Next Article in Journal
The Ferritin, Hepcidin and Cytokines Link in the Diagnoses of Iron Deficiency Anaemia during Pregnancy: A Review
Previous Article in Journal
Relapse/Refractory Paediatric B-ALL Case with CD19 Phenotype Switching Indicating the Importance of Appropriate Diagnostic Approach and Targeted Treatment Adjustment—Case Report
Previous Article in Special Issue
Immunomodulatory Role of Interferons in Viral and Bacterial Infections
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Significance of the cGAS-STING Pathway in Health and Disease

1
Fujian Key Laboratory of Innate Immune Biology, Biomedical Research Center of South China, College of Life Science, Fujian Normal University Qishan Campus, Fuzhou 350117, China
2
Key Laboratory of College of First Clinical Medicine, College of First Clinical Medicine, Fujian Medical University, Taijiang Campus, Fuzhou 350001, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(17), 13316; https://doi.org/10.3390/ijms241713316
Submission received: 30 June 2023 / Revised: 18 August 2023 / Accepted: 22 August 2023 / Published: 28 August 2023

Abstract

:
The cyclic GMP-AMP synthase (cGAS)-stimulator of interferon genes (STING) pathway plays a significant role in health and disease. In this pathway, cGAS, one of the major cytosolic DNA sensors in mammalian cells, regulates innate immunity and the STING-dependent production of pro-inflammatory cytokines, including type-I interferon. Moreover, the cGAS–STING pathway is integral to other cellular processes, such as cell death, cell senescence, and autophagy. Activation of the cGAS–STING pathway by “self” DNA is also attributed to various infectious diseases and autoimmune or inflammatory conditions. In addition, the cGAS–STING pathway activation functions as a link between innate and adaptive immunity, leading to the inhibition or facilitation of tumorigenesis; therefore, research targeting this pathway can provide novel clues for clinical applications to treat infectious, inflammatory, and autoimmune diseases and even cancer. In this review, we focus on the cGAS–STING pathway and its corresponding cellular and molecular mechanisms in health and disease.

1. Introduction

The innate immune system, which constitutes part of the defense system of mammals, can recognize the “non-self” genetic material of invading pathogens. In particular, the germline-encoded pattern-recognition receptors (PRRs) that are usually located on the cell surface or in the cellular compartments of the cytosol distinguish the pathogen-associated molecular patterns (PAMPs) of extracellular pathogens [1]. Furthermore, “self” DNA molecules, which are vital genetic material found within specific regions of cells, can occasionally occur as misplaced DNA within the cytosol. This DNA is also recognized by PRRs via damage-associated molecular patterns (DAMPs) [2]. Consequently, the recognition of both PAMPs and DAMPs by PRRs elicits the innate defense mechanism, as well as the production of soluble mediators, such as type-I interferon (IFN) and other cytokines, to impede viral or microbial infection and maintain cellular balance [3]; thus, the production of type-I IFN is fundamentally controlled by PRRs. Among the different PRRs, cyclic GMP-AMP synthase (cGAS) is a class of PRRs involved in recognizing cytosolic DNA. A stimulator of interferon genes (STING), which is localized on the endoplasmic reticulum (ER), was initially discovered as the upstream initiator of type-I IFN [4]; however, STING cannot directly bind to immune-stimulatory DNA (ISD). Instead, bacteria-derived cyclic diguanylate monophosphate (c-dGMP) or cyclic di-adenosine monophosphate (c-dAMP) serve as the upstream ligands of STING [5,6]. Subsequently, cGAMP (cGMP-AMP), which is synthesized by cGAS, acts as the novel second messenger of STING [7].
The cGAS–STING pathway is widely expressed in immune, non-immune, and cancer cells [8]. In addition to triggering the type-I IFN signaling cascade to mediate cellular innate immune responses, the cGAS–STING pathway is also involved in other intrinsic cellular processes, including cell death, autophagy, and cellular senescence [9]. Moreover, the cGAS–STING pathway is regulated by other DNA-sensing pathways and cellular molecules to maintain intracellular homeostasis under normal conditions [9].
The primary role of the cGAS–STING pathway is to activate host innate immunity against a DNA-containing pathogen infection [10]; however, abnormal activation of the cGAS–STING pathway can lead to the excessive and continuous production of type-I IFN, resulting in its disproportionate accumulation in tissues and organs [11]. Moreover, increasing evidence has suggested that the abnormal accumulation of type-I IFN is involved in the pathogenesis of autoimmune diseases and mediates inflammation development in other diseases [12,13]. Additionally, activating the cGAS–STING pathway can restrict tumor growth by inducing specific antitumor immunity [11]; therefore, preclinical research targeting the cGAS–STING pathway can pave the way for future translation into clinical trials, ultimately alleviating symptoms, improving the life quality of patients with various autoimmune or inflammatory disorders, and prolonging the life expectancy of patients with cancer.
In this review, we describe the mechanisms by which the cGAS–STING pathway induces type-I IFN immune responses and the regulation of the cGAS–STING pathway by other DNA-sensing pathways and cellular molecules. We further cover the role of the cGAS–STING pathway in certain cellular processes, such as cell death, cell senescence, and autophagy. Additionally, we highlight the importance of the cGAS–STING pathway in anti-pathogen immunity and summarize the advances in the current research targeting this pathway in various disorders and cancers. Finally, we discuss the existing and potential therapeutic strategies focused on the cGAS–STING pathway and provide recommendations that should be addressed in future studies.

2. Non-Canonical Activation of cGAS and cGAMP Synthesis

cGAS is not only localized in the cytosol, but has also been detected in nuclear regions [7,14]. cGAS is a 60 kDa protein with a non-conserved amino-terminal stretch composed of approximately 130–150 residues, along with a highly conserved Mab21 domain belonging to the nucleotidyl transferase (NTase) superfamily and three other major domains, including an unstructured active domain, a two-lobed inactive catalytic domain, and an extended N-terminal domain [15].
cGAS functions as a cytosolic DNA sensor; thus, its activation is determined by its recognition of cytosolic double-stranded DNA (dsDNA). Moreover, cGAS can be activated by dsDNA from multiple sources, such as microbial DNA, mitochondrial DNA (mtDNA), nuclear chromatin, extracellular self-DNA, cytosolic chromatin and micronuclei, and dysfunctional telomeres [3]. The dsDNA from microbes, including simplex virus, vaccinia virus, cytomegalovirus, Chlamydia trachomatis, Mycobacterium tuberculosis, and Francisella novicida, can enter the cytosol and be sensed by cGAS, leading to a downstream signaling cascade that induces the type-I IFN response [10]. Mitochondrial stress can be induced by the deficiency of mitochondrial transcription factor A (TFAM; an mtDNA-binding protein) or the release of mtDNA into the cytosol by herpes viruses, resulting in cytosolic mtDNA that subsequently engages cGAS [16]. Extracellular self-DNA can arise from the deficiencies of three-prime repair exonuclease 1 (TREX1; an exonuclease) or lysosomal DNases and then invade the cytosol of host cells, thereby activating cGAS [17]. Cytosolic micronuclei mainly arise from DNA double-strand breaks (DSBs), mitotic errors, and abnormal DNA replication, which result in acentric chromosome fragments and whole non-segregated chromatins that are generated due to their failure to be included in the daughter nuclei after telophase [18]. Eventually, these chromatins and chromatin fragments recruit their nuclear envelopes (NEs) and form micronuclei in the cytosol. The collapse of the NE of these micronuclei then leads to the recognition of the cytosolic chromatin DNA by cGAS, causing its activation [19]. DNA from dysfunctional telomeres is another potential source of cytosolic DNA. Telomere shortening is a hallmark of cellular senescence, which induces replicative crises, such as mitotic delay and deformation of chromosomes (e.g., chromosomes with two centromeres). This abnormal chromosomal DNA reaches the cytosol and is detected by cGAS [20]. Furthermore, only dsDNA over 40 bp in length, such as those previously mentioned, can theoretically be detected by cGAS; however, a unique structure called “stem-loop” formed during the reverse transcription of human immunodeficiency virus type 1 (HIV-1) has been reported to activate cGAS in a sequence-dependent manner due to the presence of unpaired guanosines flanking short dsDNA (Y-DNA), which are considered the cGAS recognition motif and enhancer of cGAS enzymatic activity [21]. One report indicated that cGAS is predominantly localized in the nuclear region, which is tethered tightly by a salt-resistant interaction in order to maintain the resting state of cGAS and prevent auto-reactivity [14]. In addition, another study illustrated that in the nucleus, the barrier-to-autointegration factor 1(BAF1) competitively binds to self-DNA for prohibition of the formation of DNA–cGAS complexes that are essential for the enzymatic activity of cGAS [22].
After recognizing these DNA stimulants in mammalian cells, each of the two DNA-binding sites on the catalytic domains of cGAS binds with different dsDNA, generating a 2:2 cGAS–dsDNA complex oriented in two different directions that represents the minimal active enzymatic unit [13]. Additionally, in the human cGAS (h-cGAS), the 2:2 cGAS–dsDNA complexes rearrange themselves to create a ladder-like network that enables the successive recruitment of adjacent cGAS to form more 2:2 cGAS–dsDNA and stabilize its structure [23,24]. Next, the catalytic domain of the cGAS–dsDNA complex is rearranged to catalyze its substrates (guanosine triphosphate [GTP] and adenosine triphosphate [ATP]) into cGAMP, which contains two phosphate diester bonds that have a high affinity for STING [25,26,27,28]. Furthermore, the N-terminal of cGAS mainly induces the liquid phase transition of cGAS to form liquid-like droplets with dsDNA that promote cGAMP production by enhancing the concentration of enzymatic activity and reactants [29,30]. This liquid phase transition requires an appropriate concentration of cGAS and dsDNA because cGAS is only activated when dsDNA reaches a certain concentration [26].

3. Downstream Signaling Reaction of cGAMP

The STING protein comprises four major components, including a short cytosolic N-terminal segment, a four-span transmembrane domain, a connector region, and a cytosolic ligand-binding domain (LBD) with a C-terminal tail (CCT) [13].
cGAMP can be sensed by the ER-located STING [4,31]. After detecting cGAMP, STING undergoes dimerization, followed by the LBD of the STING dimer rotating in the opposite direction to form an ordered β-sheet [32]. Subsequently, the STING dimer forms a “cryo-EM” structure, leading to a structural conformational change that causes self-activation [32,33]. After activating itself, the trafficking STING dimer translocates from the ER to the ER–Golgi intermediate compartment (ERGIC) with the collaboration of the TRAP–translocon complex [34]. Simultaneously, STING recruits TANK-binding kinase 1 (TBK1) for a downstream signaling cascade during this process [13]. Next, cytoplasmic coat protein complex-II (COPII) and translocon-associated protein β (TRAPβ) facilitate STING to reach the Golgi [35]. In the Golgi, the Ser366 site in the CCT of trafficking STING is crosswise phosphorylated by an adjacent TBK1 associated with another STING, while the TBK1 molecules phosphorylate each other [36]. This modification enables the highly efficient activation and palmitoylation of STING, and STING and TBK1 aggregate to form the STING signalosome [37]. Subsequently, STING and TBK1 rearrange to a “scaffold” structure, leading to the phosphorylation of interferon regulatory factor 3 (IRF3). Finally, the phosphorylated IRF3 undergoes dimerization and is trafficked into the nucleus as a transcriptional factor that binds the promoter of type-I IFN, causing its expression [38]. In addition, the STING–TBK1 signalosome can phosphorylate the inhibitors of the transcription factor NF-κB (IκBα), resulting in the polyubiquitination and degradation of IκBα by the ubiquitin–proteasome pathway to release NF-κB into the nucleus [4]. The NF-κB has two main roles in the nucleus. One role involves aiding the expression of type-I IFN, and the other consists of regulating the transcription of inflammatory cytokines, such as interleukin 6 (IL-6) and tumor necrosis factor (TNF) [38,39]. In summary, the non-canonical activation of the cGAS–STING pathway involves the induction of pro-inflammatory cytokine expression, particularly type-I IFN. Activation of the cGAS-STING pathway during innate immune response was showed in the Figure 1.

4. Regulation of the cGAS–STING Pathway

The cGAS–STING pathway is positively and negatively regulated by cellular molecules, enzymes, and other DNA-sensing pathways.
In the negative regulation of the cGAS–STING pathway, the most efficient mechanism is to eliminate cytosolic DNA and thereby prevent cGAS activation. Several cellular nucleases can perform this role. One such nuclease is DNase II, a lysosomal enzyme responsible for DNA digestion in endosomes or autophagosomes to prevent DNA from leaking into the cytosol [17]. Similarly, the main function of TREX1 (also called DNase III) is to eliminate free dsDNA entering the cytosol [40]. Additionally, the RNaseH2 endonuclease complex (RNaseH2A, RNaseH2B, and RNaseH2C) is responsible for separating ribonucleotides embedded in the DNA of RNA–DNA hybrids, which ensures that DNA replication occurs on the rails and restricts the release of abnormally replicated DNA into the cytosol [41]. Another nuclease is SAM domain- and HD domain-containing protein 1 (SAMHD1), which is widely characterized as a dNTPase that acts as an inhibitor to limit the reverse transcription of RNA viruses. Moreover, activated SAMHD1 can increase the enzymatic activity of MRE11 to degrade nascent ssDNA at stalled replication forks, which impedes the accumulation of ssDNA fragments in the cytosol [42]. Consequently, a deficiency in any endonucleases, particularly DNase II, TREX1, the RNaseH2 endonuclease complex, and SAMHD1, may lead to the abnormal accumulation of dsDNA, as well as excessive activation of the cGAS–STING pathway [9].
Apart from the previously mentioned endonucleases, other DNA sensors can also negatively modulate the cGAS–STING pathway. γ-interferon-inducible protein-16 (IFI16) is a typical absent melanoma 2 (AIM2)-like receptor from the PYHIN family, which can sense an invading DNA virus or damaged chromatin DNA in both the nucleus and cytosol. After DNA recognition, IFI16 initiates the STING-dependent type-I IFN signaling cascade with TNF receptor-associated factor 6 (TRAF6) and p53 [43]. Studies have also indicated that IFI16 competitively binds with DNA and adaptor STING, partially suppressing cGAMP production; yhus, IFI16 is essential for the higher-level responses of STING [44,45]. Furthermore, the HIN2 domain of IFI16 inhibits the activation of cGAS, while a portion of cGAS can enter the nucleus and maintain the stability of IFI16 [46,47]. AIM2 is another DNA sensor in the PYHIN family located in the cytosol. After recognizing dsDNA, AIM2 associates with apoptosis-associated speck-like proteins, such as a CARD (ASC) and procaspase-1, to form inflammasomes [48]. The inflammasomes lead to a multi-protein signaling cascade that activates caspase-1 and further facilitates the transformation of pro-inflammatory cytokines, such as IL-1β and IL-18, into active forms, ultimately inducing pyroptosis [49]. Simultaneously, the AIM2-induced inflammasomes interfere with the activation of the cGAS–STING pathway via the cleavage of cGAS and reduce cell viability [50,51]. Furthermore, the activated caspase-1 cleaves gasdermin D and promotes its pore-forming activity, resulting in an intracellular potassium (K+) efflux that prevents cytosolic DNA from engaging cGAS [52]. Alternatively, activating the cGAS–STING pathway may dampen other DNA sensor-induced type-I IFN production. Toll-like receptor 9 (TLR9) is an important endosomal PRR composed of a leucine-rich repeat (LRR) domain and a cytosolic c-terminal Toll/IL-1 receptor (TIR) domain [53]. During infection, TLR9 recognizes cytosine–phosphate–guanosine (CpG)-rich sequences formed from bacterial or viral DNA via its LRR domain, which leads to the recruitment of MyD88 to the cytosolic C-terminal TIR domain for the nuclear translocation of IFN transcription factor 7 (IRF7), eventually enhancing type-I IFN production [54]. In plasmacytoid dendritic cells (pDCs), activation of the cGAS–STING pathway triggers the expression of suppressor of cytokine signaling 1 (SOCS1) and SOCS3, which subsequently inhibits the TLR9-mediated type-I IFN signaling cascade [55].
The downstream ligands of cGAS, including cGAMP and STING, can also negatively regulate cGAS activity through modifications such as deubiquitylation, glutamylation, phosphodiesterase-catalyzed hydrolysis, sumoylation, and phosphorylation. In addition, cGAS activity can be modulated by direct protein–protein interactions, while intracellular molecules can also modulate cGAS activity. Table 1 summarizes the cellular molecules that can suppress the activity of cGAS, cGAMP, or STING by modification, direct protein binding, or degradation.
The positive regulation of the cGAS–STING pathway is mainly achieved by intracellular modulators during the post-translational stages via modifications, direct interactions, or indirect assistance. These modulators are summarized in Table 2.

5. cGAS–STING Pathway in Autophagy

Autophagy is an intracellular, self-protective mechanism to maintain energy balance in response to nutrient stress in cells. This process can degrade misfolded or aggregated proteins, damaged organelles, and invading pathogens [82]. Studies have shown that under conditions of pathogen infection, the cGAS–STING pathway can induce autophagy without triggering type-I IFN or NF-κB signaling [83]. Specifically, in mammalian cells, cGAMP-bound STING translocates to the ERGIC independently of the recruitment of TBK1. The STING-containing ERGIC then acts as the membrane source of LC3B (microtubule-associated proteins 1A/1B light chain 3B) lipidation to form autophagosomes that engulf cytosolic DNA, subcellular organelles, or misfolded proteins with the collaboration of WD repeat domain phosphoinositide-interacting protein 2 (WIPI2) and autophagy protein 5 (ATG5) [83]. In contrast, the formation of canonical autophagosomes requires the recruitment of the unc-51-like kinase 1 (ULK1) complex that phosphorylates components of the class III PI3K (PI3KC3) complex I and the negative modulation of the mammalian target of rapamycin (mTOR); therefore, the cGAS–STING pathway-induced autophagy is a novel bypass mechanism to initiate autophagy. Moreover, in certain ancient invertebrate species, such as the anemone Nematostella vectensis that emerged 500 million years before human beings, the cGAS–STING pathway-induced autophagy plays an essential role in resisting microbe infection [84]. In the case of mammals, the cGAS–STING pathway-induced autophagy may be mainly aimed at protecting against the invasion of extracellular pathogens, including M. tuberculosis, certain Gram-positive bacteria, and the Zika virus; however, its excessive signaling may cause irreversible apoptosis during cellular replicative stress [20,84,85,86,87]. Nevertheless, the cGAS–STING pathway-dependent autophagy is generally conserved and can reduce the over-reaction of the cGAS–STING pathway-induced type-I IFN signaling cascade.

6. cGAS–STING Pathway in Cell Death

Another functional response mediated by the cGAS–STING pathway is triggering cell death. The cGAS–STING pathway mainly initiates three types of cell death pathways (i.e., lysosomal-dependent pyroptosis [LDCP], apoptosis, and necroptosis) to eliminate infected, damaged, or transformed cells and maintain organismal homeostasis.
The hallmark of LDCP is the release of lysosomal hydrolases, such as cathepsins, into the cytosol due to lysosomal membrane permeabilization and the ensuing pyroptosis [88]. In specific cell types, particularly human myeloid cells, STING is sorted into lysosomes after activating the downstream signaling cascade, causing permeabilization of the lysosome membrane and consequent leakage of cathepsins into the cytosol. Next, the NLRP3 inflammasome and caspase-1 are activated, ultimately resulting in a K+ efflux and pyroptosis. Pyroptosis has also been suggested to be a secondary inflammatory response in dying cells involving the secretion of cytokines, such as IL-1β and pro-IL-1β [89].
Apoptosis is a programmed cell death process that activates neighboring cells or macrophages to eliminate damaged cells [90]. The initiation of apoptosis relies on activating any three signals: extrinsic signals, intrinsic mitochondrial signals, and granzyme-mediated signals [91]. The intrinsic mitochondrial signals, including cellular stresses such as genotoxic stress, activate two proapoptotic proteins, BCL-2-associated X protein (BAX) and BCL-2 homologous killer (BAK), by stimulating only BH3 domain family members (BIM). These activated proapoptotic proteins cooperatively form a pore-like conformation, inducing mitochondrial outer membrane permeabilization, which in turn results in the release of mitochondrial contents, including cytochrome c and mtDNA [92]. Cytochrome c then binds with an adapter apoptotic protease activating factor-1 (APAF1) to form a cytochrome c-APAF1 association. Subsequently, this association recruits and activates caspase-9 to form an apoptosome that triggers apoptosis [93]. Additionally, the activated caspase-9 activates caspase-7 and -3, which results in the inhibition of the cGAS–STING pathway due to the cleavage of cGAS and IRF3 by the activated caspase-7 and -3 [94]. Although the cGAS–STING pathway is inhibited during apoptosis, independent studies have confirmed that certain mechanisms ensure that moderate cGAS–STING-mediated type-I IFN production is triggered by mtDNA in response to pathogen infection [95,96]. Considering the findings above, the function of caspase-9-mediated apoptosis may involve avoiding excessive type-I IFN immune responses. Another study has reported that the infection of Mycobacterium bovis in murine macrophages can lead to the release of activated STING into the cytosol due to ER stress [97]. In this context, STING can directly recruit phosphorylated TBK1, which then activates IRF3 that can directly trigger apoptosis in a BAX–BAK-dependent manner [97,98]; thus, the cGAS–STING pathway-induced activation of IRF3 results in apoptosis rather than a type-I IFN signaling cascade under viral infection or irreversible damage. Furthermore, excessive activation of autophagy mediated by cGAS–STING may induce the irreversible apoptosis process termed autophagy-dependent apoptosis [87].
Necroptosis is a type of cellular necrosis that serves as a supplementary mechanism for regulating cell death to ensure normal development and homeostasis of multicellular organisms [99]. Excessive production of TNF and type-I IFN has been shown to contribute to necroptosis initiation when apoptosis is moderated by various conditions, including genetic defects, caspase-9 inhibition, and pathogen invasion [100]. Another study illustrated that in bone marrow-derived macrophages (BMDMs), type-I IFN and TNF produced in a cGAS–STING pathway-dependent manner are maintained at a critical threshold level after the recognition of cytosolic DNA. This threshold maintenance facilitates the subsequent activation of mixed lineage kinase domain-like protein (MLKL) or receptor-interacting serine/threonine protein kinase 3 (RIPK3) that causes plasma membrane disruption and eventually enables cells to undergo necroptosis, during which caspase-9 is inhibited [101,102].

7. cGAS–STING Pathway in Cellular Senescence

The cell cycle is vital in regulating cell proliferation, growth, and division. Cellular senescence is characterized by the permanent arrest of the cell cycle via telomere erosion or various stress, generally comprising ionizing radiation, oxidative stress, or oncogene signaling [103]. Certain hallmarks can help distinguish senescent cells from normal cells, including an enlarged size, increased granularity, increased senescence-associated (SA)-β-galactosidase activity, an increased level of cyclin-dependent kinase inhibitors, the expression of anti-proliferative molecules (p16 and p21), and DNA damage foci [104]. A report using mouse embryonic fibroblasts (MEFs) demonstrated that when cells undergo senescence by ionizing radiation, changes in the nuclear structure lead to the disintegration of the nuclear lamina (NL), which is responsible for maintaining the NE structure. Furthermore, the chromatin fragments generated by senescence-related DNA damage enter the cytosol, activating cGAS [105]. Interestingly, senescent cells have been indicated to contain higher levels of chromatin fragments than normal cells, and the depletion of cGAS can result in the spontaneous immortalization of senescent cells [105,106]. Senescent cells express various SA secretory phenotype (SASP)-related genes, including those of cytokines, chemokines, and proteases, in an autocrine and paracrine manner, which reinforces cell cycle arrest and mediates chronic inflammation [107]. Studies have further shown that although the expression of SASP-related genes in senescent cells is controlled at multiple levels, the cGAS–STING pathway-mediated expression of several genes is related to the regulation of the SASP-related genes. The cGAS–STING pathway recognizes cytosolic chromatin fragments in senescent cells and mediates the expression of cytokines (IL-6 and TNF-α) and chemokines (Cxcl-10, Cxcl-2, Ccl-3, and Ccl-5) to shape the microenvironment of chronic inflammation in senescent cells [108].
Hematopoietic stem cells (HSCs) are usually dormant in the bone marrow niche. A study revealed that HSCs trigger the type-I IFN immune response via the cGAS–STING pathway during the invasion of DNA-containing bacteria in the bone marrow. The stimulation of the type-I IFN immune response, in turn, leads to the exhaustion of HSCs [109]. Moreover, the excessive secretion of type-I IFN in response to DNA damage can accelerate cellular senescence and inhibit cellular function via the activation of p53–p21 signals and increased p16INK4 levels in HSCs [110]. Interestingly, another investigation highlighted that long-term HSCs produce a peculiar circular RNA termed “cia-cGAS” under homeostatic conditions in the nucleus, which exhibits a higher binding affinity for cGAS than self-DNA. Furthermore, cia-cGAS was found to suppress cGAS–DNA binding, resulting in the inhibition of cGAS-mediated type-I IFN production and the maintenance of a dormant state [60]. All these findings indicate that the cGAS–STING pathway has a predominant role in facilitating cellular senescence. The cGAS–STING pathway in cell death, autophagy, and senescence was illustrated in Figure 2.

8. cGAS–STING Pathway in Anti-Pathogen Immunity

Theoretically, all nucleic acids, such as self-DNA and DNA–RNA hybrids, can be bound by cGAS; therefore, the cGAS–STING pathway is a major innate immune pathway triggered against infections by numerous diverse pathogens, such as DNA viruses, retroviruses, DNA-containing bacteria, and parasites. cGAMP has also been reported as an adjuvant that boosts adaptive immune responses by accelerating antigen-specific T-cell activation and antibody production [10,111].
Viral invasion, such as that by the herpes simplex virus (HSV), elicits mtDNA stress, causing detrimental effects on mtDNA stability and eventually leading to mtDNA entry into the cytosol. The resulting cytosolic mtDNA promotes the production of type-I IFN and pro-inflammatory cytokines via the cGAS–STING pathway [16]. Moreover, an infection of the Dengue virus and SARS-CoV-2 also causes the activation of the cGAS–STING pathway through cytosolic mtDNA rather than its genome [112,113]. In the case of retroviruses (including the typical HIV-1 retrovirus), cGAS is recruited to the capsid of HIV-1 through binding of polyglutamine-binding protein 1 (PQBP1), which is the adaptor of cGAS, and then induces a STING-dependent anti-viral immune response. [21,114,115,116]. Nevertheless, viruses have been reported to evolve corresponding proteins to inhibit the cGAS–STING pathway and escape immune surveillance. Table 3 displays these viral proteins from previous literature.
During bacterial invasion, including M. tuberculosis infection, cGAS in the macrophages detects M. tuberculosis genomic DNA in the cytosol. This recognition not only induces significantly higher levels of type-IFN, but also initiates the ubiquitin-mediated selective autophagy to degrade the bacilli [117]. Furthermore, after a Burkholderia pseudomallei or B. thailandensis invasion, cGAS can induce host cell fusion, forming multinucleated giant cells (MNGCs) via the bacterial type VI secretion system 5 (T6SS-5). This process occurs without interference from extracellular host defenses, facilitating the diffusion of these Burkholderia species among the host cells [118]. Furthermore, mitotic events in the MNGCs are abnormally disrupted, resulting in micronuclei formation. cGAS then recognizes the chromatin fragments in the micronuclei by colocalization and consecutively activates STING-dependent autophagy rather than a type-I IFN signaling cascade, ultimately causing autophagy-dependent apoptosis [119]. Interestingly, during the initial formation of these MNGCs, type-I IFN produced in the cGAS–STING pathway induces the expression of guanylate-binding proteins (GBPs), which reduces MNGC formation by inhibiting bacterial Arp2/3-dependent actin motility [120]; therefore, two cGAS–STING pathway mechanisms of genomic replication inhibition are involved in B. pseudomallei or B. thailandensis infection. Moreover, during F. novicida infection, cGAS together with IFI16 binds cytosolic Francisella DNA and induces the highest level of STING-dependent type-I IFN immune response [121]. Conversely, certain bacteria have evolved strategies to limit the activation of the cGAS–STING pathway. These bacteria and their mechanisms have been summarized in Table 4.
The cGAS–STING pathway is also vital in combating parasites. This role has been particularly highlighted in malaria infections caused by the introduction of parasites of the Plasmodium genus into the blood cells after mosquito bites. In the blood cells, the Plasmodium parasites create extracellular vesicles (EVs) to preserve their genomic DNA; however, monocytes absorb these EVs and cause the parasitic genomic DNA to leak into the cytosol of the host cells [122]. After detecting the genomic DNA of the Plasmodium species, cGAS induces type-I IFN production and facilitates the germinal center (GC)-mediated humoral immunity to eliminate the parasites [123]. Furthermore, when Toxoplasma gondii enters normal individuals, these parasites are dealt with by the immune system and rendered dormant in the brain. More specifically, after infection with T. gondii, the T. gondii genome undergoes rapid replication in the cytosol of cerebral cells, activating cGAS. The Toxoplasma gondii dense granule protein 15 (TgGRA15) then facilitates the ubiquitination of STING, further enhancing the production of STING-dependent type-I IFN and chemokines to suppress T. gondii growth and establish latent infection [124,125].
Table 3. Proteins of various viruses restricting the cGAS–STING pathway.
Table 3. Proteins of various viruses restricting the cGAS–STING pathway.
VirusProteinsTargetsMechanismsRefs.
Herpesviridae
HSV-1UL37
VP22
γ134.5
cGAS
cGAS
STING
Blocks the synthesis of cGAMP
Reduces the activity of cGAS
Disrupts STING trafficking
[126]
[127]
[128]
KSHVORF52
vIRF1
cGAS
STING
Blocks the synthesis of cGAMP
Obstructs STING trafficking
[129]
[130]
Coronaviridae
PEDVPLP2STINGImpedes polyubiquitination of STING[131]
SARS-CoVPLproSTINGDisrupts the dimerization of STING[132]
Flaviviridae
HCVNS4BSTINGBlocks the interaction between STING and TBK1[133]
DENVNS2BcGASInitiates autophagy to degrade cGAS[134]
ZIKVNS1cGASCleaves cGAS[135]
Papillomaviridae
HPVE7STINGDecreases the activity of STING[136]
Adenoviridae
ADEVE1ASTINGDecreases the activity of STING[136]
Hepadnaviridae
HBVPolSTINGImpedes polyubiquitination of STING[137]
Poxviridae
POXVPoxinscGAMPHydrolyzes cGAMP[138]
Retroviridae
HIVVpxSTINGRestrains STING-induced NF-κB signaling[139]
Abbreviations: HSV-1, herpes simplex virus 1; KSHV, Kaposi’s sarcoma-associated herpes virus; PEDV, porcine epidemic diarrhea virus; SARS-CoV, severe acute respiratory syndrome-coronavirus; HCV, hepatitis C virus; DENV, dengue virus; ZIKV, Zika virus; HPV, human papilloma virus; ADEV, adenovirus; HBV, hepatitis B virus; Pol, polymerase; POXV, poxvirus; HIV, human immunodeficiency virus; cGAS–STING, cyclic GMP-AMP synthase-stimulator of interferon genes; cGAMP, cGMP-AMP; TBK1, TANK-binding kinase 1.
Table 4. Proteins of different bacteria limiting the cGAS–STING pathway.
Table 4. Proteins of different bacteria limiting the cGAS–STING pathway.
BacteriaProteinsTargetsMechanismsRefs.
GBSCdnPCDNsHydrolyzes cyclic-di-AMP[140]
C. trachomatisCpoSSTINGPrevents STING-induced
cell death and type-I IFN production
[141]
YersiniaYopJSTINGRestrains the formation of
STING-TBK1 signalosome
[142]
M. tuberculosisCpsASTINGInhibits STING-dependent
autophagy
[143]
Abbreviations: GBS, Group B Streptococcus; CdnP, a type of ectonucleotidase; CDNs, cyclic dinucleotides; C. trachomatis, Chlamydia trachomatis; CpoS, Chlamydia promoter of survival inclusion membrane protein; YopJ, Yersinia outer protein J; M. tuberculosis, Mycobacterium tuberculosis; CpsA, a member of the LytR-CpsA-Psr (LCP) protein family; cGAS–STING, cyclic GMP-AMP synthase-stimulator of interferon genes; cGAMP, cGMP-AMP; TBK1, TANK-binding kinase 1.

9. cGAS–STING Pathway in Autoimmune Disorders and Inflammation

Although the morbidity factors of various autoimmune disorders remain unclear, the abnormal upregulation of type-I IFN has been linked to a spectrum of autoimmune disorders termed “type-I IFN interferonopathies” [144]; thus, the cGAS–STING pathway, which is primarily responsible for producing type-I IFN, is likely to be involved in these disorders.
One such autoimmune disorder is STING-associated vasculopathy with infantile onset (SAVI), which is characterized by ulcerating acral skin lesions, fever episodes, and lung fibrosis. Patients with SAVI present with IgM (Immunoglobulin M) deposition, variable autoantibody titers, and excessive type-I IFN production [145]. This disease is caused by the gain-of-function mutation in the STING-encoding gene that leads to the spontaneous translocation of STING from the ER to the GR in the absence of stimulation by cGAMP, resulting in persistent IRF3 phosphorylation and type-I IFN expression [145]. Another autoimmune disorder is COPA syndrome, which is an early-onset autosomal disease distinguished by arthritis and interstitial lung disease, along with pulmonary hemorrhage as a striking feature. Patients with COPA syndrome have increased levels of T helper cell 17 (Th17 cell) and abnormal expression of several cytokines, such as IL-1β, IL-6, and type-I IFN [146]. This syndrome is attributed to heterozygous mutations in a subunit of coat protein complex I (COPI), resulting in impaired binding and sorting of the proteins targeted for ER retrieval [147]. Consistent with this process, STING is abnormally accumulated in the ER, inducing the intensive type-I IFN signaling cascade [148]. Additionally, familial chilblain lupus (FCL) is a form of cutaneous lupus erythematosus with onset in childhood. Patients with FCL experience cold-induced bluish-red skin lesions in peripheral locations and manifest elevated levels of mucin formation and immunoglobulin deposits [149]. FCL is caused by the heterozygous gain-of-function mutation in STING, which enables STING to undergo dimerization without cGAMP interaction and constitutively activates the type-I IFN signature [150]. In addition to the previously mentioned autoimmune disorders, systemic lupus erythematosus (SLE) is a chronically autoimmune disorder with systemic manifestations, including fever, fatigue, multiple organ failure, and perturbations of the hematopoietic system [151]. Patients with SLE demonstrate deposition of the immune complex formed by anti-nuclear autoantibodies in tissues and abnormal T-cell activation; however, limited evidence has suggested that the cGAS–STING pathway has a central role in SLE progression, despite SLE being a multifactorial disease. Nevertheless, some studies have reported that the loss-of-function mutations in TREX1, RNaseH2C, and DNase I can lead to the aberrant accumulation of cytosolic DNA, which amplifies the type-I IFN signaling cascade in the cGAS–STING pathway [40,152,153]. Moreover, other studies of the serum of patients with SLE have indicated that a defect in eliminating apoptotic cells by macrophages results in the formation of apoptosis-derived membrane vesicles (AdMVs), which contain a significant amount of dsDNA. These AdMVs then induce increased production of type-I IFN to trigger an immune response that causes tissue damage in multiple organs and the regeneration of AdMVs, triggering a positive-loop of type-I IFN production [154]. Aicardi–Goutières syndrome (AGS) is another autoimmune disorder that presents as a progressive brain disease with onset in infancy. AGS is manifested by leukoencephalopathy with basal ganglia calcifications and progressive cerebral atrophy, as well as symptoms overlapping with SLE [41]. Furthermore, patients with AGS exhibit chronic cerebrospinal fluid lymphocytosis and immune complex deposition in tissues, similar to SLE [41]. AGS is caused by autosomal recessive heterogeneous mutations in any nine distinct genes, including LSM11, RNU7-1, TREX1, the RNaseH2 endonuclease complex (RNaseH2A, RNaseH2B, and RNaseH2C), SAMHD1, ADAR, and IFIH1, which are responsible for clearing diverse cytosolic genomic fragments [155,156]. Another study reported that approximately 25% of patients with AGS have mutations in the RNaseH2 endonuclease complex [157]. Furthermore, research has indicated that the loss-of-function mutations in the RNaseH2 catalytic core may adversely affect genome stability, resulting in the presence of cytosolic dsDNA and ensuing excessive cGAS–STING pathway-dependent expression of type-I IFN [158,159]. Finally, the autoimmune disorder rheumatoid arthritis (RA) is a chronic systemic inflammation typically associated with irreversible joint damage, disability, and cardiovascular comorbidities [160]. The etiology of RA is correlated with increased expression of several factors, including pro-inflammatory cytokines, chemokines, and matrix metalloproteinase (MMP), accompanied by the aberrant activation of T cells [161]. Moreover, the accumulation of cytosolic DNA plays a central role in the negative regulation of the inflammatory response in fibroblast-like synoviocytes (FLS) of patients with RA. This is because cytosolic DNA accumulation leads to the continuous activation of the cGAS–STING-dependent type-I IFN signaling cascade and the expression of cytokines and chemokines [161].
The inflammatory responses induced by the cGAS–STING pathway also contribute to the development of inflammation in other diseases. The first disease is amyotrophic lateral sclerosis (ALS), which has been suggested to be mainly caused by a mutation in the nuclear DNA/RNA binding protein called TAR DNA-binding protein 43 (TDP-43) [162]. Particularly, the missense mutation in a low-complexity glycine-rich region in TDP-43 enables TDP-43 to accumulate in the cytosol, and it is then absorbed into the mitochondria by mitochondrial import inner membrane translocase 22 (TIM22), causing the leakage of mtDNA into the cytosol via the permeability transition pore. The cytosolic mtDNA is then sensed by cGAS, triggering the production of type-I IFN and NF-κB to initiate hyperinflammatory responses [163]. The second disorder is Huntington disease (HD), which is caused by a mutation of the N-terminal polyglutamine in the huntingtin protein (mHTT) that induces damage to the brain’s striatum [164]. Research has suggested that the striatum injury is associated with DNA damage and upregulation of cGAS in HD cells, triggering type-I IFN inflammatory responses and STING-dependent autophagy [165]. The third condition is myocardial infarction (MI), wherein massive synchronous cell death in the heart and strong cardiac inflammatory response in related tissues are observed during the post-MI period [166]. Additionally, synchronously dying cells can be sensed by cardiac macrophages. This recognition by the cardiac macrophages induces the robust activation of the type-I IFN axis via cGAS–STING, which then maintains an inflammatory microenvironment by inhibiting the transformation of cardiac macrophages from inflammatory cells to a reparative phenotype [167]. The fourth disease is acute pancreatitis (AP), in which many pancreatic acinar cells die, leading to severe inflammation [168]. Furthermore, an independent investigation revealed that DNA released into circulation from necrotic pancreatic cells is recognized and internalized by leukocytes, triggering the cGAS–STING pathway to promote inflammation [169]. The fifth disorder showing inflammation development associated with the cGAS–STING pathway is silicosis. In silicosis, silica microparticles invade the lungs and adhere to lung parenchyma during inhalation, resulting in chronic progressive fibrotic inflammation in the lungs and other complications [170]. Moreover, the etiopathology of lung inflammation in patients with silicosis is associated with a significant increase in the release of chromatin fragments and mtDNA from dying cells into the bronchoalveolar lavage fluid (BALF) after silica exposure, inducing excess production of type-I IFN and CXCL10 via cGAS–STING activation [171]. The sixth ailment associated with the cGAS–STING pathway is nonalcoholic steatohepatitis (NASH). Lipotoxicity in NASH leads to mitochondrial stress and the release of mtDNA into the cytosol of Kupffer cells, which activates cGAS–STING to trigger adipose tissue inflammation in liver that further causes liver obesity, insulin resistance (IR), and glucose intolerance [172,173,174]. Finally, liver ischemia-reperfusion injury (IRI) is the last disorder to be linked with the cGAS–STING pathway. IRI has been reported to induce the generation of reactive oxygen species (ROS) in Kupffer cells, resulting in oxidative mitochondrial damage and the release of mtDNA into the cytosol, triggering the stimulation of NLRP3 inflammasome-driven cell death via cGAS–STING activation [175]. Additionally, another study indicated that the cGAS-STING pathway plays an important role in aging-related inflammation and neurodegeneration [176]. In aged microglia, an increased abundance of mtDNA in cytosol activates chronic inflammatory responses through the cGAS–STING pathway, which contributes to the impaired neuronal function, ensuing damaged brain homeostasis [176]. The different roles of the cGAS–STING pathway in host defense and various diseases were illustrated in Figure 3.

10. cGAS–STING Pathway in Cancer

The cGAS–STING pathway plays an indispensable role in radiotherapy, chemotherapy, and anti-tumor immune checkpoint therapy [177,178,179]. The paradigm of cancer is that irreparable damage occurs in normal tissue, which is strongly linked with persistent inflammation [180]. Furthermore, the cGAS–STING pathway has been proposed to not only contribute to the restriction of tumorigenesis, but also facilitate metastasis under certain conditions [181].
Moreover, cancer cells are packed with ectopic cytosolic dsDNA, and its accumulation may cause the formation and rupture of micronuclei [182]. In addition, oxidative stress and mitochondrial dysfunction induce the release of mtDNA into the cytosol, thus acting as another source of cytosolic DNA in malignant cells [183]. This cytosolic dsDNA is then recognized by cGAS, and tumor-suppressing biological activities are initiated. A study on early neoplastic progression demonstrated that after cGAS binds to cytosolic dsDNA in cancer cells, the cGAS–STING pathway mediates the secretion of type-I IFN to elicit a robust immune response and recruits immune cells to inhibit tumorigenesis in a cancer cell-autonomous manner [184]. Moreover, tumor DNA and cGAMP can be delivered into tumor-infiltrating dendritic cells (DCs) by tumor-derived exosomes (TEX) and the gap junction, respectively, which results in the production of type-I IFN that facilitates cross-priming with anti-tumor CD8+ T cells for the elimination of tumorigenic cells [185,186,187]. Additionally, cGAMP formed in melanoma cells can be transported to proximal non-tumor bystander cells via the gap junction, which results in the activation of STING in these cells and the recruitment of natural killer (NK) cells into the tumor tissue, thereby enhancing tumor regression [188,189]. Activation of the cGAS–STING pathway also induces the expression of anti-proliferative molecules that facilitate senescence in cancer cells [177]. Furthermore, autophagy induced by cGAS–STING activation can halt the transformation of normal cells into cancerous cells by initiating autophagy-dependent cell death in response to abnormal mitotic processes in normal cells [20].
The cGAS–STING pathway can also contribute to metastasis in a non-cell-autonomous manner. A study of patients with metastatic breast or lung cancers found that cGAMP in cancer cells is transferred to astrocytes via the gap junctions, which then activates astrocytic STING to stimulate inflammatory cytokine production. This production of inflammatory cytokines further causes the paracrine activation of the STAT1 and NF-κB pathways in brain metastatic cells, ultimately facilitating the development of metastatic brain cancer [190]. Another research study revealed that the chronic stimulation of the STING pathway contributes to the growth of a 7, 12-dimethylbenz(a)anthracene (DMBA)-induced skin tumor, indicating that STING activation must be maintained appropriately to avoid inflammation-driven tumorigenesis [191]. Moreover, etoposide, camptothecin, and H2O2 treatment have been shown to cause DNA damage and recruit cGAS to translocate into the nucleus in normal cells. The nuclear cGAS then attenuates the DNA damage response (DDR) mediated by homologous recombination, thereby facilitating cancer development in these cells [192].
Conversely, tumor cells can also escape the immune surveillance mediated by the cGAS–STING pathway. In specific cancer cell lines, the promoters of cGAS and STING are prone to undergo loss-of-function mutation or epigenetic silencing, which leads to the suppression of the cGAS–STING pathway [193]. In addition, the increased lactic acid levels in the lung tumor microenvironment have been revealed to inhibit STING-dependent type-I IFN production in DCs, hampering the capacity of tumor-conditioned DCs to present tumor-associated antigens [194]. Additionally, in breast cancer, activating the human epidermal growth factor receptor 2-RAC-alpha serine/threonine protein kinase (HER2-AKT1) axis inhibits cGAS and TBK1 enzymatic activity [56,195]. Some cancer cell lines also cause proliferation by a unique mechanism that utilizes extrachromosomal telomere repeat (ECTR) DNA to extend telomeres via the alternative lengthening of the telomeres (ALT) pathway [196], which additionally suppresses STING expression [197]. The cGAS-STING pathway was showed in Figure 4.

11. Therapeutic Strategies Targeting the cGAS–STING Pathway

As discussed in this review, the cGAS–STING pathway is involved in the pathogenesis of various diseases; therefore, drugs targeting cGAS–STING may provide therapeutic options. For example, inhibitors that negatively modulate the cGAS–STING pathway may be used to reduce the development of autoimmune disorders and local inflammation. In contrast, agonists that enhance the cGAS–STING pathway can improve immune responses and restrict the invasion of extracellular pathogens. In line with this notion, STING agonists can be applied as vaccine adjuvants against infectious diseases [10]. In the case of cancer therapy, cancer immunotherapy targeting the cGAS–STING pathway depends on multiple factors, such as the immune state, the magnitude of the cGAS–STING pathway, different cancer cell types, and tumor stages. Furthermore, the appropriate activation of the cGAS–STING pathway in immune cells provides beneficial effects that can restrain tumor growth, whereas the persistent activation of this pathway may contribute to the formation of carcinogen-induced tumors and arrest the development of T cell-driven adaptive immunity [198]; thus, examining the context of the cancer conditions is crucial in determining the applicability of the cGAS–STING pathway agonists in cancer immunotherapy. Nevertheless, the threshold of cGAS or STING activities must be maintained at normal levels to ensure effective anti-tumor responses and avoid facilitating the development of malignancy and side effects; therefore, developing STING agonists as adjuvants of cancer vaccines may be beneficial [199].
Inhibitors targeting the cGAS–STING pathway can be divided into two categories: cGAS inhibitors and STING inhibitors. The cGAS inhibitors include catalytic site inhibitors that block the activate site of cGAS and DNA-binding inhibitors that competitively bind to dsDNA to inhibit the interaction between DNA and cGAS. The STING inhibitors target the cyclic dinucleotide (CDN)-binding site and act as competitive antagonists of STING activators and palmitoylation inhibitors, which bind with the palmitoylation sites on STING to reduce the recruitment of downstream signal factors and the transformation of STING [13]. Table 5 summarizes these inhibitors and their biological effects. In terms of the cGAS–STING pathway agonists, studies have concentrated on the STING agonists due to their anti-tumor effect. These agonists include CDN, non-CDN, and indirect agonists [200]. Table 6 displays the typical STING agonists and their preclinical effects.

12. Conclusions and Future Prospects

Since the discovery of the cGAS–STING pathway, numerous advances have been made in identifying its components the mechanisms by which DNA activates this pathway to induce the secretion of type-I IFN; however, further research is required to address the inconsistent findings regarding the binding of cGAS to ligands from different sources and the underlying subcellular compartmentalization of cGAS. Furthermore, large gaps still exist in understanding the modulation patterns of the cGAS–STING pathway because these alterations involve various modifications by numerous intracellular molecules and signaling networks during the post-translational stages; therefore, future investigations targeting the modulation patterns of the cGAS–STING pathway are required. Moreover, additional studies are required to explore the physiological roles of cGAS and its potential implementation in other intracellular processes involving the induction of type-I IFN.
In summary, our review indicates that cGAS has been well-established as a major sensor that activates the immune response against pathogens. In addition, research on the cGAS–STING pathway can help its application as a novel treatment strategy for cancer, considering its role in inducing cell-intrinsic processes, such as autophagy, cellular senescence, and cell death, or in enhancing innate immunity for the activation of adaptive immunity. In contrast, aberrant or excessive activation of the cGAS–STING pathway can cause primary pathogenesis and the manifestation of several autoimmune diseases; thus, identifying compounds, delivery pathways, and treatment regimens targeting the suppression of the cGAS–STING pathway can provide novel approaches to alleviate the symptoms of autoimmune disorders or inflammation. Conversely, immunopharmacological research on developing cGAS–STING pathway agonists can help formulate therapies for infectious diseases and cancer, such as anti-pathogen or anti-cancer drugs and adjuvants; however, these drugs should be optimized to augment the desirable effect and should not induce any unwanted effects. A reasonable approach may involve controlling the pharmacokinetic parameters appropriately to avoid chronic inflammation triggered by cGAS–STING-induced cytokine storms and to ensure that the drugs reach the target area without side effects. Additionally, considering that discrepancies may arise during drug development due to the differences in the component structures of the cGAS–STING pathway between humans and other species, the preclinical drugs should be carefully examined to determine their suitability in clinical trials. Nevertheless, we anticipate that the development of drugs targeting the cGAS–STING pathway and their translation into clinical applications in the near future will help further ameliorate the condition of patients with cancer, inflammation, and autoimmune diseases.

Author Contributions

J.Z. and Z.Z. contributed equally to this article. J.Z. and Z.Z. drafted the original article and prepared the figures. J.L. searched for relevant literature and wrote this paper. Z.F. proposed the research design, searched for relevant literature, and wrote this paper. Z.F. offered professional advice and revision. All authors critically reviewed the content and approved the final version of this paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Roers, A.; Hiller, B.; Hornung, V. Recognition of Endogenous Nucleic Acids by the Innate Immune System. Immunity 2016, 44, 739–754. [Google Scholar] [CrossRef]
  2. Gong, T.; Liu, L.; Jiang, W.; Zhou, R. DAMP-sensing receptors in sterile inflammation and inflammatory diseases. Nat. Rev. Immunol. 2020, 20, 95–112. [Google Scholar] [CrossRef]
  3. Hopfner, K.P.; Hornung, V. Molecular mechanisms and cellular functions of cGAS-STING signalling. Nat. Rev. Mol. Cell Biol. 2020, 21, 501–521. [Google Scholar] [CrossRef]
  4. Ishikawa, H.; Barber, G.N. STING is an endoplasmic reticulum adaptor that facilitates innate immune signalling. Nature 2008, 455, 674–678. [Google Scholar] [CrossRef]
  5. 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]
  6. 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]
  7. 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, 786–791. [Google Scholar] [CrossRef]
  8. Mabbott, N.A.; Baillie, J.K.; Brown, H.; Freeman, T.C.; Hume, D.A. An expression atlas of human primary cells: Inference of gene function from coexpression networks. BMC Genom. 2013, 14, 632. [Google Scholar] [CrossRef]
  9. Wan, D.; Jiang, W.; Hao, J. Research Advances in How the cGAS-STING Pathway Controls the Cellular Inflammatory Response. Front. Immunol. 2020, 11, 615. [Google Scholar] [CrossRef]
  10. Li, X.D.; Wu, J.; Gao, D.; Wang, H.; Sun, L.; Chen, Z.J. Pivotal roles of cGAS-cGAMP signaling in antiviral defense and immune adjuvant effects. Science 2013, 341, 1390–1394. [Google Scholar] [CrossRef]
  11. Li, Q.; Tian, S.; Liang, J.; Fan, J.; Lai, J.; Chen, Q. Therapeutic Development by Targeting the cGAS-STING Pathway in Autoimmune Disease and Cancer. Front. Pharmacol. 2021, 12, 779425. [Google Scholar] [CrossRef] [PubMed]
  12. Crow, Y.J.; Stetson, D.B. The type I interferonopathies: 10 years on. Nat. Rev. Immunol. 2021, 22, 471–483. [Google Scholar] [CrossRef] [PubMed]
  13. Decout, A.; Katz, J.D.; Venkatraman, S.; Ablasser, A. The cGAS-STING pathway as a therapeutic target in inflammatory diseases. Nat. Rev. Immunol. 2021, 21, 548–569. [Google Scholar] [CrossRef]
  14. Volkman, H.E.; Cambier, S.; Gray, E.E.; Stetson, D.B. Tight nuclear tethering of cGAS is essential for preventing autoreactivity. Elife 2019, 8, e47491. [Google Scholar] [CrossRef] [PubMed]
  15. Civril, F.; Deimling, T.; Mann, C.C.d.O.; 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] [PubMed]
  16. West, A.P.; Khoury-Hanold, W.; Staron, M.; Tal, M.C.; Pineda, C.M.; Lang, S.M.; Bestwick, M.; Duguay, B.A.; Raimundo, N.; MacDuff, D.A.; et al. Mitochondrial DNA stress primes the antiviral innate immune response. Nature 2015, 520, 553–557. [Google Scholar] [CrossRef]
  17. 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]
  18. Fenech, M.; Kirsch-Volders, M.; Natarajan, A.T.; Surralles, J.; Crott, J.W.; Parry, J.; Norppa, H.; Eastmond, D.A.; Tucker, J.D.; Thomas, P. Molecular mechanisms of micronucleus, nucleoplasmic bridge and nuclear bud formation in mammalian and human cells. Mutagenesis 2011, 26, 125–132. [Google Scholar] [CrossRef]
  19. Hatch, E.M.; Fischer, A.H.; Deerinck, T.J.; Hetzer, M.W. Catastrophic Nuclear Envelope Collapse in Cancer Cell Micronuclei. Cell 2013, 154, 47–60. [Google Scholar] [CrossRef]
  20. Nassour, J.; Radford, R.; Correia, A.; Fusté, J.M.; Schoell, B.; Jauch, A.; Shaw, R.J.; Karlseder, J. Autophagic cell death restricts chromosomal instability during replicative crisis. Nature 2019, 565, 659–663. [Google Scholar] [CrossRef]
  21. Herzner, A.-M.; Hagmann, C.A.; Goldeck, M.; Wolter, S.; Kübler, K.; Wittmann, S.; Gramberg, T.; Andreeva, L.; Hopfner, K.-P.; Mertens, C.; et al. Sequence-specific activation of the DNA sensor cGAS by Y-form DNA structures as found in primary HIV-1 cDNA. Nat. Immunol. 2015, 16, 1025–1033. [Google Scholar] [CrossRef] [PubMed]
  22. Guey, B.; Wischnewski, M.; Decout, A.; Makasheva, K.; Kaynak, M.; Sakar, M.S.; Fierz, B.; Ablasser, A. BAF restricts cGAS on nuclear DNA to prevent innate immune activation. Science 2020, 369, 823–828. [Google Scholar] [CrossRef] [PubMed]
  23. Zhou, W.; Whiteley, A.T.; de Oliveira Mann, C.C.; Morehouse, B.R.; Nowak, R.P.; Fischer, E.S.; Gray, N.S.; Mekalanos, J.J.; Kranzusch, P.J. Structure of the Human cGAS-DNA Complex Reveals Enhanced Control of Immune Surveillance. Cell 2018, 174, 300–311.e11. [Google Scholar] [CrossRef] [PubMed]
  24. Andreeva, L.; Hiller, B.; Kostrewa, D.; Lässig, C.; Mann, C.C.d.O.; Drexler, D.J.; Maiser, A.; Gaidt, M.; Leonhardt, H.; Hornung, V.; et al. cGAS senses long and HMGB/TFAM-bound U-turn DNA by forming protein–DNA ladders. Nature 2017, 549, 394–398. [Google Scholar] [CrossRef]
  25. Ablasser, 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]
  26. Wang, Y.; Luo, J.; Alu, A.; Han, X.; Wei, Y.; Wei, X. cGAS-STING pathway in cancer biotherapy. Mol. Cancer 2020, 19, 136. [Google Scholar] [CrossRef]
  27. 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]
  28. 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]
  29. Du, M.; Chen, Z.J. DNA-induced liquid phase condensation of cGAS activates innate immune signaling. Science 2018, 361, 704–709. [Google Scholar] [CrossRef]
  30. Xie, W.; Lama, L.; Adura, C.; Tomita, D.; Glickman, J.F.; Tuschl, T.; Patel, D.J. Human cGAS catalytic domain has an additional DNA-binding interface that enhances enzymatic activity and liquid-phase condensation. Proc. Natl. Acad. Sci. USA 2019, 116, 11946–11955. [Google Scholar] [CrossRef]
  31. Diner, E.J.; Burdette, D.L.; Wilson, S.C.; Monroe, K.M.; Kellenberger, C.A.; Hyodo, M.; Hayakawa, Y.; Hammond, M.C.; Vance, R.E. The Innate Immune DNA Sensor cGAS Produces a Noncanonical Cyclic Dinucleotide that Activates Human STING. Cell Rep. 2013, 3, 1355–1361. [Google Scholar] [CrossRef]
  32. Ergun, S.L.; Fernandez, D.; Weiss, T.M.; Li, L. STING Polymer Structure Reveals Mechanisms for Activation, Hyperactivation, and Inhibition. Cell 2019, 178, 290–301.e10. [Google Scholar] [CrossRef]
  33. 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] [PubMed]
  34. Luo, W.-W.; Li, S.; Li, C.; Lian, H.; Yang, Q.; Zhong, B.; Shu, H.-B. iRhom2 is essential for innate immunity to DNA viruses by mediating trafficking and stability of the adaptor STING. Nat. Immunol. 2016, 17, 1057–1066. [Google Scholar] [CrossRef] [PubMed]
  35. Dobbs, N.; Burnaevskiy, N.; Chen, D.; Gonugunta, V.K.; Alto, N.M.; Yan, N. STING Activation by Translocation from the ER Is Associated with Infection and Autoinflammatory Disease. Cell Host Microbe 2015, 18, 157–168. [Google Scholar] [CrossRef]
  36. 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]
  37. Jiang, M.; Chen, P.; Wang, L.; Li, W.; Chen, B.; Liu, Y.; Wang, H.; Zhao, S.; Ye, L.; He, Y.; et al. cGAS-STING, an important pathway in cancer immunotherapy. J. Hematol. Oncol. 2020, 13, 1–11. [Google Scholar] [CrossRef]
  38. Chen, Q.; Sun, L.; Chen, Z. Regulation and function of the cGAS-STING pathway of cytosolic DNA sensing. Nat. Immunol. 2016, 17, 1142–1149. [Google Scholar] [CrossRef]
  39. Hou, Y.; Liang, H.; Rao, E.; Zheng, W.; Huang, X.; Deng, L.; Zhang, Y.; Yu, X.; Xu, M.; Mauceri, H.; et al. Non-canonical NF-κB Antagonizes STING Sensor-Mediated DNA Sensing in Radiotherapy. Immunity 2018, 49, 490–503.e4. [Google Scholar] [CrossRef]
  40. Xiao, N.; Wei, J.; Xu, S.; Du, H.; Huang, M.; Zhang, S.; Ye, W.; Sun, L.; Chen, Q. cGAS activation causes lupus-like autoimmune disorders in a TREX1 mutant mouse model. J. Autoimmun. 2019, 100, 84–94. [Google Scholar] [CrossRef]
  41. Crow, Y.J.; Manel, N. Aicardi-Goutières syndrome and the type I interferonopathies. Nat. Rev. Immunol. 2015, 15, 429–440. [Google Scholar] [CrossRef] [PubMed]
  42. Coquel, F.; Silva, M.-J.; Técher, H.; Zadorozhny, K.; Sharma, S.; Nieminuszczy, J.; Mettling, C.; Dardillac, E.; Barthe, A.; Schmitz, A.-L.; et al. SAMHD1 acts at stalled replication forks to prevent interferon induction. Nature 2018, 557, 57–61. [Google Scholar] [CrossRef] [PubMed]
  43. Grabosch, S.; Bulatovic, M.; Zeng, F.; Ma, T.; Zhang, L.; Ross, M.; Brozick, J.; Fang, Y.; Tseng, G.; Kim, E.; et al. Cisplatin-induced immune modulation in ovarian cancer mouse models with distinct inflammation profiles. Oncogene 2018, 38, 2380–2393. [Google Scholar] [CrossRef]
  44. Jønsson, K.L.; Laustsen, A.; Krapp, C.; Skipper, K.A.; Thavachelvam, K.; Hotter, D.; Egedal, J.H.; Kjolby, M.; Mohammadi, P.; Prabakaran, T.; et al. IFI16 is required for DNA sensing in human macrophages by promoting production and function of cGAMP. Nat. Commun. 2017, 8, 14391. [Google Scholar] [CrossRef] [PubMed]
  45. Almine, J.F.; O’hare, C.A.J.; Dunphy, G.; Haga, I.R.; Naik, R.J.; Atrih, A.; Connolly, D.J.; Taylor, J.; Kelsall, I.R.; Bowie, A.G.; et al. IFI16 and cGAS cooperate in the activation of STING during DNA sensing in human keratinocytes. Nat. Commun. 2017, 8, 14392. [Google Scholar] [CrossRef]
  46. Orzalli, M.H.; Broekema, N.M.; Diner, B.A.; Hancks, D.C.; Elde, N.C.; Cristea, I.M.; Knipe, D.M. cGAS-mediated stabilization of IFI16 promotes innate signaling during herpes simplex virus infection. Proc. Natl. Acad. Sci. USA 2015, 112, E1773–E1781. [Google Scholar] [CrossRef]
  47. Zheng, W.; Zhou, R.; Li, S.; He, S.; Luo, J.; Zhu, M.; Chen, N.; Chen, H.; Meurens, F.; Zhu, J. Porcine IFI16 Negatively Regulates cGAS Signaling Through the Restriction of DNA Binding and Stimulation. Front. Immunol. 2020, 11, 1669. [Google Scholar] [CrossRef]
  48. Li, Y.-K.; Chen, J.-G.; Wang, F. The emerging roles of absent in melanoma 2 (AIM2) inflammasome in central nervous system disorders. Neurochem. Int. 2021, 149, 105122. [Google Scholar] [CrossRef]
  49. Zhu, W.; Zu, X.; Liu, S.; Zhang, H. The absent in melanoma 2 (AIM2) inflammasome in microbial infection. Clin. Chim. Acta 2019, 495, 100–108. [Google Scholar] [CrossRef]
  50. Corrales, L.; Woo, S.-R.; Williams, J.B.; McWhirter, S.M.; Dubensky, T.W.; Gajewski, T.F. Antagonism of the STING Pathway via Activation of the AIM2 Inflammasome by Intracellular DNA. J. Immunol. 2016, 196, 3191–3198. [Google Scholar] [CrossRef]
  51. Wang, Y.; Ning, X.; Gao, P.; Wu, S.; Sha, M.; Lv, M.; Zhou, X.; Gao, J.; Fang, R.; Meng, G.; et al. Inflammasome Activation Triggers Caspase-1-Mediated Cleavage of cGAS to Regulate Responses to DNA Virus Infection. Immunity 2017, 46, 393–404. [Google Scholar] [CrossRef] [PubMed]
  52. Banerjee, I.; Behl, B.; Mendonca, M.; Shrivastava, G.; Russo, A.J.; Menoret, A.; Ghosh, A.; Vella, A.T.; Vanaja, S.K.; Sarkar, S.N.; et al. Gasdermin D Restrains Type I Interferon Response to Cytosolic DNA by Disrupting Ionic Homeostasis. Immunity 2018, 49, 413–426.e5. [Google Scholar] [CrossRef] [PubMed]
  53. Briard, B.; Place, D.E.; Kanneganti, T.-D. DNA Sensing in the Innate Immune Response. Physiology 2020, 35, 112–124. [Google Scholar] [CrossRef]
  54. Amadio, R.; Piperno, G.; Benvenuti, F. Self-DNA Sensing by cGAS-STING and TLR9 in Autoimmunity: Is the Cytoskeleton in Control? Front. Immunol. 2021, 12, 657344. [Google Scholar] [CrossRef] [PubMed]
  55. Deb, P.; Dai, J.; Singh, S.; Kalyoussef, E.; Fitzgerald-Bocarsly, P. Triggering of the cGAS-STING Pathway in Human Plasmacytoid Dendritic Cells Inhibits TLR9-Mediated IFN Production. J. Immunol. 2020, 205, 223–236. [Google Scholar] [CrossRef]
  56. Seo, G.J.; Yang, A.; Tan, B.; Kim, S.; Liang, Q.; Choi, Y.; Yuan, W.; Feng, P.; Park, H.-S.; Jung, J.U. Akt Kinase-Mediated Checkpoint of cGAS DNA Sensing Pathway. Cell Rep. 2015, 13, 440–449. [Google Scholar] [CrossRef]
  57. Jiang, H.; Xue, X.; Panda, S.; Kawale, A.; Hooy, R.M.; Liang, F.; Sohn, J.; Sung, P.; Gekara, N.O. Chromatin-bound cGAS is an inhibitor of DNA repair and hence accelerates genome destabilization and cell death. EMBO J. 2019, 38, e102718. [Google Scholar] [CrossRef]
  58. Hu, M.-M.; Yang, Q.; Xie, X.-Q.; Liao, C.-Y.; Lin, H.; Liu, T.-T.; Yin, L.; Shu, H.-B. Sumoylation Promotes the Stability of the DNA Sensor cGAS and the Adaptor STING to Regulate the Kinetics of Response to DNA Virus. Immunity 2016, 45, 555–569. [Google Scholar] [CrossRef]
  59. Xia, P.; Ye, B.; Wang, S.; Zhu, X.; Du, Y.; Xiong, Z.; Tian, Y.; Fan, Z. Glutamylation of the DNA sensor cGAS regulates its binding and synthase activity in antiviral immunity. Nat. Immunol. 2016, 17, 369–378. [Google Scholar] [CrossRef]
  60. Xia, P.; Wang, S.; Ye, B.; Du, Y.; Li, C.; Xiong, Z.; Qu, Y.; Fan, Z. A Circular RNA Protects Dormant Hematopoietic Stem Cells from DNA Sensor cGAS-Mediated Exhaustion. Immunity 2018, 48, 688–701.e7. [Google Scholar] [CrossRef]
  61. Ghosh, A.; Shao, L.; Sampath, P.; Zhao, B.; Patel, N.V.; Zhu, J.; Behl, B.; Parise, R.A.; Beumer, J.H.; O’sullivan, R.J.; et al. Oligoadenylate-Synthetase-Family Protein OASL Inhibits Activity of the DNA Sensor cGAS during DNA Virus Infection to Limit Interferon Production. Immunity 2019, 50, 51–63.e5. [Google Scholar] [CrossRef] [PubMed]
  62. Luteijn, R.D.; Zaver, S.A.; Gowen, B.G.; Wyman, S.K.; Garelis, N.E.; Onia, L.; McWhirter, S.M.; Katibah, G.E.; Corn, J.E.; Woodward, J.J.; et al. SLC19A1 transports immunoreactive cyclic dinucleotides. Nature 2019, 573, 434–438. [Google Scholar] [CrossRef] [PubMed]
  63. Sun, H.; Zhang, Q.; Jing, Y.-Y.; Zhang, M.; Wang, H.-Y.; Cai, Z.; Liuyu, T.; Zhang, Z.-D.; Xiong, T.-C.; Wu, Y.; et al. USP13 negatively regulates antiviral responses by deubiquitinating STING. Nat. Commun. 2017, 8, 15534. [Google Scholar] [CrossRef] [PubMed]
  64. Zhang, L.; Mo, J.; Swanson, K.V.; Wen, H.; Petrucelli, A.; Gregory, S.M.; Zhang, Z.; Schneider, M.; Jiang, Y.; Fitzgerald, K.A.; et al. NLRC3, a Member of the NLR Family of Proteins, Is a Negative Regulator of Innate Immune Signaling Induced by the DNA Sensor STING. Immunity 2014, 40, 329–341. [Google Scholar] [CrossRef] [PubMed]
  65. Guo, H.; König, R.; Deng, M.; Riess, M.; Mo, J.; Zhang, L.; Petrucelli, A.; Yoh, S.M.; Barefoot, B.; Samo, M.; et al. NLRX1 Sequesters STING to Negatively Regulate the Interferon Response, Thereby Facilitating the Replication of HIV-1 and DNA Viruses. Cell Host Microbe 2016, 19, 515–528. [Google Scholar] [CrossRef]
  66. Konno, H.; Konno, K.; Barber, G.N. Cyclic Dinucleotides Trigger ULK1 (ATG1) Phosphorylation of STING to Prevent Sustained Innate Immune Signaling. Cell 2013, 155, 688–698. [Google Scholar] [CrossRef]
  67. Zhong, B.; Zhang, L.; Lei, C.; Li, Y.; Mao, A.-P.; Yang, Y.; Wang, Y.-Y.; Zhang, X.-L.; Shu, H.-B. The Ubiquitin Ligase RNF5 Regulates Antiviral Responses by Mediating Degradation of the Adaptor Protein MITA. Immunity 2009, 30, 397–407. [Google Scholar] [CrossRef]
  68. Liu, Y.; Xu, P.; Rivara, S.; Liu, C.; Ricci, J.; Ren, X.; Hurley, J.H.; Ablasser, A. Clathrin-associated AP-1 controls termination of STING signalling. Nature 2022, 610, 761–767. [Google Scholar] [CrossRef]
  69. 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]
  70. Cui, Y.; Yu, H.; Zheng, X.; Peng, R.; Wang, Q.; Zhou, Y.; Wang, R.; Wang, J.; Qu, B.; Shen, N.; et al. SENP7 Potentiates cGAS Activation by Relieving SUMO-Mediated Inhibition of Cytosolic DNA Sensing. PLoS Pathog. 2017, 13, e1006156. [Google Scholar] [CrossRef]
  71. Dai, J.; Huang, Y.-J.; He, X.; Zhao, M.; Wang, X.; Liu, Z.-S.; Xue, W.; Cai, H.; Zhan, X.-Y.; Huang, S.-Y.; et al. Acetylation Blocks cGAS Activity and Inhibits Self-DNA-Induced Autoimmunity. Cell 2019, 176, 1447–1460.e14. [Google Scholar] [CrossRef] [PubMed]
  72. Wang, Q.; Huang, L.; Hong, Z.; Lv, Z.; Mao, Z.; Tang, Y.; Kong, X.; Li, S.; Cui, Y.; Liu, H.; et al. The E3 ubiquitin ligase RNF185 facilitates the cGAS-mediated innate immune response. PLoS Pathog. 2017, 13, e1006264. [Google Scholar] [CrossRef] [PubMed]
  73. Wang, C.; Guan, Y.; Lv, M.; Zhang, R.; Guo, Z.; Wei, X.; Du, X.; Yang, J.; Li, T.; Wan, Y.; et al. Manganese Increases the Sensitivity of the cGAS-STING Pathway for Double-Stranded DNA and Is Required for the Host Defense against DNA Viruses. Immunity 2018, 48, 675–687.e7. [Google Scholar] [CrossRef] [PubMed]
  74. Liu, Z.S.; Cai, H.; Xue, W.; Wang, M.; Xia, T.; Li, W.J.; Xing, J.Q.; Zhao, M.; Huang, Y.J.; Chen, S.; et al. G3BP1 promotes DNA binding and activation of cGAS. Nat. Immunol. 2019, 20, 18–28. [Google Scholar] [CrossRef] [PubMed]
  75. Johnson, J.S.; Lucas, S.Y.; Amon, L.M.; Skelton, S.; Nazitto, R.; Carbonetti, S.; Sather, D.N.; Littman, D.R.; Aderem, A. Reshaping of the Dendritic Cell Chromatin Landscape and Interferon Pathways during HIV Infection. Cell Host Microbe 2018, 23, 366–381.e9. [Google Scholar] [CrossRef] [PubMed]
  76. Siddiqui, M.A.; Yamashita, M. Toll-Like Receptor (TLR) Signaling Enables Cyclic GMP-AMP Synthase (cGAS) Sensing of HIV-1 Infection in Macrophages. mBio 2021, 12, e0281721. [Google Scholar] [CrossRef] [PubMed]
  77. Liu, Z.-S.; Zhang, Z.-Y.; Cai, H.; Zhao, M.; Mao, J.; Dai, J.; Xia, T.; Zhang, X.-M.; Li, T. RINCK-mediated monoubiquitination of cGAS promotes antiviral innate immune responses. Cell Biosci. 2018, 8, 35. [Google Scholar] [CrossRef]
  78. Chen, M.; Meng, Q.; Qin, Y.; Liang, P.; Tan, P.; He, L.; Zhou, Y.; Chen, Y.; Huang, J.; Wang, R.-F.; et al. TRIM14 Inhibits cGAS Degradation Mediated by Selective Autophagy Receptor p62 to Promote Innate Immune Responses. Mol. Cell 2016, 64, 105–119. [Google Scholar] [CrossRef]
  79. Wang, Q.; Liu, X.; Cui, Y.; Tang, Y.; Chen, W.; Li, S.; Yu, H.; Pan, Y.; Wang, C. The E3 Ubiquitin Ligase AMFR and INSIG1 Bridge the Activation of TBK1 Kinase by Modifying the Adaptor STING. Immunity 2014, 41, 919–933. [Google Scholar] [CrossRef]
  80. Ni, G.; Konno, H.; Barber, G.N. Ubiquitination of STING at lysine 224 controls IRF3 activation. Sci. Immunol. 2017, 2. [Google Scholar] [CrossRef]
  81. Li, Y.; James, S.J.; Wyllie, D.H.; Wynne, C.; Czibula, A.; Bukhari, A.; Pye, K.; Mustafah, S.M.B.; Fajka-Boja, R.; Szabo, E.; et al. TMEM203 is a binding partner and regulator of STING-mediated inflammatory signaling in macrophages. Proc. Natl. Acad. Sci. USA 2019, 116, 16479–16488. [Google Scholar] [CrossRef] [PubMed]
  82. Glick, D.; Barth, S.; MacLeod, K.F. Autophagy: Cellular and molecular mechanisms. J. Pathol. 2010, 221, 3–12. [Google Scholar] [CrossRef] [PubMed]
  83. Gui, X.; Yang, H.; Li, T.; Tan, X.; Shi, P.; Li, M.; Du, F.; Chen, Z.J. Autophagy induction via STING trafficking is a primordial function of the cGAS pathway. Nature 2019, 567, 262–266. [Google Scholar] [CrossRef]
  84. Motwani, M.; Pesiridis, S.; Fitzgerald, K. DNA sensing by the cGAS-STING pathway in health and disease. Nat. Rev. Genet. 2019, 20, 657–674. [Google Scholar] [CrossRef] [PubMed]
  85. Moretti, J.; Roy, S.; Bozec, D.; Martinez, J.; Chapman, J.R.; Ueberheide, B.; Lamming, D.W.; Chen, Z.J.; Horng, T.; Yeretssian, G.; et al. Faculty Opinions recommendation of STING Senses Microbial Viability to Orchestrate Stress-Mediated Autophagy of the Endoplasmic Reticulum. Cell 2017, 171, 809–823.e13. [Google Scholar] [CrossRef]
  86. Liu, Y.; Gordesky-Gold, B.; Leney-Greene, M.; Weinbren, N.L.; Tudor, M.; Cherry, S. Inflammation-Induced, STING-Dependent Autophagy Restricts Zika Virus Infection in the Drosophila Brain. Cell Host Microbe 2018, 24, 57–68.e3. [Google Scholar] [CrossRef]
  87. Mariño, G.; Niso-Santano, M.; Baehrecke, E.H.; Kroemer, G. Self-consumption: The interplay of autophagy and apoptosis. Nat. Rev. Mol. Cell Biol. 2014, 15, 81–94. [Google Scholar] [CrossRef]
  88. Wang, F.; Gómez-Sintes, R.; Boya, P. Lysosomal membrane permeabilization and cell death. Traffic 2018, 19, 918–931. [Google Scholar] [CrossRef]
  89. Gaidt, M.M.; Ebert, T.S.; Chauhan, D.; Ramshorn, K.; Pinci, F.; Zuber, S.; O’Duill, F.; Schmid-Burgk, J.L.; Hoss, F.; Buhmann, R.; et al. Faculty Opinions recommendation of The DNA Inflammasome in Human Myeloid Cells Is Initiated by a STING-Cell Death Program Upstream of NLRP3. Cell 2017, 171, 1110–1124.e18. [Google Scholar] [CrossRef]
  90. Pistritto, G.; Trisciuoglio, D.; Ceci, C.; Garufi, A.; D'Orazi, G. Apoptosis as anticancer mechanism: Function and dysfunction of its modulators and targeted therapeutic strategies. Aging 2016, 8, 603–619. [Google Scholar] [CrossRef]
  91. Elmore, S. Apoptosis: A review of programmed cell death. Toxicol. Pathol. 2007, 35, 495–516. [Google Scholar] [CrossRef] [PubMed]
  92. Tait, S.W.G.; Green, D.R. Mitochondria and cell death: Outer membrane permeabilization and beyond. Nat. Rev. Mol. Cell Biol. 2010, 11, 621–632. [Google Scholar] [CrossRef] [PubMed]
  93. Paludan, S.R.; Reinert, L.S.; Hornung, V. DNA-stimulated cell death: Implications for host defence, inflammatory diseases and cancer. Nat. Rev. Immunol. 2019, 19, 141–153. [Google Scholar] [CrossRef] [PubMed]
  94. Murthy, A.M.V.; Robinson, N.; Kumar, S. Crosstalk between cGAS-STING signaling and cell death. Cell Death Differ. 2020, 27, 2989–3003. [Google Scholar] [CrossRef]
  95. Rongvaux, A.; Jackson, R.; Harman, C.C.; Li, T.; West, A.P.; de Zoete, M.R.; Wu, Y.; Yordy, B.; Lakhani, S.A.; Kuan, C.-Y.; et al. Apoptotic Caspases Prevent the Induction of Type I Interferons by Mitochondrial DNA. Cell 2014, 159, 1563–1577. [Google Scholar] [CrossRef]
  96. Ning, X.; Wang, Y.; Jing, M.; Sha, M.; Lv, M.; Gao, P.; Zhang, R.; Huang, X.; Feng, J.-M.; Jiang, Z. Apoptotic Caspases Suppress Type I Interferon Production via the Cleavage of cGAS, MAVS, and IRF3. Mol. Cell 2019, 74, 19–31.e7. [Google Scholar] [CrossRef]
  97. Cui, Y.; Zhao, D.; Sreevatsan, S.; Liu, C.; Yang, W.; Song, Z.; Yang, L.; Barrow, P.; Zhou, X. Mycobacterium bovis Induces Endoplasmic Reticulum Stress Mediated-Apoptosis by Activating IRF3 in a Murine Macrophage Cell Line. Front. Cell. Infect. Microbiol. 2016, 6, 182. [Google Scholar] [CrossRef]
  98. Chattopadhyay, S.; Marques, J.T.; Yamashita, M.; Peters, K.L.; Smith, K.; Desai, A.; Williams, B.R.G.; Sen, G.C. Viral apoptosis is induced by IRF-3-mediated activation of Bax. EMBO J. 2010, 29, 1762–1773. [Google Scholar] [CrossRef]
  99. Frank, D.; Vince, J.E. Pyroptosis versus necroptosis: Similarities, differences, and crosstalk. Cell Death Differ. 2019, 26, 99–114. [Google Scholar] [CrossRef]
  100. Robinson, N.; McComb, S.; Mulligan, R.; Dudani, R.; Krishnan, L.; Sad, S. Type I interferon induces necroptosis in macrophages during infection with Salmonella enterica serovar Typhimurium. Nat. Immunol. 2012, 13, 954–962. [Google Scholar] [CrossRef]
  101. Brault, M.; Olsen, T.M.; Martinez, J.; Stetson, D.B.; Oberst, A. Intracellular Nucleic Acid Sensing Triggers Necroptosis through Synergistic Type I IFN and TNF Signaling. J. Immunol. 2018, 200, 2748–2756. [Google Scholar] [CrossRef] [PubMed]
  102. Sarhan, J.; Liu, B.C.; Muendlein, H.I.; Weindel, C.G.; Smirnova, I.; Tang, A.Y.; Ilyukha, V.; Sorokin, M.; Buzdin, A.; Fitzgerald, K.A.; et al. Constitutive interferon signaling maintains critical threshold of MLKL expression to license necroptosis. Cell Death Differ. 2019, 26, 332–347. [Google Scholar] [CrossRef] [PubMed]
  103. Regulski, M.J. Cellular Senescence: What, Why, and How. Wounds 2017, 29, 168–174. [Google Scholar] [PubMed]
  104. Sikora, E.; Bielak-Żmijewska, A.; Mosieniak, G. What is and what is not cell senescence. Postepy Biochem. 2018, 64, 110–118. [Google Scholar] [CrossRef] [PubMed]
  105. Yang, H.; Wang, H.; Ren, J.; Chen, Q.; Chen, Z.J. cGAS is essential for cellular senescence. Proc. Natl. Acad. Sci. USA 2017, 114, E4612–E4620. [Google Scholar] [CrossRef]
  106. Lan, Y.Y.; Heather, J.M.; Eisenhaure, T.; Garris, C.S.; Lieb, D.; Raychowdhury, R.; Hacohen, N. Extranuclear DNA accumulates in aged cells and contributes to senescence and inflammation. Aging Cell 2019, 18, e12901. [Google Scholar] [CrossRef]
  107. Lopes-Paciencia, S.; Saint-Germain, E.; Rowell, M.-C.; Ruiz, A.F.; Kalegari, P.; Ferbeyre, G. The senescence-associated secretory phenotype and its regulation. Cytokine 2019, 117, 15–22. [Google Scholar] [CrossRef]
  108. Glück, S.; Guey, B.; Gulen, M.F.; Wolter, K.; Kang, T.-W.; Schmacke, N.A.; Bridgeman, A.; Rehwinkel, J.; Zender, L.; Ablasser, A. Innate immune sensing of cytosolic chromatin fragments through cGAS promotes senescence. Nature 2017, 19, 1061–1070. [Google Scholar] [CrossRef]
  109. Kobayashi, H.; Kobayashi, C.I.; Nakamura-Ishizu, A.; Karigane, D.; Haeno, H.; Yamamoto, K.N.; Sato, T.; Ohteki, T.; Hayakawa, Y.; Barber, G.N.; et al. Bacterial c-di-GMP Affects Hematopoietic Stem/Progenitors and Their Niches through STING. Cell Rep. 2015, 11, 71–84. [Google Scholar] [CrossRef]
  110. Yu, Q.; Katlinskaya, Y.V.; Carbone, C.J.; Zhao, B.; Katlinski, K.V.; Zheng, H.; Guha, M.; Li, N.; Chen, Q.; Yang, T.; et al. DNA-Damage-Induced Type I Interferon Promotes Senescence and Inhibits Stem Cell Function. Cell Rep. 2015, 11, 785–797. [Google Scholar] [CrossRef]
  111. Zhang, X.; Bai, X.C.; Chen, Z.J. Structures and Mechanisms in the cGAS-STING Innate Immunity Pathway. Immunity 2020, 53, 43–53. [Google Scholar] [CrossRef]
  112. Sun, B.; Sundström, K.B.; Chew, J.J.; Bist, P.; Gan, E.S.; Tan, H.C.; Goh, K.C.; Chawla, T.; Tang, C.K.; Ooi, E.E. Dengue virus activates cGAS through the release of mitochondrial DNA. Sci. Rep. 2017, 7, 3594. [Google Scholar] [CrossRef]
  113. Domizio, J.D.; Gulen, M.F.; Saidoune, F.; Thacker, V.V.; Yatim, A.; Sharma, K.; Nass, T.; Guenova, E.; Schaller, M.; Conrad, C.; et al. The cGAS-STING pathway drives type I IFN immunopathology in COVID-19. Nature 2022, 603, 145–151. [Google Scholar] [CrossRef]
  114. Ma, Y.; Wang, X.; Luo, W.; Xiao, J.; Song, X.; Wang, Y.; Shuai, H.; Ren, Z.; Wang, Y. Roles of Emerging RNA-Binding Activity of cGAS in Innate Antiviral Response. Front. Immunol. 2021, 12, 741599. [Google Scholar] [CrossRef]
  115. Yoh, S.M.; Mamede, J.I.; Lau, D.; Ahn, N.; Sánchez-Aparicio, M.T.; Temple, J.; Tuckwell, A.; Fuchs, N.V.; Cianci, G.C.; Riva, L.; et al. Recognition of HIV-1 capsid by PQBP1 licenses an innate immune sensing of nascent HIV-1 DNA. Mol. Cell 2022, 82, 2871–2884.e6. [Google Scholar] [CrossRef]
  116. Yoh, S.M.; Schneider, M.; Seifried, J.; Soonthornvacharin, S.; Akleh, R.E.; Olivieri, K.C.; De Jesus, P.D.; Ruan, C.; de Castro, E.; Ruiz, P.A.; et al. PQBP1 Is a Proximal Sensor of the cGAS-Dependent Innate Response to HIV-1. Cell 2015, 161, 1293–1305. [Google Scholar] [CrossRef]
  117. Watson, R.O.; Bell, S.L.; MacDuff, D.A.; Kimmey, J.M.; Diner, E.J.; Olivas, J.; Vance, R.E.; Stallings, C.L.; Virgin, H.W.; Cox, J.S. The Cytosolic Sensor cGAS Detects Mycobacterium tuberculosis DNA to Induce Type I Interferons and Activate Autophagy. Cell Host Microbe 2015, 17, 811–819. [Google Scholar] [CrossRef]
  118. Whiteley, L.; Meffert, T.; Haug, M.; Weidenmaier, C.; Hopf, V.; Bitschar, K.; Schittek, B.; Kohler, C.; Steinmetz, I.; West, T.E.; et al. Entry, Intracellular Survival, and Multinucleated-Giant-Cell-Forming Activity of Burkholderia pseudomallei in Human Primary Phagocytic and Nonphagocytic Cells. Infect. Immun. 2017, 85. [Google Scholar] [CrossRef]
  119. Ku, J.W.K.; Chen, Y.; Lim, B.J.W.; Gasser, S.; Crasta, K.C.; Gan, Y.H. Bacterial-induced cell fusion is a danger signal triggering cGAS-STING pathway via micronuclei formation. Proc. Natl. Acad. Sci. USA 2020, 117, 15923–15934. [Google Scholar] [CrossRef]
  120. Place, D.E.; Briard, B.; Samir, P.; Karki, R.; Bhattacharya, A.; Guy, C.S.; Peters, J.L.; Frase, S.; Vogel, P.; Neale, G.; et al. Interferon inducible GBPs restrict Burkholderia thailandensis motility induced cell-cell fusion. PLOS Pathog. 2020, 16, e1008364. [Google Scholar] [CrossRef]
  121. Liu, N.; Pang, X.; Zhang, H.; Ji, P. The cGAS-STING Pathway in Bacterial Infection and Bacterial Immunity. Front. Immunol. 2021, 12, 814709. [Google Scholar] [PubMed]
  122. Ahn, J.; Barber, G.N. STING signaling and host defense against microbial infection. Exp. Mol. Med. 2019, 51, 1–10. [Google Scholar] [CrossRef] [PubMed]
  123. Hahn, W.O.; Butler, N.S.; Lindner, S.E.; Akilesh, H.M.; Sather, D.N.; Kappe, S.H.; Hamerman, J.A.; Gale, M., Jr.; Liles, W.C.; Pepper, M. cGAS-mediated control of blood-stage malaria promotes Plasmodium-specific germinal center responses. JCI Insight 2018, 3, e94142. [Google Scholar] [CrossRef] [PubMed]
  124. Sun, Y.; Cheng, Y. STING or Sting: cGAS-STING-Mediated Immune Response to Protozoan Parasites. Trends Parasitol. 2020, 36, 773–784. [Google Scholar] [CrossRef] [PubMed]
  125. Wang, P.; Li, S.; Zhao, Y.; Zhang, B.; Li, Y.; Liu, S.; Du, H.; Cao, L.; Ou, M.; Ye, X.; et al. The GRA15 protein from Toxoplasma gondii enhances host defense responses by activating the interferon stimulator STING. J. Biol. Chem. 2019, 294, 16494–16508. [Google Scholar] [CrossRef] [PubMed]
  126. Zhang, J.; Zhao, J.; Xu, S.; Li, J.; He, S.; Zeng, Y.; Xie, L.; Xie, N.; Liu, T.; Lee, K.; et al. Faculty Opinions recommendation of Species-Specific Deamidation of cGAS by Herpes Simplex Virus UL37 Protein Facilitates Viral Replication. Cell Host Microbe 2018, 24, 234–248.e5. [Google Scholar] [CrossRef] [PubMed]
  127. Huang, J.; You, H.; Su, C.; Li, Y.; Chen, S.; Zheng, C. Herpes Simplex Virus 1 Tegument Protein VP22 Abrogates cGAS/STING-Mediated Antiviral Innate Immunity. J. Virol. 2018, 92, 10–1128. [Google Scholar] [CrossRef]
  128. Pan, S.; Liu, X.; Ma, Y.; Cao, Y.; He, B. Herpes Simplex Virus 1 γ1 34.5 Protein Inhibits STING Activation That Restricts Viral Replication. J. Virol. 2018, 92, e01015-18. [Google Scholar] [CrossRef]
  129. Wu, J.-J.; Li, W.; Shao, Y.; Avey, D.; Fu, B.; Gillen, J.; Hand, T.; Ma, S.; Liu, X.; Miley, W.; et al. Inhibition of cGAS DNA Sensing by a Herpesvirus Virion Protein. Cell Host Microbe 2015, 18, 333–344. [Google Scholar] [CrossRef]
  130. Ma, Z.; Jacobs, S.R.; West, J.A.; Stopford, C.; Zhang, Z.; Davis, Z.; Barber, G.N.; Glaunsinger, B.A.; Dittmer, D.P.; Damania, B. Modulation of the cGAS-STING DNA sensing pathway by gammaherpesviruses. Proc. Natl. Acad. Sci. USA 2015, 112, E4306–E4315. [Google Scholar] [CrossRef]
  131. Xing, Y.; Chen, J.; Tu, J.; Zhang, B.; Chen, X.; Shi, H.; Baker, S.C.; Feng, L.; Chen, Z. The papain-like protease of porcine epidemic diarrhea virus negatively regulates type I interferon pathway by acting as a viral deubiquitinase. J. Gen. Virol. 2013, 94, 1554–1567. [Google Scholar] [CrossRef] [PubMed]
  132. Sun, L.; Xing, Y.; Chen, X.; Zheng, Y.; Yang, Y.; Nichols, D.B.; Clementz, M.A.; Banach, B.S.; Li, K.; Baker, S.C.; et al. Coronavirus Papain-like Proteases Negatively Regulate Antiviral Innate Immune Response through Disruption of STING-Mediated Signaling. PLoS ONE 2012, 7, e30802. [Google Scholar] [CrossRef] [PubMed]
  133. Ding, Q.; Cao, X.; Lu, J.; Huang, B.; Liu, Y.-J.; Kato, N.; Shu, H.-B.; Zhong, J. Hepatitis C virus NS4B blocks the interaction of STING and TBK1 to evade host innate immunity. J. Hepatol. 2013, 59, 52–58. [Google Scholar] [CrossRef] [PubMed]
  134. Aguirre, S.; Luthra, P.; Sanchez-Aparicio, M.T.; Maestre, A.M.; Patel, J.; Lamothe, F.; Fredericks, A.C.; Tripathi, S.; Zhu, T.; Pintado-Silva, J.; et al. Dengue virus NS2B protein targets cGAS for degradation and prevents mitochondrial DNA sensing during infection. Nat. Microbiol. 2017, 2, 17037. [Google Scholar] [CrossRef] [PubMed]
  135. Zheng, Y.; Liu, Q.; Wu, Y.; Ma, L.; Zhang, Z.; Liu, T.; Jin, S.; She, Y.; Li, Y.-P.; Cui, J. Zika virus elicits inflammation to evade antiviral response by cleaving cGAS via NS 1-caspase-1 axis. EMBO J. 2018, 37, e99347. [Google Scholar] [CrossRef]
  136. Lau, L.; Gray, E.E.; Brunette, R.L.; Stetson, D.B. DNA tumor virus oncogenes antagonize the cGAS-STING DNA-sensing pathway. Science 2015, 350, 568–571. [Google Scholar] [CrossRef]
  137. Liu, Y.; Li, J.; Chen, J.; Li, Y.; Wang, W.; Du, X.; Song, W.; Zhang, W.; Lin, L.; Yuan, Z. Hepatitis B Virus Polymerase Disrupts K63-Linked Ubiquitination of STING To Block Innate Cytosolic DNA-Sensing Pathways. J. Virol. 2015, 89, 2287–2300. [Google Scholar] [CrossRef]
  138. Eaglesham, J.B.; Pan, Y.; Kupper, T.S.; Kranzusch, P.J. Viral and metazoan poxins are cGAMP-specific nucleases that restrict cGAS-STING signalling. Nature 2019, 566, 259–263. [Google Scholar] [CrossRef]
  139. Su, J.; Rui, Y.; Lou, M.; Yin, L.; Xiong, H.; Zhou, Z.; Shen, S.; Chen, T.; Zhang, Z.; Zhao, N.; et al. HIV-2/SIV Vpx targets a novel functional domain of STING to selectively inhibit cGAS-STING-mediated NF-κB signalling. Nat. Microbiol. 2019, 4, 2552–2564. [Google Scholar]
  140. Andrade, W.A.; Firon, A.; Schmidt, T.; Hornung, V.; Fitzgerald, K.A.; Kurt-Jones, E.A.; Trieu-Cuot, P.; Golenbock, D.T.; Kaminski, P.-A. Group B Streptococcus Degrades Cyclic-di-AMP to Modulate STING-Dependent Type I Interferon Production. Cell Host Microbe 2016, 20, 49–59. [Google Scholar] [CrossRef]
  141. Sixt, B.S.; Bastidas, R.J.; Finethy, R.; Baxter, R.M.; Carpenter, V.K.; Kroemer, G.; Coers, J.; Valdivia, R.H. The Chlamydia trachomatis Inclusion Membrane Protein CpoS Counteracts STING-Mediated Cellular Surveillance and Suicide Programs. Cell Host Microbe 2016, 21, 113–121. [Google Scholar] [CrossRef] [PubMed]
  142. Cao, Y.; Guan, K.; He, X.; Wei, C.; Zheng, Z.; Zhang, Y.; Ma, S.; Zhong, H.; Shi, W. Yersinia YopJ negatively regulates IRF3-mediated antibacterial response through disruption of STING-mediated cytosolic DNA signaling. Biochim. et Biophys. Acta (BBA) Mol. Cell Res. 2016, 1863, 3148–3159. [Google Scholar] [CrossRef]
  143. Köster, S.; Upadhyay, S.; Chandra, P.; Papavinasasundaram, K.; Yang, G.; Hassan, A.; Grigsby, S.J.; Mittal, E.; Park, H.S.; Jones, V.; et al. Mycobacterium tuberculosis is protected from NADPH oxidase and LC3-associated phagocytosis by the LCP protein CpsA. Proc. Natl. Acad. Sci. USA 2017, 114, E8711–E8720. [Google Scholar] [CrossRef] [PubMed]
  144. Lee-Kirsch, M.A. The Type I Interferonopathies. Annu. Rev. Med. 2017, 68, 297–315. [Google Scholar] [CrossRef] [PubMed]
  145. Liu, Y.; Jesus, A.A.; Marrero, B.; Yang, D.; Ramsey, S.E.; Montealegre Sanchez, G.A.; Tenbrock, K.; Wittkowski, H.; Jones, O.Y.; Kuehn, H.S.; et al. Activated STING in a Vascular and Pulmonary Syndrome. N. Engl. J. Med. 2014, 371, 507–518. [Google Scholar] [CrossRef]
  146. Vece, T.J.; Watkin, L.B.; Nicholas, S.K.; Canter, D.; Braun, M.C.; Guillerman, R.P.; Eldin, K.W.; Bertolet, G.; McKinley, S.D.; de Guzman, M.; et al. Copa Syndrome: A Novel Autosomal Dominant Immune Dysregulatory Disease. J. Clin. Immunol. 2016, 36, 377–387. [Google Scholar] [CrossRef]
  147. Deng, Z.; Chong, Z.; Law, C.S.; Mukai, K.; Ho, F.O.; Martinu, T.; Backes, B.J.; Eckalbar, W.L.; Taguchi, T.; Shum, A.K. A defect in COPI-mediated transport of STING causes immune dysregulation in COPA syndrome. J. Exp. Med. 2020, 217, e20201045. [Google Scholar] [CrossRef]
  148. Lepelley, A.; Martin-Niclós, M.J.; Le Bihan, M.; Marsh, J.A.; Uggenti, C.; Rice, G.I.; Bondet, V.; Duffy, D.; Hertzog, J.; Rehwinkel, J.; et al. Mutations in COPA lead to abnormal trafficking of STING to the Golgi and interferon signaling. J. Exp. Med. 2020, 217, e20200600. [Google Scholar] [CrossRef]
  149. Lee-Kirsch, M.A.; Gong, M.; Schulz, H.; Rüschendorf, F.; Stein, A.; Pfeiffer, C.; Ballarini, A.; Gahr, M.; Hubner, N.; Linné, M. Familial Chilblain Lupus, a Monogenic Form of Cutaneous Lupus Erythematosus, Maps to Chromosome 3p. Am. J. Hum. Genet. 2006, 79, 731–737. [Google Scholar] [CrossRef]
  150. König, N.; Fiehn, C.; Wolf, C.; Schuster, M.; Costa, E.C.; Tüngler, V.; Alvarez, H.A.; Chara, O.; Engel, K.; Goldbach-Mansky, R.; et al. Familial chilblain lupus due to a gain-of-function mutation in STING. Ann. Rheum. Dis. 2016, 76, 468–472. [Google Scholar] [CrossRef]
  151. Rahman, A.; Isenberg, D.A. Systemic lupus erythematosus. N. Engl. J. Med. 2008, 358, 929–939. [Google Scholar] [CrossRef] [PubMed]
  152. Günther, C.; Kind, B.; Reijns, M.A.; Berndt, N.; Martinez-Bueno, M.; Wolf, C.; Tüngler, V.; Chara, O.; Lee, Y.A.; Hübner, N.; et al. Defective removal of ribonucleotides from DNA promotes systemic autoimmunity. J. Clin. Investig. 2014, 125, 413–424. [Google Scholar] [CrossRef] [PubMed]
  153. Yasutomo, K.; Horiuchi, T.; Kagami, S.; Tsukamoto, H.; Hashimura, C.; Urushihara, M.; Kuroda, Y. Mutation of DNASE1 in people with systemic lupus erythematosus. Nat. Genet. 2001, 28, 313–314. [Google Scholar] [CrossRef]
  154. Kato, Y.; Park, J.; Takamatsu, H.; Konaka, H.; Aoki, W.; Aburaya, S.; Ueda, M.; Nishide, M.; Koyama, S.; Hayama, Y.; et al. Apoptosis-derived membrane vesicles drive the cGAS-STING pathway and enhance type I IFN production in systemic lupus erythematosus. Ann. Rheum. Dis. 2018, 77, 1507–1515. [Google Scholar]
  155. Crow, Y.J.; Shetty, J.; Livingston, J.H. Treatments in Aicardi-Goutières syndrome. Dev. Med. Child Neurol. 2020, 62, 42–47. [Google Scholar] [CrossRef]
  156. Naesens, L.; Nemegeer, J.; Roelens, F.; Vallaeys, L.; Meuwissen, M.; Janssens, K.; Verloo, P.; Ogunjimi, B.; Hemelsoet, D. Mutations in RNU7-1 Weaken Secondary RNA Structure, Induce MCP-1 and CXCL10 in CSF, and Result in Aicardi-Goutières Syndrome with Severe End-Organ Involvement. J. Clin. Immunol. 2022, 42, 962–974. [Google Scholar] [CrossRef]
  157. Rice, G.I.; Forte, G.M.; Szynkiewicz, M.; Chase, D.S.; Aeby, A.; Abdel-Hamid, M.S.; Ackroyd, S.; Allcock, R.; Bailey, K.M.; Balottin, U.; et al. Assessment of interferon-related biomarkers in Aicardi-Goutières syndrome associated with mutations in TREX1, RNASEH2A, RNASEH2B, RNASEH2C, SAMHD1, and ADAR: A case-control study. Lancet Neurol. 2013, 12, 1159–1169. [Google Scholar] [CrossRef] [PubMed]
  158. Pokatayev, V.; Hasin, N.; Chon, H.; Cerritelli, S.M.; Sakhuja, K.; Ward, J.M.; Morris, H.D.; Yan, N.; Crouch, R.J. RNase H2 catalytic core Aicardi-Goutières syndrome-related mutant invokes cGAS-STING innate immune-sensing pathway in mice. J. Exp. Med. 2016, 213, 329–336. [Google Scholar] [CrossRef]
  159. Mackenzie, K.J.; Carroll, P.; Lettice, L.; Tarnauskaitė, Ž.; Reddy, K.; Dix, F.; Revuelta, A.; Abbondati, E.; Rigby, R.E.; Rabe, B.; et al. Ribonuclease H2 mutations induce a cGAS/STING -dependent innate immune response. EMBO J. 2016, 35, 831–844. [Google Scholar] [CrossRef]
  160. Scott, D.L.; Wolfe, F.; Huizinga, T.W. Rheumatoid arthritis. Lancet 2010, 376, 1094–1108. [Google Scholar] [CrossRef]
  161. Wang, J.; Li, R.; Lin, H.; Qiu, Q.; Lao, M.; Zeng, S.; Wang, C.; Xu, S.; Zou, Y.; Shi, M.; et al. Accumulation of cytosolic dsDNA contributes to fibroblast-like synoviocytes-mediated rheumatoid arthritis synovial inflammation. Int. Immunopharmacol. 2019, 76, 105791. [Google Scholar] [CrossRef] [PubMed]
  162. Wang, W.; Arakawa, H.; Wang, L.; Okolo, O.; Siedlak, S.L.; Jiang, Y.; Gao, J.; Xie, F.; Petersen, R.B.; Wang, X. Motor-Coordinative and Cognitive Dysfunction Caused by Mutant TDP-43 Could Be Reversed by Inhibiting Its Mitochondrial Localization. Mol. Ther. 2017, 25, 127–139. [Google Scholar] [CrossRef] [PubMed]
  163. Yu, C.-H.; Davidson, S.; Harapas, C.R.; Hilton, J.B.; Mlodzianoski, M.J.; Laohamonthonkul, P.; Louis, C.; Low, R.R.J.; Moecking, J.; De Nardo, D.; et al. TDP-43 Triggers Mitochondrial DNA Release via mPTP to Activate cGAS/STING in ALS. Cell 2020, 183, 636–649.e18. [Google Scholar] [CrossRef] [PubMed]
  164. Crotti, A.; Glass, C.K. The choreography of neuroinflammation in Huntington's disease. Trends Immunol. 2015, 36, 364–373. [Google Scholar]
  165. Sharma, M.; Rajendrarao, S.; Shahani, N.; Ramírez-Jarquín, U.N.; Subramaniam, S. Cyclic GMP-AMP synthase promotes the inflammatory and autophagy responses in Huntington disease. Proc. Natl. Acad. Sci. USA 2020, 117, 15989–15999. [Google Scholar] [CrossRef]
  166. Wang, D.; Zhao, H.; Shen, Y.; Chen, Q. A Variety of Nucleic Acid Species Are Sensed by cGAS, Implications for Its Diverse Functions. Front. Immunol. 2022, 13, 826880. [Google Scholar] [CrossRef]
  167. Cao, D.J.; Schiattarella, G.G.; Villalobos, E.; Jiang, N.; May, H.I.; Li, T.; Chen, Z.J.; Gillette, T.G.; Hill, J.A. Cytosolic DNA Sensing Promotes Macrophage Transformation and Governs Myocardial Ischemic Injury. Circulation 2018, 137, 2613–2634. [Google Scholar] [CrossRef]
  168. Lankisch, P.G.; Apte, M.; Banks, P.A. Acute pancreatitis. Lancet 2015, 386, 85–96. [Google Scholar] [CrossRef]
  169. Zhao, Q.; Wei, Y.; Pandol, S.J.; Li, L.; Habtezion, A. STING Signaling Promotes Inflammation in Experimental Acute Pancreatitis. Gastroenterology 2018, 154, 1822–1835.e2. [Google Scholar] [CrossRef]
  170. Leung, C.C.; Yu, I.T.; Chen, W. Silicosis. Lancet 2012, 379, 2008–2018. [Google Scholar]
  171. Benmerzoug, S.; Rose, S.; Bounab, B.; Gosset, D.; Duneau, L.; Chenuet, P.; Mollet, L.; Le Bert, M.; Lambers, C.; Geleff, S.; et al. STING-dependent sensing of self-DNA drives silica-induced lung inflammation. Nat. Commun. 2018, 9, 5226. [Google Scholar] [CrossRef]
  172. Mao, Y.; Luo, W.; Zhang, L.; Wu, W.; Yuan, L.; Xu, H.; Song, J.; Fujiwara, K.; Abe, J.-I.; Lemaire, S.A.; et al. STING–IRF3 Triggers Endothelial Inflammation in Response to Free Fatty Acid-Induced Mitochondrial Damage in Diet-Induced Obesity. Arterioscler. Thromb. Vasc. Biol. 2017, 37, 920–929. [Google Scholar] [CrossRef] [PubMed]
  173. Yu, Y.; Liu, Y.; An, W.; Song, J.; Zhang, Y.; Zhao, X. STING-mediated inflammation in Kupffer cells contributes to progression of nonalcoholic steatohepatitis. J. Clin. Investig. 2019, 129, 546–555. [Google Scholar] [CrossRef] [PubMed]
  174. Xu, D.; Tian, Y.; Xia, Q.; Ke, B. The cGAS-STING Pathway: Novel Perspectives in Liver Diseases. Front. Immunol. 2021, 12, 682736. [Google Scholar] [CrossRef] [PubMed]
  175. Zhong, W.; Rao, Z.; Rao, J.; Han, G.; Wang, P.; Jiang, T.; Pan, X.; Zhou, S.; Zhou, H.; Wang, X. Aging aggravated liver ischemia and reperfusion injury by promoting STING-mediated NLRP3 activation in macrophages. Aging Cell 2020, 19, e13186. [Google Scholar] [CrossRef] [PubMed]
  176. Gulen, M.F.; Samson, N.; Keller, A.; Schwabenland, M.; Liu, C.; Glück, S.; Thacker, V.V.; Favre, L.; Mangeat, B.; Kroese, L.J.; et al. cGAS-STING drives ageing-related inflammation and neurodegeneration. Nature 2023, 620, 374–380. [Google Scholar]
  177. Dou, Z.; Ghosh, K.; Vizioli, M.G.; Zhu, J.; Sen, P.; Wangensteen, K.J.; Simithy, J.; Lan, Y.; Lin, Y.; Zhou, Z.; et al. Cytoplasmic chromatin triggers inflammation in senescence and cancer. Nature 2017, 550, 402–406. [Google Scholar] [CrossRef]
  178. Chabanon, R.M.; Muirhead, G.; Krastev, D.B.; Adam, J.; Morel, D.; Garrido, M.; Lamb, A.; Hénon, C.; Dorvault, N.; Rouanne, M.; et al. PARP inhibition enhances tumor cell–intrinsic immunity in ERCC1-deficient non–small cell lung cancer. J. Clin. Investig. 2019, 129, 1211–1228. [Google Scholar] [CrossRef]
  179. Baird, J.R.; Friedman, D.; Cottam, B.; Dubensky, T.W.; Kanne, D.B.; Bambina, S.; Bahjat, K.; Crittenden, M.R.; Gough, M.J. Radiotherapy Combined with Novel STING-Targeting Oligonucleotides Results in Regression of Established Tumors. Cancer Res 2016, 76, 50–61. [Google Scholar] [CrossRef]
  180. Dvorak, H.F. Tumors: Wounds that do not heal. Similarities between tumor stroma generation and wound healing. N. Engl. J. Med. 1986, 315, 1650–1659. [Google Scholar]
  181. Kwon, J.; Bakhoum, S.F. The Cytosolic DNA-Sensing cGAS-STING Pathway in Cancer. Cancer Discov. 2020, 10, 26–39. [Google Scholar] [CrossRef] [PubMed]
  182. Crasta, K.; Ganem, N.J.; Dagher, R.; Lantermann, A.B.; Ivanova, E.V.; Pan, Y.; Nezi, L.; Protopopov, A.; Chowdhury, D.; Pellman, D. DNA breaks and chromosome pulverization from errors in mitosis. Nature 2012, 482, 53–58. [Google Scholar] [CrossRef]
  183. Riley, J.S.; Quarato, G.; Cloix, C.; Lopez, J.; O'Prey, J.; Pearson, M.; Chapman, J.; Sesaki, H.; Carlin, L.M.; Passos, J.F.; et al. Mitochondrial inner membrane permeabilisation enables mtDNA release during apoptosis. EMBO J. 2018, 37, e99238. [Google Scholar] [CrossRef] [PubMed]
  184. Khoo, L.T.; Chen, L.Y. Role of the cGAS-STING pathway in cancer development and oncotherapeutic approaches. EMBO Rep. 2018, 19, e46935. [Google Scholar] [CrossRef] [PubMed]
  185. Diamond, J.M.; Vanpouille-Box, C.; Spada, S.; Rudqvist, N.-P.; Chapman, J.R.; Ueberheide, B.M.; Pilones, K.A.; Sarfraz, Y.; Formenti, S.C.; Demaria, S. Exosomes Shuttle TREX1-Sensitive IFN-Stimulatory dsDNA from Irradiated Cancer Cells to DCs. Cancer Immunol. Res. 2018, 6, 910–920. [Google Scholar] [CrossRef]
  186. Schadt, L.; Sparano, C.; Schweiger, N.A.; Silina, K.; Cecconi, V.; Lucchiari, G.; Yagita, H.; Guggisberg, E.; Saba, S.; Nascakova, Z.; et al. Cancer-Cell-Intrinsic cGAS Expression Mediates Tumor Immunogenicity. Cell Rep. 2019, 29, 1236–1248.e7. [Google Scholar] [CrossRef] [PubMed]
  187. Pépin, G.; Gantier, M.P. cGAS-STING Activation in the Tumor Microenvironment and Its Role in Cancer Immunity. Adv. Exp. Med. Biol. 2017, 1024, 175–194. [Google Scholar]
  188. Marcus, A.; Mao, A.J.; Lensink-Vasan, M.; Wang, L.; Vance, R.E.; Raulet, D.H. Tumor-Derived cGAMP Triggers a STING-Mediated Interferon Response in Non-tumor Cells to Activate the NK Cell Response. Immunity 2018, 49, 754–763.e4. [Google Scholar] [CrossRef]
  189. Lam, A.R.; Le Bert, N.; Ho, S.S.W.; Shen, Y.J.; Tang, M.L.F.; Xiong, G.M.; Croxford, J.L.; Koo, C.X.; Ishii, K.J.; Akira, S.; et al. RAE1 Ligands for the NKG2D Receptor Are Regulated by STING-Dependent DNA Sensor Pathways in Lymphoma. Cancer Res. 2014, 74, 2193–2203. [Google Scholar] [CrossRef]
  190. Chen, Q.; Boire, A.; Jin, X.; Valiente, M.; Er, E.E.; Lopez-Soto, A.; Jacob, L.S.; Patwa, R.; Shah, H.; Xu, K.; et al. Carcinoma–astrocyte gap junctions promote brain metastasis by cGAMP transfer. Nature 2016, 533, 493–498. [Google Scholar] [CrossRef]
  191. Ahn, J.; Xia, T.; Konno, H.; Konno, K.; Ruiz, P.; Barber, G.N. Inflammation-driven carcinogenesis is mediated through STING. Nat. Commun. 2014, 5, 5166. [Google Scholar] [CrossRef] [PubMed]
  192. Liu, H.; Zhang, H.; Wu, X.; Ma, D.; Wu, J.; Wang, L.; Jiang, Y.; Fei, Y.; Zhu, C.; Tan, R.; et al. Nuclear cGAS suppresses DNA repair and promotes tumorigenesis. Nature 2018, 563, 131–136. [Google Scholar] [CrossRef] [PubMed]
  193. Konno, H.; Yamauchi, S.; Berglund, A.; Putney, R.M.; Mulé, J.J.; Barber, G.N. Suppression of STING signaling through epigenetic silencing and missense mutation impedes DNA damage mediated cytokine production. Oncogene 2018, 37, 2037–2051. [Google Scholar] [CrossRef] [PubMed]
  194. Caronni, N.; Simoncello, F.; Stafetta, F.; Guarnaccia, C.; Ruiz-Moreno, J.S.; Opitz, B.; Galli, T.; Proux-Gillardeaux, V.; Benvenuti, F. Downregulation of Membrane Trafficking Proteins and Lactate Conditioning Determine Loss of Dendritic Cell Function in Lung Cancer. Cancer Res 2018, 78, 1685–1699. [Google Scholar] [CrossRef]
  195. Wu, S.; Zhang, Q.; Zhang, F.; Meng, F.; Liu, S.; Zhou, R.; Wu, Q.; Li, X.; Shen, L.; Huang, J.; et al. HER2 recruits AKT1 to disrupt STING signalling and suppress antiviral defence and antitumour immunity. Nat. Cell Biol. 2019, 21, 1027–1040. [Google Scholar] [CrossRef] [PubMed]
  196. De Vitis, M.; Berardinelli, F.; Sgura, A. Telomere Length Maintenance in Cancer: At the Crossroad between Telomerase and Alternative Lengthening of Telomeres (ALT). Int. J. Mol. Sci. 2018, 19, 606. [Google Scholar] [CrossRef]
  197. Chen, Y.A.; Shen, Y.L.; Hsia, H.Y.; Tiang, Y.P.; Sung, T.L.; Chen, L.Y. Extrachromosomal telomere repeat DNA is linked to ALT development via cGAS-STING DNA sensing pathway. Nat. Struct. Mol. Biol. 2017, 24, 1124–1131. [Google Scholar]
  198. Sivick, K.E.; Desbien, A.L.; Glickman, L.H.; Reiner, G.L.; Corrales, L.; Surh, N.H.; Hudson, T.E.; Vu, U.T.; Francica, B.J.; Banda, T.; et al. Magnitude of Therapeutic STING Activation Determines CD8+ T Cell-Mediated Anti-tumor Immunity. Cell Rep. 2018, 25, 3074–3085.e5. [Google Scholar] [CrossRef]
  199. Da Hoong, B.Y.; Gan, Y.H.; Liu, H.; Chen, E.S. cGAS-STING pathway in oncogenesis and cancer therapeutics. Oncotarget 2020, 11, 2930–2955. [Google Scholar] [CrossRef]
  200. Zheng, J.; Mo, J.; Zhu, T.; Zhuo, W.; Yi, Y.; Hu, S.; Yin, J.; Zhang, W.; Zhou, H.; Liu, Z. Comprehensive elaboration of the cGAS-STING signaling axis in cancer development and immunotherapy. Mol. Cancer 2020, 19, 133. [Google Scholar] [CrossRef]
  201. Hall, J.; Brault, A.; Vincent, F.; Weng, S.; Wang, H.; Dumlao, D.; Aulabaugh, A.; Aivazian, D.; Castro, D.; Chen, M.; et al. Discovery of PF-06928215 as a high affinity inhibitor of cGAS enabled by a novel fluorescence polarization assay. PLoS ONE 2017, 12, e0184843. [Google Scholar] [CrossRef] [PubMed]
  202. Hall, J.; Ralph, E.C.; Shanker, S.; Wang, H.; Byrnes, L.J.; Horst, R.; Wong, J.; Brault, A.; Dumlao, D.; Smith, J.F.; et al. The catalytic mechanism of cyclic GMP-AMP synthase (cGAS) and implications for innate immunity and inhibition. Protein Sci. 2017, 26, 2367–2380. [Google Scholar] [CrossRef] [PubMed]
  203. Vincent, J.; Adura, C.; Gao, P.; Luz, A.; Lama, L.; Asano, Y.; Okamoto, R.; Imaeda, T.; Aida, J.; Rothamel, K.; et al. Small molecule inhibition of cGAS reduces interferon expression in primary macrophages from autoimmune mice. Nat. Commun. 2017, 8, 750. [Google Scholar] [CrossRef]
  204. Lama, L.; Adura, C.; Xie, W.; Tomita, D.; Kamei, T.; Kuryavyi, V.; Gogakos, T.; Steinberg, J.I.; Miller, M.; Ramos-Espiritu, L.; et al. Development of human cGAS-specific small-molecule inhibitors for repression of dsDNA-triggered interferon expression. Nat. Commun. 2019, 10, 2261. [Google Scholar] [CrossRef]
  205. Padilla-Salinas, R.; Sun, L.; Anderson, R.; Yang, X.; Zhang, S.; Chen, Z.J.; Yin, H.H. Discovery of Small-Molecule Cyclic GMP-AMP Synthase Inhibitors. J. Org. Chem. 2020, 85, 1579–1600. [Google Scholar] [CrossRef] [PubMed]
  206. Lai, J.; Luo, X.; Tian, S.; Zhang, X.; Huang, S.; Wang, H.; Li, Q.; Cai, S.; Chen, Q. Compound C Reducing Interferon Expression by Inhibiting cGAMP Accumulation. Front. Pharmacol. 2020, 11, 88. [Google Scholar] [CrossRef]
  207. An, J.; Woodward, J.J.; Sasaki, T.; Minie, M.; Elkon, K.B. Cutting Edge: Antimalarial Drugs Inhibit IFN-β Production through Blockade of Cyclic GMP-AMP Synthase–DNA Interaction. J. Immunol. 2015, 194, 4089–4093. [Google Scholar] [CrossRef]
  208. Bose, D.; Su, Y.; Marcus, A.; Raulet, D.H.; Hammond, M.C. An RNA-Based Fluorescent Biosensor for High-Throughput Analysis of the cGAS-cGAMP-STING Pathway. Cell Chem. Biol. 2016, 23, 1539–1549. [Google Scholar] [CrossRef]
  209. Wang, M.; Sooreshjani, M.A.; Mikek, C.; Opoku-Temeng, C.; Sintim, H.O. Suramin potently inhibits cGAMP synthase, cGAS, in THP1 cells to modulate IFN-β levels. Future Med. Chem. 2018, 10, 1301–1317. [Google Scholar] [CrossRef]
  210. Steinhagen, F.; Zillinger, T.; Peukert, K.; Fox, M.; Thudium, M.; Barchet, W.; Putensen, C.; Klinman, D.; Latz, E.; Bode, C. Suppressive oligodeoxynucleotides containing TTAGGG motifs inhibit cGAS activation in human monocytes. Eur. J. Immunol. 2018, 48, 605–611. [Google Scholar] [CrossRef]
  211. Li, S.; Hong, Z.; Wang, Z.; Li, F.; Mei, J.; Huang, L.; Lou, X.; Zhao, S.; Song, L.; Chen, W.; et al. The Cyclopeptide Astin C Specifically Inhibits the Innate Immune CDN Sensor STING. Cell Rep. 2018, 25, 3405–3421.e7. [Google Scholar] [CrossRef] [PubMed]
  212. Siu, T.; Altman, M.D.; Baltus, G.A.; Childers, M.; Ellis, J.; 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]
  213. Haag, S.M.; Gulen, M.F.; Reymond, L.; Gibelin, A.; Abrami, L.; Decout, A.; Heymann, M.; van der Goot, F.G.; Turcatti, G.; Behrendt, R.; et al. Targeting STING with covalent small-molecule inhibitors. Nature 2018, 559, 269–273. [Google Scholar] [CrossRef] [PubMed]
  214. Hansen, A.L.; Buchan, G.J.; Rühl, M.; Mukai, K.; Salvatore, S.R.; Ogawa, E.; Andersen, S.D.; Iversen, M.B.; Thielke, A.L.; Gunderstofte, C.; et al. Nitro-fatty acids are formed in response to virus infection and are potent inhibitors of STING palmitoylation and signaling. Proc. Natl. Acad. Sci. USA 2018, 115, E7768–E7775. [Google Scholar] [CrossRef] [PubMed]
  215. Corrales, L.; Glickman, L.H.; McWhirter, S.M.; Kanne, D.B.; Sivick, K.E.; Katibah, G.E.; Woo, S.-R.; Lemmens, 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] [PubMed]
  216. Ohkuri, T.; Kosaka, A.; Ishibashi, K.; Kumai, T.; Hirata, Y.; Ohara, K.; Nagato, T.; Oikawa, K.; Aoki, N.; Harabuchi, Y.; et al. Intratumoral administration of cGAMP transiently accumulates potent macrophages for anti-tumor immunity at a mouse tumor site. Cancer Immunol. Immunother. 2017, 66, 705–716. [Google Scholar] [CrossRef]
  217. Kim, S.; Li, L.; Maliga, Z.; Yin, Q.; Wu, H.; Mitchison, T.J. Anticancer Flavonoids Are Mouse-Selective STING Agonists. ACS Chem. Biol. 2013, 8, 1396–1401. [Google Scholar] [CrossRef]
  218. Lai, J.; Fu, Y.; Tian, S.; Huang, S.; Luo, X.; Lin, L.; Zhang, X.; Wang, H.; Lin, Z.; Zhao, H.; et al. Zebularine elevates STING expression and enhances cGAMP cancer immunotherapy in mice. Mol. Ther. 2021, 29, 1758–1771. [Google Scholar] [CrossRef]
  219. Banerjee, M.; Middya, S.; Shrivastava, R.; Basu, S.; Ghosh, R.; Pryde, D.C.; Yadav, D.B.; Surya, A. G10 is a direct activator of human STING. PLoS ONE 2020, 15, e0237743. [Google Scholar] [CrossRef]
  220. 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]
  221. Liu, B.; Tang, L.; Zhang, X.; Ma, J.; Sehgal, M.; Cheng, J.; Zhang, X.; Zhou, Y.; Du, Y.; Kulp, J.; et al. A cell-based high throughput screening assay for the discovery of cGAS-STING pathway agonists. Antiviral Res. 2017, 147, 37–46. [Google Scholar] [CrossRef] [PubMed]
  222. Zhang, X.; Liu, B.; Tang, L.; Su, Q.; Hwang, N.; Sehgal, M.; Cheng, J.; Ma, J.; Zhang, X.; Tan, Y.; et al. Discovery and Mechanistic Study of a Novel Human-Stimulator-of-Interferon-Genes Agonist. ACS Infect. Dis. 2019, 5, 1139–1149. [Google Scholar] [CrossRef] [PubMed]
  223. Deng, L.; Liang, H.; Xu, M.; Yang, X.; Burnette, B.; Arina, A.; Li, X.-D.; Mauceri, H.; Beckett, M.; Darga, T.; et al. STING-Dependent Cytosolic DNA Sensing Promotes Radiation-Induced Type I Interferon-Dependent Antitumor Immunity in Immunogenic Tumors. Immunity 2014, 41, 843–852. [Google Scholar] [CrossRef] [PubMed]
  224. Wang, Z.; Chen, J.; Hu, J.; Zhang, H.; Xu, F.; He, W.; Wang, X.; Li, M.; Lu, W.; Zeng, G.; et al. cGAS/STING axis mediates a topoisomerase II inhibitor–induced tumor immunogenicity. J. Clin. Investig. 2019, 129, 4850–4862. [Google Scholar] [CrossRef]
Figure 1. Activation of the cGAS–STING pathway during innate immune response. The DNA from intracellular and extracellular sources is sensed by cGAS. After DNA recognition, cGAS catalyzes its substrates, GTP and ATP, into cGAMP. cGAMP is then sensed by STING located at the ER. Upon cGAMP binding, STING translocates from the ER to the ERGIC and recruits TBK1, followed by which STING and TBK1 are trafficked to the Golgi with the assistance of COPII and TRAPβ. Finally, STING and TBK1 interact with IRF3 and IκBα, culminating in type-I IFN and cytokine production. Abbreviations: mtDNA, mitochondrial DNA; cGAS, cyclic GMP-AMP synthase; 2′3′-cGAMP, 2′3′-cGMP-AMP; STING, stimulator of interferon genes; ER, endoplasmic reticulum; ERGIC, ER–Golgi intermediate compartment; TRAPβ, translocon-associated protein β; COPII, cytoplasmic coat protein complex-II; TBK1, TANK-binding kinase 1; IRF-3, interferon regulatory factor-3; IκBα: inhibitors of transcription factor NF-κB; type-I IFN, type-I interferon; IL-6, interleukin 6; TNF-α, tumor necrosis factor-α; GTP, guanosine triphosphate; ATP, adenosine triphosphate.
Figure 1. Activation of the cGAS–STING pathway during innate immune response. The DNA from intracellular and extracellular sources is sensed by cGAS. After DNA recognition, cGAS catalyzes its substrates, GTP and ATP, into cGAMP. cGAMP is then sensed by STING located at the ER. Upon cGAMP binding, STING translocates from the ER to the ERGIC and recruits TBK1, followed by which STING and TBK1 are trafficked to the Golgi with the assistance of COPII and TRAPβ. Finally, STING and TBK1 interact with IRF3 and IκBα, culminating in type-I IFN and cytokine production. Abbreviations: mtDNA, mitochondrial DNA; cGAS, cyclic GMP-AMP synthase; 2′3′-cGAMP, 2′3′-cGMP-AMP; STING, stimulator of interferon genes; ER, endoplasmic reticulum; ERGIC, ER–Golgi intermediate compartment; TRAPβ, translocon-associated protein β; COPII, cytoplasmic coat protein complex-II; TBK1, TANK-binding kinase 1; IRF-3, interferon regulatory factor-3; IκBα: inhibitors of transcription factor NF-κB; type-I IFN, type-I interferon; IL-6, interleukin 6; TNF-α, tumor necrosis factor-α; GTP, guanosine triphosphate; ATP, adenosine triphosphate.
Ijms 24 13316 g001
Figure 2. The cGAS–STING pathway in cell death, autophagy, and senescence. (A) The cGAS–STING pathway in MT stress-derived apoptosis. Permeabilization of the mitochondrial outer membrane induces the release of mtDNA and cytochrome c into the cytosol. The mtDNA then triggers the activation of the cGAS–STING pathway, leading to the production of type-I IFN and the association of cytochrome c with APAF1 to form the caspase-9-induced apoptosome. This structure further drives apoptosis and promotes caspase-3 and -7 production, which in turn causes the cleavage of cGAS and IRF-3. (B) The cGAS–STING pathway in virus-derived apoptosis. During infection with extracellular pathogens, the pathogen-related ER stress results in the direct recruitment of phosphorylated TBK1 by STING to trigger apoptosis via a BAX–BAK-dependent pathway. (C) The cGAS–STING pathway in pyroptosis. In specific cell types, STING is sorted into lysosomes after the activation of the cGAS–STING pathway, which leads to the permeabilization of the lysosome membrane and the release of cathepsins into the cytosol. Next, the cytosolic cathepsins trigger the activation of caspase-1 and NLRP3 inflammasome to drive K+ efflux-dependent pyroptosis. (D) The cGAS–STING pathway in necroptosis. The activation of the cGAS–STING pathway by mtDNA leads to the overexpression of type-I IFN and TNF-α. This overexpression further stimulates the activation of MLKL and RIPK to execute necroptosis. (E) The cGAS–STING pathway in autophagy. To prevent the overexpression of type-I IFN, STING traffics to the ERGIC after binding to cGAMP, independently of the recruitment of TBK1. This STING-containing ERGIC then acts as a source of LC3B, which is essential to autophagosome formation to induce autophagy. (F) The cGAS–STING pathway in senescence. When cells undergo ionizing radiation, the NL cannot maintain the NE structure, leading to the leakage of chromatin fragments (created by senescence-associated DNA damage) from the nucleus to the cytosol to initiate the overproduction of type-I IFN and SASP via the cGAS–STING pathway. Consequently, the overproduction of type-I IFN activates the p53-p21 signaling pathway to increase p16INK4 levels, a hallmark of cellular senescence. Abbreviations: MT stress, mitochondrial stress; mt DNA, mitochondrial DNA; APAF1, adapter apoptotic protease activating factor-1; BAX, BCL-2 associated X protein; BAK, BCL-2 homologous killer; K+ efflux, potassium efflux; type-I IFN, type-I interferon; TNF-α, tumor necrosis factor-α; MLKL, mixed lineage kinase domain-like protein; RIPK3, receptor-interacting serine/threonine protein kinase 3; LC3B, microtubule-associated proteins 1A/1B light chain 3B; SASP, senescence-associated secretory phenotype; cGAS–STING, cyclic GMP-AMP synthase-stimulator of interferon genes; IRF-3, interferon regulatory factor-3; ER, endoplasmic reticulum; TBK1, TANK-binding kinase 1; ERGIC, ER–Golgi intermediate compartment; NL, nuclear lamina; NE, nuclear envelope.
Figure 2. The cGAS–STING pathway in cell death, autophagy, and senescence. (A) The cGAS–STING pathway in MT stress-derived apoptosis. Permeabilization of the mitochondrial outer membrane induces the release of mtDNA and cytochrome c into the cytosol. The mtDNA then triggers the activation of the cGAS–STING pathway, leading to the production of type-I IFN and the association of cytochrome c with APAF1 to form the caspase-9-induced apoptosome. This structure further drives apoptosis and promotes caspase-3 and -7 production, which in turn causes the cleavage of cGAS and IRF-3. (B) The cGAS–STING pathway in virus-derived apoptosis. During infection with extracellular pathogens, the pathogen-related ER stress results in the direct recruitment of phosphorylated TBK1 by STING to trigger apoptosis via a BAX–BAK-dependent pathway. (C) The cGAS–STING pathway in pyroptosis. In specific cell types, STING is sorted into lysosomes after the activation of the cGAS–STING pathway, which leads to the permeabilization of the lysosome membrane and the release of cathepsins into the cytosol. Next, the cytosolic cathepsins trigger the activation of caspase-1 and NLRP3 inflammasome to drive K+ efflux-dependent pyroptosis. (D) The cGAS–STING pathway in necroptosis. The activation of the cGAS–STING pathway by mtDNA leads to the overexpression of type-I IFN and TNF-α. This overexpression further stimulates the activation of MLKL and RIPK to execute necroptosis. (E) The cGAS–STING pathway in autophagy. To prevent the overexpression of type-I IFN, STING traffics to the ERGIC after binding to cGAMP, independently of the recruitment of TBK1. This STING-containing ERGIC then acts as a source of LC3B, which is essential to autophagosome formation to induce autophagy. (F) The cGAS–STING pathway in senescence. When cells undergo ionizing radiation, the NL cannot maintain the NE structure, leading to the leakage of chromatin fragments (created by senescence-associated DNA damage) from the nucleus to the cytosol to initiate the overproduction of type-I IFN and SASP via the cGAS–STING pathway. Consequently, the overproduction of type-I IFN activates the p53-p21 signaling pathway to increase p16INK4 levels, a hallmark of cellular senescence. Abbreviations: MT stress, mitochondrial stress; mt DNA, mitochondrial DNA; APAF1, adapter apoptotic protease activating factor-1; BAX, BCL-2 associated X protein; BAK, BCL-2 homologous killer; K+ efflux, potassium efflux; type-I IFN, type-I interferon; TNF-α, tumor necrosis factor-α; MLKL, mixed lineage kinase domain-like protein; RIPK3, receptor-interacting serine/threonine protein kinase 3; LC3B, microtubule-associated proteins 1A/1B light chain 3B; SASP, senescence-associated secretory phenotype; cGAS–STING, cyclic GMP-AMP synthase-stimulator of interferon genes; IRF-3, interferon regulatory factor-3; ER, endoplasmic reticulum; TBK1, TANK-binding kinase 1; ERGIC, ER–Golgi intermediate compartment; NL, nuclear lamina; NE, nuclear envelope.
Ijms 24 13316 g002
Figure 3. The different roles of the cGAS–STING pathway in host defense and various diseases. (A) The role of the cGAS–STING pathway in host defense. The genome of invading pathogens, such as viruses, bacteria and parasites, in the cytosol can be sensed by cGAS, triggering the cGAS–STING pathway-induced production of type-I IFN and other cytokines to eliminate the genomes of invading pathogens. (B) The role of the cGAS–STING pathway in ALS. The missense mutation in the TDP-43 protein enables its absorption into the mitochondria and causes mtDNA leakage into the cytosol, prompting overproduction of type-I IFN and NF-κB to initiate hyperinflammatory responses. (C) The role of the cGAS–STING pathway in HD. The mutation of the N-terminal polyglutamine in the huntingtin protein (mHTT) leads to the abnormal accumulation of damaged DNA in the brain’s striatum, initiating the STING-dependent type-I IFN production and autophagy. (D) The role of the cGAS–STING pathway in MI. Synchronously dying cells are sensed by cardiac macrophages, and damaged DNA from these cells is internalized by cardiac macrophages. Furthermore, this internalization process leads to the inhibition of the transformation of cardiac macrophages from inflammatory cells to a reparative phenotype via the activation of the cGAS–STING pathway. (E) The role of the cGAS–STING pathway in AP. Damaged DNA released by the dying pancreatic acinar cells is internalized by leukocytes, resulting in severe inflammation due to the cGAS–STING pathway-induced overproduction of type-I IFN and other cytokines. (F) The role of the cGAS–STING pathway in silicosis. Silica microparticles invade the lungs, causing an increased release of chromatin fragments and mtDNA from dying cells into the BALF. These chromatin fragments and mtDNA then engage the cGAS–STING pathway to increase type-I IFN and other cytokine levels, causing chronic progressive fibrotic inflammation in the lungs. (G) The role of the cGAS–STING pathway in NASH and liver IRI. Lipotoxicity in NASH and ROS in IRI both lead to mitochondrial stress and the release of mtDNA into the cytosol of Kupffer cells. Furthermore, these changes result in adipose tissue inflammation via the cGAS–STING pathway and cGAS–STING–NLRP3 inflammasome-driven cell death in NASH and IRI, respectively. (H) The role of the cGAS–STING pathway in SLE. The loss-of-function mutation in TREX1 is suggested as the main cause of SLE. TREXI mutations lead to the abnormal accumulation of cytosolic DNA, which continuously activates the type-I IFN signaling cascade and amplifies inflammation via cGAS–STING. (I) The role of the cGAS–STING pathway in AGS. The loss-of-function mutation in TREX1 and the RNaseH2 complex are considered the major components of AGS. In this context, mutations in TREX1 and the RNaseH2 complex lead to the failure to eliminate genomic fragments in the cytosol and maintain genome stability, causing increased cGAS–STING pathway-dependent expression of type-I IFN. (J) The role of the cGAS–STING pathway in COPA, SAVI, and FCL. COPA is caused by the abnormal accumulation of STING in the GR. SAVI is attributed to the gain-of-function mutation in the STING-encoding gene, which leads to the spontaneous translocation of STING from the ER to the GR. This abnormal translocation activates intensive type-I IFN signals independently of cGAMP stimulation. Finally, FCL is caused by the heterozygous gain-of-function mutation in STING, resulting in the spontaneous dimerization of STING and constitutive activation of the type-I IFN signature. (K) The role of the cGAS–STING pathway in RA. The abnormal accumulation of cytosolic DNA leads to the expression of cytokines and chemokines via cGAS–STING activation, contributing to RA development. (L) The role of the cGAS–STING pathway in neurodegeneration. In microglia, mtDNA in the cytosol of microglia contributes to the chronic inflammatory response, which leads to neurodegeneration. Abbreviations: ALS, amyotrophic lateral sclerosis; MT stress, mitochondrial stress; mt DNA, mitochondrial DNA; TDP-43 protein, TAR DNA-binding protein 43; HD, Huntington disease; mHTT, mutation of the N-terminal polyglutamine in huntingtin protein; MI, myocardial infarction; AP, acute pancreatitis; BALF, bronchoalveolar lavage fluid; NASH, nonalcoholic steatohepatitis; IRI, ischemia-reperfusion injury; SLE, systemic lupus erythematosus; AGS, Aicardi–Goutières syndrome; COPA, COPA syndrome; SAVI, STING-associated vasculopathy with infantile-onset; FCL, familial chilblain lupus; RA, rheumatoid arthritis; cGAS–STING, cyclic GMP-AMP synthase-stimulator of interferon genes; type-I IFN, type-I interferon.
Figure 3. The different roles of the cGAS–STING pathway in host defense and various diseases. (A) The role of the cGAS–STING pathway in host defense. The genome of invading pathogens, such as viruses, bacteria and parasites, in the cytosol can be sensed by cGAS, triggering the cGAS–STING pathway-induced production of type-I IFN and other cytokines to eliminate the genomes of invading pathogens. (B) The role of the cGAS–STING pathway in ALS. The missense mutation in the TDP-43 protein enables its absorption into the mitochondria and causes mtDNA leakage into the cytosol, prompting overproduction of type-I IFN and NF-κB to initiate hyperinflammatory responses. (C) The role of the cGAS–STING pathway in HD. The mutation of the N-terminal polyglutamine in the huntingtin protein (mHTT) leads to the abnormal accumulation of damaged DNA in the brain’s striatum, initiating the STING-dependent type-I IFN production and autophagy. (D) The role of the cGAS–STING pathway in MI. Synchronously dying cells are sensed by cardiac macrophages, and damaged DNA from these cells is internalized by cardiac macrophages. Furthermore, this internalization process leads to the inhibition of the transformation of cardiac macrophages from inflammatory cells to a reparative phenotype via the activation of the cGAS–STING pathway. (E) The role of the cGAS–STING pathway in AP. Damaged DNA released by the dying pancreatic acinar cells is internalized by leukocytes, resulting in severe inflammation due to the cGAS–STING pathway-induced overproduction of type-I IFN and other cytokines. (F) The role of the cGAS–STING pathway in silicosis. Silica microparticles invade the lungs, causing an increased release of chromatin fragments and mtDNA from dying cells into the BALF. These chromatin fragments and mtDNA then engage the cGAS–STING pathway to increase type-I IFN and other cytokine levels, causing chronic progressive fibrotic inflammation in the lungs. (G) The role of the cGAS–STING pathway in NASH and liver IRI. Lipotoxicity in NASH and ROS in IRI both lead to mitochondrial stress and the release of mtDNA into the cytosol of Kupffer cells. Furthermore, these changes result in adipose tissue inflammation via the cGAS–STING pathway and cGAS–STING–NLRP3 inflammasome-driven cell death in NASH and IRI, respectively. (H) The role of the cGAS–STING pathway in SLE. The loss-of-function mutation in TREX1 is suggested as the main cause of SLE. TREXI mutations lead to the abnormal accumulation of cytosolic DNA, which continuously activates the type-I IFN signaling cascade and amplifies inflammation via cGAS–STING. (I) The role of the cGAS–STING pathway in AGS. The loss-of-function mutation in TREX1 and the RNaseH2 complex are considered the major components of AGS. In this context, mutations in TREX1 and the RNaseH2 complex lead to the failure to eliminate genomic fragments in the cytosol and maintain genome stability, causing increased cGAS–STING pathway-dependent expression of type-I IFN. (J) The role of the cGAS–STING pathway in COPA, SAVI, and FCL. COPA is caused by the abnormal accumulation of STING in the GR. SAVI is attributed to the gain-of-function mutation in the STING-encoding gene, which leads to the spontaneous translocation of STING from the ER to the GR. This abnormal translocation activates intensive type-I IFN signals independently of cGAMP stimulation. Finally, FCL is caused by the heterozygous gain-of-function mutation in STING, resulting in the spontaneous dimerization of STING and constitutive activation of the type-I IFN signature. (K) The role of the cGAS–STING pathway in RA. The abnormal accumulation of cytosolic DNA leads to the expression of cytokines and chemokines via cGAS–STING activation, contributing to RA development. (L) The role of the cGAS–STING pathway in neurodegeneration. In microglia, mtDNA in the cytosol of microglia contributes to the chronic inflammatory response, which leads to neurodegeneration. Abbreviations: ALS, amyotrophic lateral sclerosis; MT stress, mitochondrial stress; mt DNA, mitochondrial DNA; TDP-43 protein, TAR DNA-binding protein 43; HD, Huntington disease; mHTT, mutation of the N-terminal polyglutamine in huntingtin protein; MI, myocardial infarction; AP, acute pancreatitis; BALF, bronchoalveolar lavage fluid; NASH, nonalcoholic steatohepatitis; IRI, ischemia-reperfusion injury; SLE, systemic lupus erythematosus; AGS, Aicardi–Goutières syndrome; COPA, COPA syndrome; SAVI, STING-associated vasculopathy with infantile-onset; FCL, familial chilblain lupus; RA, rheumatoid arthritis; cGAS–STING, cyclic GMP-AMP synthase-stimulator of interferon genes; type-I IFN, type-I interferon.
Ijms 24 13316 g003
Figure 4. The cGAS–STING pathway in cancer. (A) In certain cancers, the cGAS–STING pathway induces autophagy-dependent cell death to limit the transformation of normal cells into cancerous cells. (B) The cGAS–STING pathway inhibits the growth of certain cancer cells by inducing the secretion of anti-proliferative molecules that promote senescence. (C) In specific cancers, 2′3′-cGAMP from the cancer cells enters the cytosol of DCs through the gap junction or TEX, activating STING to produce type-I IFN. The type-I IFN from the DCs then cross-primes CD8+ T cells to eliminate tumorigenic cells. (D) Within some cancers, 2′3′-cGAMP from the cancer cells is transferred into non-tumor bystander cells via the gap junction and activates STING, resulting in the recruitment of NK cells to restrain tumor growth. (E) In the case of certain cancer cells, chemical treatments cause the introduction of damaged DNA into the cytosol. This damaged DNA further engages the cGAS–STING pathway, inducing the secretion of DMBA to facilitate inflammation and maintain the tumor microenvironment. (F) STING activation also occurs after 2′3′-cGAMP from specific cancer cells are trafficked to astrocytes via the gap junction. The activated STING in the astrocytes promotes STAT1 and NF-κB production, further facilitating the development of metastatic cancer cells. (G) The loss of function of cGAS or STING leads to immunosuppression in cancer cells due to the hindered production of type-I IFN and other cytokines. Abbreviations: TEX, tumor-derived exosomes; DCs, infiltrating dendritic cells; NK cells, natural killer cells; DMBA, 7, 12-dimethylbenz(a)anthracene; cGAS–STING, cyclic GMP-AMP synthase-stimulator of interferon genes; 2′3′-cGAMP, 2′3′-cGMP-AMP; type-I IFN, type-I interferon.
Figure 4. The cGAS–STING pathway in cancer. (A) In certain cancers, the cGAS–STING pathway induces autophagy-dependent cell death to limit the transformation of normal cells into cancerous cells. (B) The cGAS–STING pathway inhibits the growth of certain cancer cells by inducing the secretion of anti-proliferative molecules that promote senescence. (C) In specific cancers, 2′3′-cGAMP from the cancer cells enters the cytosol of DCs through the gap junction or TEX, activating STING to produce type-I IFN. The type-I IFN from the DCs then cross-primes CD8+ T cells to eliminate tumorigenic cells. (D) Within some cancers, 2′3′-cGAMP from the cancer cells is transferred into non-tumor bystander cells via the gap junction and activates STING, resulting in the recruitment of NK cells to restrain tumor growth. (E) In the case of certain cancer cells, chemical treatments cause the introduction of damaged DNA into the cytosol. This damaged DNA further engages the cGAS–STING pathway, inducing the secretion of DMBA to facilitate inflammation and maintain the tumor microenvironment. (F) STING activation also occurs after 2′3′-cGAMP from specific cancer cells are trafficked to astrocytes via the gap junction. The activated STING in the astrocytes promotes STAT1 and NF-κB production, further facilitating the development of metastatic cancer cells. (G) The loss of function of cGAS or STING leads to immunosuppression in cancer cells due to the hindered production of type-I IFN and other cytokines. Abbreviations: TEX, tumor-derived exosomes; DCs, infiltrating dendritic cells; NK cells, natural killer cells; DMBA, 7, 12-dimethylbenz(a)anthracene; cGAS–STING, cyclic GMP-AMP synthase-stimulator of interferon genes; 2′3′-cGAMP, 2′3′-cGMP-AMP; type-I IFN, type-I interferon.
Ijms 24 13316 g004
Table 1. Intracellular modulators that suppress the activity of cGAS, cGAMP, and STING.
Table 1. Intracellular modulators that suppress the activity of cGAS, cGAMP, and STING.
TargetModulatorsModificationMechanismsRefs.
cGASAKTPhosphorylationPhosphorylates the active domain of cGAS[56]
BTK Reduces the binding efficiency of DNA[57]
Caspase-1Cleavage (degradation)Dampens the production of type-I
IFN
[51]
TRIM38SumoylationInhibits the activation of cGAS
without microbial invasion
[58]
TTLL6PolyglutamylationImpedes the cGAS–DNA synthesis
activity
[59]
TTLL4MonoglutamylationReduces the enzymatic activity of cGAS[59]
Cia-cGASDirect bindingPrevents the binding of cGAS to
DNA
[60]
OASLDirect bindingDecreases the synthase activity of
cGAS
[61]
cGAMPENP11PhosphodiesteraseHydrolyzes cGAMP[62]
STINGUSP13DeubiquitylationPrevents the recruitment of TBK1
by STING
[63]
NLRC3Direct bindingImpairs the interaction between
STING and TBK1
[64]
NLRX1Direct bindingImpairs the interaction between
STING and TBK1
[65]
ULK1PhosphorylationThwarts sustained type I IFN production by STING[66]
RNF5UbiquitylationDegrades STING at mitochondria [67]
AP-1Directly bindingTerminates the sorting of STING in Golgi[68]
2-BPDirectly bindingSuppresses palmitoylation of STING in Golgi[69]
Abbreviations: AKT, RACα serine/threonine-protein kinase (also known as PKB); BTK, Bruton’s tyrosine kinase; ENPP1, ectonucleotide pyrophosphatase/phosphodiesterase family member 1; TRIM38, tripartite motif 38; TTLL6, tubulin-tyrosine ligase-like member 6; TTLL4, tubulin-tyrosine ligase-like member 4; USP13, ubiquitin specific peptidase 13; cia-cGAS, circular RNA with higher cGAS affinity than dsDNA; OASL, 20–50 oligoadenylate synthase; NLRC3, NOD-like receptor family 3 containing CARD domain; NLRX1, NLR family member X1; ULK1, serine/threonine UNC-51-like kinase (ULK1/ATG1); RNF5, RING finger domain; AP-1, adaptor protein complex 1; 2-BP, 2-bromopalmitate; cGAS, cyclic GMP-AMP synthase; cGAMP, cGMP-AMP; STING, stimulator of interferon genes.
Table 2. Intracellular modulators that enhance the activity of cGAS, cGAMP, and STING.
Table 2. Intracellular modulators that enhance the activity of cGAS, cGAMP, and STING.
TargetModulatorsModificationMechanismsRefs
cGASSENP7DesumoylationCleaves sumoylated cGAS to increase its activity[70]
HDAC3DeacetylationEnhances the activity of cGAS[71]
RNF185E3 ubiquitylationPromotes the catalytic activity of cGAS[72]
Mn2+ ionDirect bindingAugments cGAMP–STING binding affinity[73]
G3BP1Direct bindingRequired for the high-level activation of cGAS[74]
R848Indirect assistanceEnables the recognition of HIV-1 infection by cGAS[75,76]
PAM3Indirect assistanceEnables the recognition of HIV-1 infection by cGAS[75,76]
TRIM41Monoubiquitylation
E3 ubiquitylation
Promotes the activation of cGAS
against viruses infection
Protects cGAS degradation from
autophagy
[77,78]
cGAMPZn2+ ionDirect bindingAccelerates cGAS–DNA synthesis for liquid phase condensation[29]
STINGAMFRE3 ubiquitylationFacilitates recruitment of TBK1 by STING and its translocation[79]
MUL1PolyubiquitylationPromotes STING trafficking[80]
TMEM203Direct bindingCooperates with STING to activate TBK1 and IRF3[81]
Abbreviations: SENP7, sentrin/SUMO-specific protease 7; HDAC3, histone deacetylase 3; RNF185, RING finger protein 185; AMFR, autocrine motility factor receptor; MUL1, mitochondrial E3 ubiquitin protein ligase 1; Zn2+ ion, zinc ion; Mn2+ ion, manganese ion; G3BP1, GTPase-activating protein SH3 domain-binding protein 1; TMEM203, transmembrane protein 203; R848, TLR7/8 agonist; PAM3, TLR2 agonist,; TRIM41, tripartite motif protein 41; cGAS, cyclic GMP-AMP synthase; STING, stimulator of interferon genes.
Table 5. Inhibitors of cGAS or STING.
Table 5. Inhibitors of cGAS or STING.
CompoundCharacteristicsBiological EffectsRefs
cGAS inhibitors
Catalytic site inhibitors
PF-06928215Binds to the active site of cGAS
occupied by ATP
Attenuates type-I IFN signaling in
AGS mouse models
[201,202]
RU.521Competitively occupies the catalytic
site of cGAS
Reduces expression levels of Ifnb1
in BMDMs
[203]
G150Binds to the active site of cGASThe IC50 is 0.62
μM in primary H-macrophages
[204]
CU-76
CU-32
Inhibit the dimerization of cGASSpecifically target the inhibition of the
cGAS–STING pathway
[205]
AspirinAcetylates Lys384, Lys394, and Lys414
amino acid residues of cGAS
Suppresses immune responses in AGS mouse models[71]
Compound CReduces cGAMP accumulationSuppresses type-I IFN induction[206]
DNA-binding inhibitors
AMDsBlock cGAS–dsDNA interactionThe IC50 in THP-1 cells is 3–25 μM[207,208]
SuraminDisrupts the formation of cGAS–dsDNA complexModulates type-I IFN level in THP-1 cells[209]
A151Competitively binds to DNA-binding domain of cGASInhibits type-I IFN signaling in TREX1-deficient cells[210]
STING inhibitors
CDN-binding site inhibitors
Astin CInhibits the recruitment of IRF3 to
the STING-TBK1 signalosome
Inhibits the expression of Ifnb, Cxcl0, and Tnf mRNA in multiple tissues of mouse models[211]
THIQTransforms STING into an inactive, open conformationInhibits cGAMP-induced type-I IFN secretion in THP-1 cells[212]
Palmitoylation inhibitors
NitrofuransBind to Cys91 to inhibit the
palmitoylation of STING
Strongly suppress inflammatory response in mouse models[213]
Indole ureaseForms a covalent bond with Cys91 of STINGReduces the production of type-I IFN in TREX1-deficient mouse tumor model[213]
NO2-FAsCovalently modify STING in Cys88 and Cys91Reduce type-I IFN production in response to HSV-2 infection in THP-1 cells[214]
Abbreviations: BMDMs, bone marrow-derived macrophages; IC50, half-maximal inhibitory concentration; primary H-macrophages, primary human macrophages; AMDs, antimalarial drugs (such as hydroxychloroquine and quinacrine); THP-1 cells, human myeloid leukemia mononuclear cells; THIQ, tetrahydroisoquinoline; NO2-FAs, nitro fatty acids; cGAS, cyclic GMP-AMP synthase; STING, stimulator of interferon genes; type-I IFN, type-I interferon; AGS, Aicardi–Goutières syndrome; cGAMP, cGMP-AMP; dsDNA, double-stranded DNA; CDNs, cyclic dinucleotides; IRF3, interferon regulatory factor-3; TBK1, TANK-binding kinase 1.
Table 6. Agonists of STING.
Table 6. Agonists of STING.
CompoundMechanismsPreclinical EffectsRefs
CDN agonists
Natural CDNsActivate APCs and CD8+ T cellsEnhance antitumor signals[215]
ADU-S100Activates all STING variants and improves their stabilityInduces durable tumor regression[215]
cGAMP-NPsInserted liposomal NPs can
deliver cGAMP more efficiently
Create anti-tumor
microenvironment
[216]
Non-CDN agonists
DMXAAHigher affinity than cGAMP and activates STING efficientlyRestricts tumorigenesis[217]
ZebularineEnhances the STING gene expression by demethylationReduces tumor burden[218]
G10Stabilizes the structure of STINGSuppresses tumor growth[219]
ABZIInduces type-I IFN production that is 400 times higher than that of cGAMPTumor volume regression[220]
DSDPInduces STING-dependent
cytokine responses
Triggers antiviral responses[221]
BNBCSpecifically activates STINGTriggers antiviral responses[222]
Indirect agonists
RadiotherapyCauses DSBs and the accumulation of cytosolic DNAActivates adaptive immune
response
[223]
CisplatinInhibits DDR and the release of chromatin fragments into the cytosolActivates CD8+ T cells[43]
TeniposideInduces DNA damage in cancer cellsActivates DCs and T cells[224]
Abbreviations: CDNs, cyclic dinucleotides; APCs, antigen-presenting cells; cGAMP-NPs, cGAMP-nanoparticles; DMXAA, 5, 6-dimethylxanthenone-4-acetic acid; ABZI, amide benzimidazole; DSDP, dispiro diketopiperzine; BNBC, 6-bromo-N-(naphthalen-1-yl)benzo[d][1,3] dioxole-5-carboxamide; STING, stimulator of interferon genes; cGAMP, cGMP-AMP; type-I IFN, type-I interferon; DSBs, double-strand breaks; DCs, dendritic cells.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhou, J.; Zhuang, Z.; Li, J.; Feng, Z. Significance of the cGAS-STING Pathway in Health and Disease. Int. J. Mol. Sci. 2023, 24, 13316. https://doi.org/10.3390/ijms241713316

AMA Style

Zhou J, Zhuang Z, Li J, Feng Z. Significance of the cGAS-STING Pathway in Health and Disease. International Journal of Molecular Sciences. 2023; 24(17):13316. https://doi.org/10.3390/ijms241713316

Chicago/Turabian Style

Zhou, Jinglin, Zhan Zhuang, Jiamian Li, and Zhihua Feng. 2023. "Significance of the cGAS-STING Pathway in Health and Disease" International Journal of Molecular Sciences 24, no. 17: 13316. https://doi.org/10.3390/ijms241713316

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