, A.; Fang, F.G.; Huang, K.C.; Bao, X.; Choi, H.; Majumder, U.; Shen, Y.Y.; Mathieu, S.; Zhu, X.; et al. Title E7766, a Macrocycle-Bridged Stimulator of Interferon Genes (STING) Agonist with Potent Pan-Genotypic Activity. ChemMedChem 2021, 16, 1741–1744. [Google Scholar] [CrossRef] [PubMed]
  33. Saito-Tarashima, N.; Kinoshita, M.; Igata, Y.; Kashiwabaraa, Y.; Minakawa, N. Replacement of oxygen with sulfur on the furanose ring of cyclic dinucleotides enhances the immunostimulatory effect via STING activation. RSC Med. Chem. 2021, 12, 1519–1524. [Google Scholar] [CrossRef] [PubMed]
  34. Li, T.; Cheng, H.; Yuan, H.; Xu, Q.; Shu, C.; Zhang, Y.; Xu, P.; Tan, J.; Rui, Y.; Li, P.; et al. Antitumor Activity of cGAMP via Stimulation of cGAS-cGAMP-STING-IRF3 Mediated Innate Immune Response. Sci. Rep. 2016, 6, 19049. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Mullard, A. Can innate immune system targets turn up the heat on ‘cold’ tumours? Nat. Rev. Drug Discov. 2018, 17, 3–5. [Google Scholar] [CrossRef] [PubMed]
  36. Shi, F.; Su, J.; Wang, J.; Liu, Z.; Wang, T. Activation of STING inhibits cervical cancer tumor growth through enhancing the anti-tumor immune response. Mol. Cell. Biochem. 2021, 476, 1015–1024. [Google Scholar] [CrossRef]
  37. Vasiyani, H.; Shinde, A.; Roy, M.; Mane, M.; Singh, K.; Singh, J.; Gohel, D.; Currim, F.; Vaidya, K.; Chhabria, M.; et al. The analog of cGAMP, c-di-AMP, activates STING mediated cell death pathway in estrogen-receptor negative breast cancer cells. Apoptosis 2021, 26, 293–306. [Google Scholar] [CrossRef]
  38. Ren, D.; Qin, G.; Zhao, J.; Sun, Y.; Zhang, B.; Li, D.; Wang, B.; Jin, X.; Wu, H. Metformin activates the STING/IRF3/IFN-β pathway by inhibiting AKT phosphorylation in pancreatic cancer. Am. J. Cancer Res. 2020, 10, 2851–2864. [Google Scholar]
  39. Li, M.; Ferretti, M.; Ying, B.; Descamps, H.; Lee, E.; Dittmar, M.; Lee, J.S.; Whig, K.; Kamalia, B.; Dohnalová, L.; et al. Pharmacological activation of STING blocks SARS-CoV-2 infection. Sci. Immunol. 2021, 6, eabi9007. [Google Scholar] [CrossRef]
  40. Rui, Y.; Su, J.; Shen, S.; Hu, Y.; Huang, D.; Zheng, W.; Lou, M.; Shi, Y.; Wang, M.; Chen, S.; et al. Unique and complementary suppression of cGAS-STING and RNA sensing-triggered innate immune responses by SARS-CoV-2 proteins. Signal Transduct. Target. Ther. 2021, 6, 123. [Google Scholar] [CrossRef]
  41. Gao, D.; Wu, J.; Wu, Y.T.; Du, F.; Aroh, C.; Yan, N.; Sun, L.; Chen, Z.J. Cyclic GMP-AMP Synthase Is an Innate Immune Sensor of HIV and Other Retroviruses. Science 2013, 341, 903–906. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Gao, D.; Li, T.; Li, X.D.; Chen, X.; Li, Q.Z.; Wight-Carter, M.; Chen, Z.J. Activation of cyclic GMP-AMP synthase by self-DNA causes autoimmune diseases. Proc. Natl. Acad. Sci. USA 2015, 112, E5699–E5705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. McCauley, M.E.; O’Rourke, J.G.; Yáñez, A.; Markman, J.L.; Ho, R.; Wang, X.; Chen, S.; Lall, D.; Jin, M.; Muhammad, A.K.M.G.; et al. C9orf72 in myeloid cells suppresses STING-induced inflammation. Nature 2020, 585, 96–101. [Google Scholar] [CrossRef] [PubMed]
  44. Papinska, J.; Bagavant, H.; Gmyrek, G.B.; Deshmukh, U.S. Pulmonary Involvement in a Mouse Model of Sjögren’s Syndrome Induced by STING Activation. Int. J. Mol. Sci. 2020, 21, 4512. [Google Scholar] [CrossRef]
  45. Haag, S.M.; Gulen, M.F.; Gmyrek, G.B.; Deshmukh, U.S. Targeting STING with covalent small-molecule inhibitors. Nature 2018, 559, 269–273. [Google Scholar] [CrossRef]
  46. Siu, T.; Altman, M.D.; Baltus, G.A.; Childers, M.; Ellis, J.M.; Gunaydin, H.; Hatch, H.; Ho, T.; Jewell, J.; Lacey, B.M.; et al. Discovery of a Novel cGAMP Competitive Ligand of the Inactive Form of STING. ACS Med. Chem. Lett. 2019, 10, 92–97. [Google Scholar] [CrossRef]
  47. Hong, Z.; Mei, J.; Li, C.; Bai, G.; Maimaiti, M.; Hu, H.; Yu, W.; Sun, L.; Zhang, L.; Cheng, D.; et al. STING inhibitors target the cyclic dinucleotide binding pocket. Proc. Natl. Acad. Sci. USA 2021, 118, e2105465118. [Google Scholar] [CrossRef]
  48. Li, L.; Yin, Q.; Kuss, P.; Maliga, Z.; Millán, J.L.; Wu, H.; Mitchison, T.J. Hydrolysis of 2′3′-cGAMP by ENPP1 and design of nonhydrolyzable analogs. Nat. Chem. Biol. 2014, 10, 1043–1048. [Google Scholar] [CrossRef] [Green Version]
  49. Gaffney, B.L.; Veliath, E.; Zhao, J.; Jones, R.A. One-flask syntheses of c-di-GMP and the [Rp,Rp] and [Rp,Sp] thiophosphate analogues. Org. Lett. 2010, 12, 3269–3271. [Google Scholar] [CrossRef] [Green Version]
  50. Corrale, L.; Glickman, L.H.; McWhirter, S.M.; Kanne, D.B.; Sivick, K.E.; Katibah, G.E.; Woo, S.R.; Lemmen, E.; Banda, T.; Leong, J.J.; et al. Direct Activation of STING in the Tumor Microenvironment Leads to Potent and Systemic Tumor Regression and Immunity. Cell Rep. 2015, 11, 1018–1030. [Google Scholar] [CrossRef] [Green Version]
  51. Hyodo, M.; Sato, Y.; Hayakawa, Y. Synthesis of cyclic bis(3′-5′)diguanylic acid (c-di-GMP) analogs. Tetrahedron 2006, 62, 3089–3094. [Google Scholar] [CrossRef]
  52. Lioux, T.; Mauny, M.A.; Lamoureux, A.; Bascoul, N.; Hays, M.; Vernejoul, F.; Baudru, A.S.; Boularan, C.; Lopes-Vicente, J.; Qushair, G.; et al. Design, Synthesis, and Biological Evaluation of Novel Cyclic Adenosine–Inosine Monophosphate (cAIMP) Analogs That Activate Stimulator of Interferon Genes (STING). J. Med. Chem. 2016, 59, 10253–10267. [Google Scholar] [CrossRef] [PubMed]
  53. Shiraishi, K.; Saito-Tarashima, N.; Igata, Y.; Murakami, K.; Okamoto, Y.; Miyake, Y.; Furukawa, K.; Minakawa, N. Synthesis and evaluation of c-di-4′-thioAMP as an artificial ligand for c-di-AMP riboswitch. Bioorg. Med. Chem. 2017, 25, 3883–3889. [Google Scholar] [CrossRef] [PubMed]
  54. Gaffney, B.L.; Jones, R.A. Synthesis of c-di-GMP Analogs with Thiourea, Urea, Carbodiimide, and Guanidinium Linkages. Org. Lett. 2013, 16, 158–161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Dialer, C.R.; Stazzoni, S.; Drexler, D.J.; Muller, F.M.; Veth, S.; Pichler, A.; Okamura, H.; Witte, G.; Hopfner, K.P.; Carell, T. A Click-Chemistry Linked 2′3′-cGAMP Analogue. Chem. Eur. J. 2019, 25, 2089–2095. [Google Scholar] [CrossRef] [Green Version]
  56. Fujino, T.; Okuda, K.; Isobe, H. Conformational restriction of cyclic dinucleotides with triazole-linked cyclophane analogues. Tetrahedron Lett. 2014, 55, 2659–2661. [Google Scholar] [CrossRef]
  57. Shu, C.; Yi, G.; Watts, T.; Kao, C.C.; Li, P. Structure of STING bound to cyclic di-GMP reveals the mechanism of cyclic dinucleotide recognition by the immune system. Nat. Struct. Mol. Biol. 2012, 19, 722–724. [Google Scholar] [CrossRef] [Green Version]
  58. Ikeda, K.; Yanase, Y.; Hayashi, K.; Hara-Kudo, Y.; Tsuji, G.; Demizu, Y. Amine skeleton-based c-di-GMP derivatives as biofilm formation inhibitors. Bioorg. Med. Chem. Lett. 2021, 32, 127713. [Google Scholar] [CrossRef]
  59. Chen, L.; Zhao, S.; Zhu, Y.; Liu, Y.; Li, H.; Zhao, Q. Molecular Dynamics Simulations Reveal the Modulated Mechanism of STING Conformation. Interdiscip. Sci. 2021, 13, 751–765. [Google Scholar] [CrossRef]
  60. Novotnà, B.; Vaneková, L.; Zavřel, M.; Buděšínský, M.; Dejmek, M.; Smola, M.; Gutten, O.; Tehrani, Z.A.; Polidarová, M.P.; Brázdová, A.; et al. Enzymatic Preparation of 2′-5′,3′-5′-Cyclic Dinucleotides, Their Binding Properties to Stimulator of Interferon Genes Adaptor Protein, and Structure/Activity Correlations. J. Med. Chem. 2019, 62, 10676–10690. [Google Scholar] [CrossRef]
  61. Woodward, J.J.; Iavarone, A.T.; Portnoy, D.A. c-di-AMP Secreted by Intracellular Listeria monocytogenes Activates a Host Type I Interferon Response. Science 2010, 328, 1703–1705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Gentili, M.; Kowal, J.; Tkach, M.; Satoh, T.; Lahaye, X.; Conrad, C.; Boyron, M.; Lombard, B.; Durand, S.; Kroemer, G.; et al. Transmission of innate immune signaling by packaging of cGAMP in viral particles. Science 2015, 349, 1232–1236. [Google Scholar] [CrossRef] [PubMed]
  63. Case, D.A.; Babin, V.; Berryman, J.T.; Betz, R.M.; Cai, Q.; Cerutti, D.S.; Cheatham, T.E., III; Darden, T.A.; Duke, R.E.; Gohlke, H.; et al. AMBER 14 Referemce Manual; University of California: San Francisco, CA, USA, 2014. [Google Scholar]
Figure 1. Chemical structures of representative CDNs activating STING.
Figure 1. Chemical structures of representative CDNs activating STING.
Ijms 23 06847 g001
Scheme 1. Synthesis of the cyclic CDN derivative. (a) (1) CF3CO2Et, CH3CN, H2O; (2) (Boc)2O, Et3N, THF; (3) NH4OH; (b) 2-NsCl, Et3N, CH2Cl2, 68% for four steps; (c) 1,2-dibromoethane, K2CO3, DMF, 63%; (d) thiophenol, K2CO3, DMF, 82%; (e) compound 36, HATU, DIPEA, DMF; (f) (1) 7 M NH3 in MeOH; (2) 4 M HCl in 1,4-dioxane; (3) RP-HPLC purification, 15% for three steps. Abbreviations: Boc; tert-butoxycarbonyl, Ns; 2-nitrobenzenesulfonyl, DMF; N,N-dimethylformamide, HATU; 1-[Bis(dimethylamino)methylene]-1H-1,2,3-triazolo[4,5-b]pyridinium 3-oxide hexafluorophosphate, DIPEA; N,N-dimethylformamide, RP; reverse phase.
Scheme 1. Synthesis of the cyclic CDN derivative. (a) (1) CF3CO2Et, CH3CN, H2O; (2) (Boc)2O, Et3N, THF; (3) NH4OH; (b) 2-NsCl, Et3N, CH2Cl2, 68% for four steps; (c) 1,2-dibromoethane, K2CO3, DMF, 63%; (d) thiophenol, K2CO3, DMF, 82%; (e) compound 36, HATU, DIPEA, DMF; (f) (1) 7 M NH3 in MeOH; (2) 4 M HCl in 1,4-dioxane; (3) RP-HPLC purification, 15% for three steps. Abbreviations: Boc; tert-butoxycarbonyl, Ns; 2-nitrobenzenesulfonyl, DMF; N,N-dimethylformamide, HATU; 1-[Bis(dimethylamino)methylene]-1H-1,2,3-triazolo[4,5-b]pyridinium 3-oxide hexafluorophosphate, DIPEA; N,N-dimethylformamide, RP; reverse phase.
Ijms 23 06847 sch001
Scheme 2. Synthesis of the linear CDN derivative. (a) compound 36, HATU, DIPEA, DMF, 37% for four steps; (b) 4 M HCl in 1,4-dioxane, quant.; (c) (1) acetyl chloride, DIPEA, CH2Cl2; (2) 7 M NH3 in MeOH; (3) reprecipitation purification, 46% for two steps.
Scheme 2. Synthesis of the linear CDN derivative. (a) compound 36, HATU, DIPEA, DMF, 37% for four steps; (b) 4 M HCl in 1,4-dioxane, quant.; (c) (1) acetyl chloride, DIPEA, CH2Cl2; (2) 7 M NH3 in MeOH; (3) reprecipitation purification, 46% for two steps.
Ijms 23 06847 sch002
Figure 2. Evaluation of the IFN-β induction activity of compounds 7 and 12 with digitonin permeabilization. HEK293T cells were transfected with vectors encoding STING, as indicated, together with an IFN-luciferase reporter, and luciferase activity was measured 6 h after stimulation. Data are presented as mean ± SEM (n = 3). * p < 0.05; ** p < 0.005; and ns mean statistically no significance when compared with the DMSO-treated control in a Student’s t-test. Abbreviations: IFN; interferon, DMSO; dimethyl sulfoxide.
Figure 2. Evaluation of the IFN-β induction activity of compounds 7 and 12 with digitonin permeabilization. HEK293T cells were transfected with vectors encoding STING, as indicated, together with an IFN-luciferase reporter, and luciferase activity was measured 6 h after stimulation. Data are presented as mean ± SEM (n = 3). * p < 0.05; ** p < 0.005; and ns mean statistically no significance when compared with the DMSO-treated control in a Student’s t-test. Abbreviations: IFN; interferon, DMSO; dimethyl sulfoxide.
Ijms 23 06847 g002
Figure 3. Evaluation of the IFN-β inhibition activity of compounds 7 and 12 with digitonin permeabilization. HEK293T cells were transfected with vectors encoding STING, as indicated, together with an IFN-luciferase reporter, and luciferase activity was measured 6 h after stimulation. Data are presented as mean ± SEM (n = 3). * p < 0.05; ** p < 0.005; and ns mean statistically no significance when compared with the cGAMP-treated control in a Student’s t-test.
Figure 3. Evaluation of the IFN-β inhibition activity of compounds 7 and 12 with digitonin permeabilization. HEK293T cells were transfected with vectors encoding STING, as indicated, together with an IFN-luciferase reporter, and luciferase activity was measured 6 h after stimulation. Data are presented as mean ± SEM (n = 3). * p < 0.05; ** p < 0.005; and ns mean statistically no significance when compared with the cGAMP-treated control in a Student’s t-test.
Ijms 23 06847 g003
Figure 4. Docking study of 7 and 12 within the ligand binding site of STING (PDB ID: 4LOH for 7, 6MXE for 12) using MOE. (A) The 3D-structure of superposed docked conformation of 7 (green) with the natural ligand, cGAMP (orange). (B) The 3D-structure of superposed docked conformation of a single molecule of 12 (green) with a reference inhibitor, cpd. 18 (orange). (C) The 3D-structure of superposed docked conformations of two molecules of 12 (green) with inhibitor 18 (orange).
Figure 4. Docking study of 7 and 12 within the ligand binding site of STING (PDB ID: 4LOH for 7, 6MXE for 12) using MOE. (A) The 3D-structure of superposed docked conformation of 7 (green) with the natural ligand, cGAMP (orange). (B) The 3D-structure of superposed docked conformation of a single molecule of 12 (green) with a reference inhibitor, cpd. 18 (orange). (C) The 3D-structure of superposed docked conformations of two molecules of 12 (green) with inhibitor 18 (orange).
Ijms 23 06847 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yanase, Y.; Tsuji, G.; Nakamura, M.; Shibata, N.; Demizu, Y. Control of STING Agonistic/Antagonistic Activity Using Amine-Skeleton-Based c-di-GMP Analogues. Int. J. Mol. Sci. 2022, 23, 6847. https://doi.org/10.3390/ijms23126847

AMA Style

Yanase Y, Tsuji G, Nakamura M, Shibata N, Demizu Y. Control of STING Agonistic/Antagonistic Activity Using Amine-Skeleton-Based c-di-GMP Analogues. International Journal of Molecular Sciences. 2022; 23(12):6847. https://doi.org/10.3390/ijms23126847

Chicago/Turabian Style

Yanase, Yuta, Genichiro Tsuji, Miki Nakamura, Norihito Shibata, and Yosuke Demizu. 2022. "Control of STING Agonistic/Antagonistic Activity Using Amine-Skeleton-Based c-di-GMP Analogues" International Journal of Molecular Sciences 23, no. 12: 6847. https://doi.org/10.3390/ijms23126847

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop