• OPEN ACCESS

Update on the STING Signaling Pathway in Developing Nonalcoholic Fatty Liver Disease

  • Wei Liu1,#,
  • Zhili Zhang Chen1,#,
  • Chenhui Yang1,
  • Yaofu Fan1,2,
  • Liang Qiao1,
  • Shaofeng Xie1,2,*  and
  • Lin Cao1,2,* 
 Author information
Journal of Clinical and Translational Hepatology 2024;12(1):91-99

DOI: 10.14218/JCTH.2023.00197

Abstract

Nonalcoholic fatty liver disease (NAFLD) is a prevalent chronic liver condition with limited treatment options. Inflammation caused by metabolic disturbances plays a significant role in NAFLD development. Stimulator of interferon gene (STING), a critical regulator of innate immunity, induces the production of interferons and other pro-inflammatory factors by recognizing cytoplasmic DNA to defend against pathogen infection. The STING-mediated signaling pathway appears to play a vital role in hepatic inflammation, metabolic disorders, and even carcinogenesis. Promisingly, pharmacological interventions targeting STING have shown improvements in the pathological state of NAFLD. Macrophages, dendritic cells, natural killer cells, and T cell pathways regulated by STING present potential novel druggable targets for NAFLD treatment. Further research and development in this area may offer new therapeutic options for managing NAFLD effectively.

Keywords

Stimulator of interferon gene, Macrophage, Innate immunity, Nonalcoholic fatty liver disease

Introduction

Owing to the rising prevalence of obesity and type 2 diabetes, nonalcoholic fatty liver disease (NAFLD) has become the most common chronic liver disease, with a global prevalence of around 25%.1 NAFLD encompasses a spectrum of conditions, ranging from simple steatosis to nonalcoholic steatohepatitis (NASH) and even leading to cirrhosis and hepatocellular carcinoma (HCC). The pathogenesis of NAFLD is complex, but inflammation has a crucial role in disease progression.2 Multiple immune cells and inflammation-related signaling pathways have been implicated in the development of NAFLD.3 Among these pathways, the stimulator of interferon gene (STING) signaling pathway, an essential regulatory protein in the innate immune system, is capable of sensing double-stranded DNA released by hepatocyte injury or the enterohepatic circulation. This activation of STING leads to liver inflammation, metabolic disturbances, and even carcinogenesis in NAFLD. Moreover, studies have shown that interventions targeting STING can effectively alleviate the pathological state of NAFLD, offering promise for potential novel targeted therapies to improve the prognosis of patients with NAFLD.4,5

Innate immunity in NAFLD

The innate immune system has a crucial role in NAFLD pathogenesis, with crosstalk between innate immune cells and liver cells contributing to the disease’s initiation and progression. Various genetic and environmental factors lead to excessive lipid accumulation in the liver, triggering pathological reactions like oxidative stress in mitochondria, endoplasmic reticulum (ER) stress and hepatocyte autophagy.6 Injured or apoptotic hepatocytes release damage-related molecules, such as nucleic acids or proteins, activating liver pattern recognition receptors and alerting the hepatic-innate immune microenvironment in NAFLD. This response involves resident Kupffer cell activation and recruitment of immune cells like neutrophils, monocytes, natural killer (NK) cells and NKT cells, releasing cytokines and chemokines that contribute to inflammation.7,8 Gut microbiota also plays a significant role in liver inflammation in NAFLD.9 The gut barrier is a physical barrier that directly prevents the translocation of luminal bacteria and bacterial-derived products or toxins. Because of changes of the intestinal flora in NAFLD, patients can develop compromised intestinal barrier integrity. Consequently, intestinal endothelial permeability, intestinal endotoxins and bacterial breakdown products known as pathogen-related molecular patterns, including nucleic acids, proteins, and flagellin, can cross the gut barrier and be transported through the portal vein to the liver. Then pattern recognition receptors recognize pathogen-related molecular patterns and activate innate immune cells, triggering intracellular signaling cascades and exacerbating inflammatory damage.10 Notably, DNA released by hepatocytes is considered a crucial factor in inducing sterile inflammation in NAFLD, while microbial DNA derived from the gut also contributes to liver inflammation in this condition.11,12 The newly discovered STING signaling pathway can identify almost all types of double-stranded DNA and is considered the main pathway for cytoplasmic DNA to induce the NAFLD inflammatory response.13,14

STING signaling pathway

STING, also known as the 173 transmembrane proteins, is a critical component of the host’s innate immune defense and can contribute to chronic autoimmune, autoinflammatory, and metabolic diseases.15 STING is primarily localized in ER and is expressed in various immune cells, including macrophages, dendritic cells, NK cells, and T cells. In addition to immune cells, STING is also expressed in endothelial and epithelial cells. These cells are exposed to the external environment and can be susceptible to infectious agents, making STING’s presence in these cells significant for detecting and responding to potential threats. Moreover, STING has been implicated in various chronic conditions, including autoimmune, autoinflammatory, and metabolic diseases, highlighting its role in regulating inflammation and immune responses throughout the body.16,17

Canonical model of STING activation

STING can be activated by cyclic dinucleotides, such as cyclic dimeric guanosine monophosphate (c-di-GMP), cyclic dimeric adenosine monophosphate (c-di-AMP) and cyclic-GMP-AMP (cGAMP), secreted by intracellular bacteria, or by noncanonical 2′3′-cGAMP generated by cGAMP synthase (cGAS).18 Involving cGAS and cGAMP in STING-mediation is considered a canonical activation pathway. When abnormal double-stranded DNA (such as pathogen DNA, host cell nucleus, or mitochondrial DNA) is phagocytosed by macrophages, cGAS, the cytoplasmic DNA sensor of pattern recognition receptors, catalyzes ATP and GTP to synthesize the endogenous second messenger cGAMP, which then activates STING in ER.19 STING then recruits and activates TANK binding kinase 1 (TBK1), which phosphorylates STING and the transcription factor interferon regulatory factor 3 (IRF3) to induce interferon-1 (IFN-1) and other cytokines to exert antiviral and antitumor effects.20 Furthermore, STING can activate the nuclear factor-kappa B (NF-κB) pathway through a redundant TBK1/inhibitor of NF-κB kinase epsilon (IKKε) mechanism to induce the expression of inflammatory cytokines, including tumor necrosis factor-alpha (TNF-α) and interleukin-6 (IL-6).21,22 STING and TBK1 activate the signal transducer and activator of transcription 6, which is consequently phosphorylated. Subsequently, phosphorylated signal transducer and activator of transcription 6 dimerizes and is delivered to the nucleus to induce the expression of chemokine ligand (CCL) 2, CCL20, and CCL26, mediating inflammatory responses.23 Moreover, the cGAS-STING signaling pathway can induce apoptosis, autophagy, and necrosis.24–27

Noncanonical model of STING activation

The STING signaling pathway can be activated independently by cGAS and cGAMP. Additionally, etoposide-induced DNA damage acts as a danger signal in the innate immune system, activating STING through the collaboration of DNA-mediated binding protein γ-interferon inducible protein 16 (IFI16), DNA damage response factors ataxia telangiectasia mutated (ATM), and poly (ADP-ribose) polymerase 1 (PARP-1). This leads to the assembly of an alternative STING signaling complex involving p53 and tumor necrosis factor receptor-associated factor 6 (TRAF6). TRAF6 catalyzes the assembly of K63-linked ubiquitin chains on STING, resulting in the activation of NF-κB, but not IRF3, leading to the expression of an alternative atypical STING-dependent gene program.28Table 1 provides a comparison between the canonical and noncanonical signaling pathways of STING, illustrating the similarities and differences between these two pathways.

Table 1

Similarities and differences of the STING canonical and noncanonical signaling pathways

STING signaling pathwayCanonical model activationNoncanonical model activation
Depend on STING; Both activate IRF3 and NF-κB
Activator typeDNA transfectionEtoposide-induced nuclear DNA damage
STING complex compositioncGAS, cGAMPATM, PARP-1, TRAF6, P53, IFI16
Transcription factor activation patternPredominantly activate IRF3 and moderately activate NF-κBPredominantly activate NF-κB and moderately activate IRF3
Gene expression profileHigher levels of the chemokine CXCL10 mRNA and the IRF3-responsive gene ISG56Higher levels of IL-6 and CCL20 mRNA

In summary, the STING signaling pathway plays a pivotal role in the immune system, contributing to antiviral and antitumor immune responses. It is also involved in various inflammatory diseases, including inflammatory bowel disease, NAFLD, unstable angina, and acute myocardial infarction.29

STING signaling pathway in NAFLD

Some studies indicate that the expression of STING increases with the progression of steatosis to NASH inflammation and fibrosis, particularly in the hepatic portal vein of patients with fibrosis.4 In addition, we performed single-cell nuclear sequencing analysis on liver tissues from three healthy individuals and nine NASH patients using the Gene Expression Omnibus dataset GSE212837. The results indicate that the expression of STING in the immune cells and hepatic stellate cells (HSCs) of NASH patients is significantly higher than that of the healthy control group (Fig. 1). This suggests that increased STING expression acts as an indicator of NAFLD progression and the severity of liver disease. However, another controversial study reported decreased expression of STING in the liver of NASH patients compared with those with steatosis alone. One possible explanation for this discrepancy could be the absence of fibrosis in the cohort participating in that specific study. Fibrosis is a critical characteristic of advanced NAFLD, and its presence may influence STING expression differently compared to earlier stages of the disease.30 Studies have also found that the levels of cGAS and STING proteins increased in liver carcinoma patients with the increase in liver fibrosis, and were higher than in the surrounding cancer tissues.31 These contrasting findings highlight the complexity of the STING signaling pathway’s involvement in NAFLD and the need for further research to fully understand its role in different stages of the disease.

Expression of STING in liver cells.
Fig. 1  Expression of STING in liver cells.

CTRL, control group; HSC, hepatic stellate cell; NASH, nonalcoholic steatohepatitis; STING, stimulator of interferon gene.

Role of STING in NAFL and NASH

The cGAS-STING pathway can activate liver macrophages to induce inflammation and lipid metabolic disorders, IR, and hepatocyte apoptosis. Accordingly, the cGAS-STING pathway likely plays a pivotal role in the progression of NAFLD and NASH (Fig. 2).

Activation of the STING signaling pathway in hepatic macrophages.
Fig. 2  Activation of the STING signaling pathway in hepatic macrophages.

AKT, protein kinase B; BAX, B-cell lymphoma-2-related protein X; cGAMP, cyclic-guanosine monophosphate-adenosine monophosphate; cGAS, cGAMP synthase; dsDNA, double-strand DNA; ER, endoplasmic reticulum; GLUT2, glucose transporter 2; IFNs, interferons; IκBα, inhibitor nuclear factor kappa B alpha; IL-6, interleukin-6; IKK, IκB-kinase; IRF3, interferon regulatory factor 3; IRS-1, insulin receptor substrate 1; JNK, c-jun N-terminal kinase; LDs, lipid droplets; MTORC1, rapamycin target protein complex 1; NF-κB, nuclear factor-kappa B; PI3K, phosphoinositide 3 kinase; RRAGA, ras-related GTP binding A; RRAGC, ras-related GTP binding C; SQSTM1, sequestosome1; STING, stimulator of interferon gene; TBK1, TANK binding kinase 1; TNF-α, tumor necrosis factor-alpha.

Inflammation

The cGAS-STING signaling pathway in the liver can be activated by various factors, including mitochondrial DNA (mtDNA), gut microbial DNA and iron deposition (Table 2). In NASH mice, the content of mtDNA in cytoplasmic mitochondria is significantly higher compared to normal mice, leading to the activation of the STING pathway in hepatic macrophages. This activation induces the expression of NF-κB and pro-inflammatory cytokines like TNF-α and IL-6, contributing to liver inflammation.32 The mechanism of accumulation of mtDNA in steatotic hepatocytes is likely to interfere with DNA replication by slowing down replication forks and activating the ATM and Rad3-related protein kinase (ATR)/cell cycle checkpoint kinase 1 (CHK1) pathway.33 Gut microbial DNA-containing extracellular vesicles can be removed by CRIg+ macrophages through complement component C3-mediated opsonization. However, obesity leads to decreased CRIg+ macrophages, and micro extracellular vesicles leakage diffuses into the liver, subsequently activating the cGAS-STING pathway and triggering liver inflammation.12 Elevated iron concentration in the livers of individuals with NAFLD can also activate the cGAS-STING pathway and cause liver inflammation.34 STING was unevenly distributed in the liver, mainly expressed in macrophages (CCR2+, S100A9+, Kupffer, and CD163+ cells). In contrast, hepatic sinusoidal endothelial cells (CD36+ cells, HSCs, and SMA+ cells), and other immune cells were poorly expressed.4,35 The findings indicate that STING is not expressed in human and mouse hepatic parenchymal cells.36,37 Subsequent experiments found that transplantation of bone marrow cells from wild-type mice into STINGgt mice could exacerbate the severity of liver inflammation and steatosis in STINGgt mice, suggesting that macrophage-derived STING is a crucial factor in promoting NAFLD.38 Under activation of the STING pathway, macrophages can be differentiated into pro-inflammatory M1 macrophages, which can act on liver cells to induce inflammation and fat deposition by secreting the inflammatory factors TNF-α, IL-1β and IL-6, consequently promoting NAFLD.39 However, a few studies suggest that STING exists in mouse hepatic parenchymal cells and activation of the STING/IRF3 pathway induces pro-apoptotic effects in hepatocytes independent of inflammation.40–42 Hepatic sinusoidal endothelial cells are the most abundant nonparenchymal cells in the liver. When the STING-IRF3 signaling pathway expressed in endothelial cells is activated, the activated IRF3 directly binds to the intercellular adhesion molecule-1 promoter, inducing intercellular adhesion molecule-1 expression in mouse vascular endothelial cells and monocyte-endothelial cell adhesion, promoting the generation of endothelial inflammation.17 Moreover, liver sinusoidal endothelial cell injury may act as a doorkeeper, implying the progression from NAFLD to the early stage of NASH.43 The cGAS-STING signaling pathway can also induce TBK1-mediated phosphorylation of p62/sequestosome1 and the formation of hepatic protein inclusions, which are vital indicators for differentiating NAFLD from NASH.44

Table 2

Pathological stimuli in NAFLD that activate the STING pathway

Pathological stimuliProduction of pathological stimuliMechanisms of promote NAFLD
Mitochondrial DNALipid overload induces nucleotide pool imbalance highlighted by a disruption of replication forks speed and activation of ATR/CHK1 pathwayInduce STING-NF-κB pathway activation and TNF-α and IL-6 expression under lipid overload in the Kupffer cells of liver
Microbial DNADeficiency of CRIg+ macrophages and leakage of intestinal EVs containing microbial DNAElevate the levels of cGAS expression and STING phosphorylation in hepatocytes and insulin target cells
Iron depositionHigh lipid induction enhances the iron accumulation by the upregulation of TFR1 and the down-regulation of FTH1Upgrade IFN-β and IL-6 expression via the cGAS-STING pathway and induce M1 polarization of macrophage

Disordered lipid metabolism

The STING signaling pathway has intricate connections with lipid metabolism, and these interactions play a crucial role in the pathogenesis of metabolic disorders like NAFLD. Sterol regulatory element binding protein cleavage-activated protein (SCAP) serves as a cholesterol sensor. A high-fat and high-cholesterol diet can lead to an abnormal increase in macrophage SCAP levels. This, in turn, recruits STING and TBK1 to the Golgi apparatus, activating NF-κB in macrophages. The activation of NF-κB promotes the release of inflammatory factors, enhancing lipid uptake and synthesis in the liver. This highlights the vital role of SCAP as a bridge molecule connecting lipid metabolism and inflammation, contributing to the development of NAFLD.45 Moreover, the release of mitochondrial DNA in hepatocytes can activate the STING-TBK1 pathway. TBK1-induced sequestosome1 phosphorylation, through STING activation, promotes activation of the rapamycin target protein complex 1 and inhibits lysosomal degradation of lipid droplets. This process leads to excessive lipid deposition in the liver, further exacerbating NAFLD.46 Furthermore, high-fat diet-induced activation of the STING-IRF3 pathway increases levels of lipid synthase sterol regulatory element binding protein-1c and lowers levels of the lipolytic enzyme peroxisome proliferator-activated receptor alpha.41 This imbalance promotes lipid synthesis and reduces lipid breakdown, contributing to lipid accumulation in the liver.47

Insulin resistance

Chronic low-grade inflammation of metabolic tissues is a central factor in developing IR. The STING signaling pathway may contribute to liver IR through multiple mechanisms, such as blocking insulin signaling pathways and promoting islet β cell senescence and apoptosis. Activating the STING-IRF3 pathway can induce hepatocyte inflammation and promote liver IR and gluconeogenesis by decreasing the expression of insulin signaling (phospho-protein kinase B (AKT)/total-AKT), phosphorylated glycogen synthase kinase 3β, and glycolytic enzymes (glucokinase, phosphofructokinase, and pyruvate kinase), and increasing the expression of glycogen synthase kinase 3β and hepatic gluconeogenic enzymes (glucose-6-phosphatase, phosphoenolpyruvate carboxykinase, and pyruvate carboxylase).41 Pro-inflammatory factors, including TNF-α and IL-6, can also suppress insulin-induced phosphorylation of AKT on Ser473, inducing IR.48 Furthermore, the STING pathway affects islet beta cells. Palmitic acid or hydrogen peroxide can activate the cGAS-STING-TBK1 pathway and upregulate pancreatic beta cell senescence indicators, including P21, P16ink4a, P53, and p53-binding protein 1. Intervention with the STING inhibitor C176 alleviates pancreatic beta cell senescence, glucose intolerance, and IR.49 Treating pancreatic beta cells with palmitic acid can activate the STING-IRF3 pathway, which promotes inflammation in pancreatic beta cells and induces apoptosis of pancreatic beta cells by upregulating the expression of apoptosis proteins: B-cell lymphoma-2 (Bcl2)-related protein X (BAX), caspase-3 (Casp-3), and PARP-1.50

Hepatocyte apoptosis

The STING-IRF3 signaling pathway in hepatocytes plays a crucial role in inducing apoptosis, and its activation can be triggered by various factors, including ER stress and free fatty acid (FFA) accumulation. In response to ER stress induced by CCl4 or FFA, the STING-IRF3 signaling pathway can lead to the upregulation of pro-apoptotic molecules, such as BAX, as well as apoptotic promoters like Casp-8 and Casp-3. This can subsequently result in hepatocyte apoptosis that contributes to liver injury and disease progression.40 Studies have shown that inhibiting STING and IRF3 using corresponding siRNAs significantly reduced the FFA-induced inflammatory cytokines and apoptotic signals, including BAX/Bcl2, clv-Casp-3/Casp-3 and clv-PARP/PARP. This suggests that inhibiting STING and blocking the interaction between phospho-IRF3 and BAX and Casp-3 can ameliorate hepatocyte apoptosis and may offer potential therapeutic strategies to mitigate liver injury and hepatocyte death.41

Role of STING in NAFLD-liver fibrosis

Liver fibrosis, a consequence of inflammation or death of hepatocytes, is a significant pathological process in various liver diseases. The crosstalk between different cellular pathways, including the STING pathway, contributes to the development of liver fibrosis. TAR DNA-binding protein 43 and X-box binding protein 1 have been identified as factors that promote macrophage self-mitochondrial DNA cytosolic leakage. This leads to macrophage STING activation, triggering intrahepatic inflammation and liver fibrosis.51,52 Additionally, ER stress-induced liver cell death has been linked to liver fibrosis through the STING-IRF3 pathway. This suggests that liver cell death independently contributes to the development of chemically induced liver fibrosis.40 HSCs are the primary cell type responsible for liver fibrosis. Studies have shown that treatment of HSCs with macrophages activated through the STING pathway results in increased activation of HSCs, leading to liver fibrosis. This suggests that STING activation in macrophages can promote liver fibrosis through a paracrine mechanism that affects HSCs and other cells involved in fibrosis.38

Role of STING in NAFLD-HCC

Indeed, NAFLD-HCC is a serious and feared liver-related complication, imposing a significant health burden. The cGAS-STING signaling pathway has been found to be involved in the occurrence and development of HCC. In HCC cells, sustained high levels of DNA damage can activate the cGAS-STING pathway, leading to the release of IFN-I. The presence of IFN-I can stimulate dendritic cells to migrate to tumor-draining lymph nodes and subsequently cross-activate tumor-specific CD8+ T cells. This process helps control both local and distant tumor growth by enhancing the body’s immune response against cancerous cells. Furthermore, the cGAS-STING pathway also has a role in recruiting NK cells and cytotoxic T lymphocytes, increasing the sensitivity of the immune system to attack cancer cells. This process contributes to enhancing the immune-mediated response against HCC cells, potentially providing novel avenues for therapeutic intervention.53 One limitation is that the upregulation of programmed cell death ligand-1 (PDL-1) and PDL-2 induced by IFNβ and IFN-γ, produced as a result of the cGAS-STING pathway activation, leads to cancer cell immune escape.54,55 Additionally, the activation of the STING pathway can induce indoleamine 2,3 dioxygenase activation, which promotes tumor immune escape and can suppress T cell activity. This further limits the immune system’s ability to target and eliminate cancer cells effectively.56,57 Moreover, STING activation leads to T lymphocyte apoptosis, allowing cancer cells to escape immune surveillance and evade detection by the immune system.58 These factors may contribute to the incomplete resolution of tumors after treatment with STING agonists. While the cGAS-STING pathway can trigger short-term inflammation to repress active oncogenes and support an antitumor immune response, prolonged or sustained inflammation leads to tissue destruction and contributes to cancer development or progression.59 Understanding the delicate balance between the beneficial and potentially detrimental effects of the cGAS-STING pathway is essential for developing targeted therapies that maximize its antitumor potential while minimizing unwanted side effects.

STING agonists and inhibitors for treating NAFLD

The STING signaling pathway has emerged as a significant target for the treatment of inflammatory and autoimmune diseases, and the development of STING agonists and inhibitors has become an active area of research. STING inhibitors can be categorized into two main types: covalent and noncovalent inhibitors. Covalent inhibitors work by forming a permanent bond with the target protein, while noncovalent inhibitors do not form such a bond and can be reversible in their action. Palmitoylation of the STING protein is necessary for its activation, and C176, C178, H151, and NO2-cLA covalently bind to STING to inhibit palmitoylation thereby leading to pathway inactivation.60,61 Noncovalent inhibitors control the activation of STING by competitively reducing the binding of 2′,3′-cGAMP to STING, mainly including astin C, SN-011, compound 18, gelsevirine, and palbociclib.62–66 Although many STING inhibitors have been discovered, none have been used for the treatment of NAFLD. However, studies have found that some drugs are effective for alleviating the pathological state of NAFLD by inhibiting the STING pathway (Table 3).66–71 Remdesivir was shown to significantly reduce liver inflammation and lipid metabolism disorders in NAFLD mice by inhibiting the STING-IRF3 signaling pathway.67 Lingguizhugan decoction and its critical components cinnamaldehyde, atractylenolide II and glycyrrhizinate can significantly reduce liver inflammation levels by inhibiting the STING-TBK1-NF-κB signaling pathway in hepatic Kupffer cells, subsequently decreasing lipid deposition in hepatocytes and effectively relieving NAFLD progression.68 Naringenin was shown to reduce liver inflammation and HSC activation by inhibiting the cGAS-STING signaling pathway, improving liver fibrosis.31 Sorafenib may attenuate the signal transduction of the STING pathway by inhibiting the dimerization of STING and the recruitment of TBK1 and IRF3, thereby alleviating liver inflammation and fat accumulation induced by palmitic acid.69,70Bifidobacterium triple live bacteria powder relieves HFD-induced NAFLD by inhibiting the expression of STING, thereby inhibiting the release of TNF-α, IL-1β, IL-6, IFN-β and p-NF-κB p65 induced by macrophages.71

Table 3

STING inhibitors for treating NASH/NAFLD liver fibrosis

DrugDosage and method of administrationMechanismReference
Remdesivir20 mg/kg/d (i.g.)Reduce liver inflammation and lipid metabolism disorders by inhibiting the STING-IRF3 signaling pathway66
Lingguizhugan decoction22 g/kg/d (i.g.)Inhibit the STING-TBK1-NF-κB signaling pathway to reduce liver inflammation67
Naringenin100 mg/kg/3 times/week (i.g.)Inhibit the cGAS-STING pathway thereby reducing the secretion of inflammatory factors by HSCs68
Sorafenib10 µmol/L (THP-1 cell culture)Inhibit the dimerization of STING and the recruitment of TBK1 and IRF369,70
Bifidobacterium triple live bacteria powder0.4 g/mouse/d (i.g.)Inhibit the STING-NF-κB pathway, thereby inhibiting the release of TNF-α, IL-1β, IL-6 and IFN-β71

STING agonists are currently under active development. Cyclic dinucleotides directly stimulate STING, including c-di-GMP, c-di-AMP, 3′,3′-cGAMP and 2′,3′-cGAMP.72 However, because of the instability, negative charge, and hydrophilicity of cyclic dinucleotides that limit their use, nonnucleotide agonists have been developed, such as 5,6-dimethylxanthone-4-acetic acid (DMXAA), flavone acetic acid (FAA), 10-carboxymethyl-9-acridone (CMA), di-amidobenzimidazole (di-ABZI), etc.73–75 STING agonists were found to improve the efficacy of NAFLD-HCC treatment (Table 4).5,76-80 The STING agonist 3′3′-cyclic adenine monophosphate-inosine monophosphate (cAIMP) was shown to reduce tumor burden by increasing tumor cell apoptosis early in HCC.5 STING agonist 2′,3′-cGAMP downregulated macrophage inhibitory receptor signal-regulatory protein alpha and inhibited the CD47/signal-regulatory protein alpha signaling axis, which mediates phagocytosis and escape of liver tumor cells, and enhances the phagocytosis of liver tumor cells by macrophages.76 To ensure the safety and efficacy of treatment, some studies have combined STING agonists with traditional antitumor therapies such as surgery and immunotherapy to have better therapeutic effectiveness, which has become a research hotspot. Irreversible electroporation, an ablation therapy, leads to tumor cell death through apoptosis. Compared with irreversible electroporation alone, STING agonist c-di-GMP combined with irreversible electroporation increased tumor-infiltrating IFN-γ/TNF-α production by CD4 and CD8 cells. The number of T cells induced a stronger antitumor immune response.77 As the first STING-based cancer vaccine in immunotherapy, STINGVAX can enhance T lymphocyte infiltration into cancer tissues by upregulating PDL-1.78 Combining DMXAA, cisplatin, tumor-specific peptides, neoantigens and an immune checkpoint inhibitor can induce the priming of tumor-specific CD8+ T cells to enhance the immune response, resulting in an anticancer immune synergistic effect.79 Synergistic immunotherapy with silk hydrogel containing interferon genes agonist, Hepa1-6 liver cancer-specific neoantigen and toll-like receptor 9 agonist, and mucin domain 3 antibody significantly reduced regulatory T cells and increased IFN-γ and IL-12p70 levels in tumor tissue, promoted IFN-γ+CD8+ T cell and 41BB+CD8+ T cell infiltration, and significantly inhibited HCC progression.80 Moreover, chemotherapy and radiotherapy combined with the STING pathway can inhibit tumor progression. Using paclitaxel, a commonly used chemotherapeutic drug in HCC, stabilizes microtubules. It interferes with mitosis, cGAS-dependent IRF3 phosphorylation accumulates, and the transcriptional mechanism promotes apoptosis. Combining cGAS-STING pathway activation and paclitaxel significantly inhibited tumor growth.81 Radiotherapy-induced DNA damage leakage into the cytosol can trigger the host antitumor immune response, using alginate as a carrier to deliver Mn2+ into radiotherapy-treated tumors. Synergistically accumulated cytosolic DNA can synergistically amplify the activation of the cGAS-STING pathway, thereby enhancing radiotherapy-induced antitumor immunity.82

Table 4

STING agonists for treating NAFLD-HCC

DrugCombinationDosage and method of administrationMechanismReference
3′3′-cAIMP/2 µg/g/3 times/week (i.p.)Increase tumor cell apoptosis by activating STING5
2′,3′-cGAMP/1 µg/mL (THP-1 cell culture)Enhance the phagocytosis of liver tumor cells by macrophages76
c-di-GMPIrreversible electroporation1 mg/mL (intratumoral injection)Increase tumor-infiltrating IFN-γ/TNF-α production of CD4 and CD877
CDA/CDGGranulocyte-macrophage colony-stimulating factor20 µg/mouse (s.c.)Enhance T lymphocyte infiltration into cancer tissues by upregulating PDL-178
DMXAACisplatin, tumor-specific peptides, neoantigens, and an immune checkpoint inhibitor100 µg/mouse (s.c.)Induce the priming of tumor-specific CD8+ T cells to enhance the immune response79
cGAMPHepa1-6 liver cancer-specific neoantigen and toll-like receptor 9 agonist, and TIM-3 antibody30 µg/mouse (i.p.)Reduce regulatory T cells and increase IFN-γ and IL-12p70 levels in tumor tissue, promote IFN-γ+CD8+ T cell and 41BB+CD8+ T cell infiltration80

Conclusion and future perspectives

Abundant evidence has established a close connection between innate immune activation and NAFLD.83 STING, being a crucial component of the innate immune system, is highly enriched in liver macrophages. Studies have explored how the STING signaling pathway mediates macrophage-mediated hepatic immune responses and metabolic regulation in NAFLD. While STING activation enhances immune surveillance in the liver during pathogen infection and tumors, its continuous activation can lead to a micro-inflammatory state in the liver. This, in turn, induces abnormalities in lipid metabolism, IR, and hepatocyte apoptosis, ultimately accelerating liver fibrosis and HCC development. Inhibiting STING has shown promise in treating inflammatory diseases, offering potential new avenues for NAFLD treatment. However, it is essential to consider potential side effects when activating the pathway, as STING’s effects can be double-edged—either suppressing cancer through early inflammation or promoting cancer through persistent chronic inflammation. Therefore, in addition to using STING agonists as monotherapy for tumors, combining STING agonists with traditional antitumor therapies such as surgery, chemotherapy, radiation therapy, and immunotherapy may minimize the negative effects. Although STING agonists or inhibitors have demonstrated encouraging results in treating NAFLD, there are still some deficiencies to be further studied. Firstly, there are currently no clinically relevant diagnostic indicators developed for the STING signaling pathway, and it is worth exploring the biochemical markers of STING in mouse models or in vitro samples. Secondly, research on the pathological mechanisms of the STING signaling pathway in NAFLD is still not comprehensive and needs further improvement. Furthermore, current research mainly focuses on basic experimental studies and lacks the application of clinical trials. Further scientific and standardized clinical research is needed to verify their effectiveness and safety, which will assist in the development of clinical treatment guidelines. Such research efforts will be critical in developing effective therapies for NAFLD and related conditions.

Abbreviations

AKT: 

protein kinase B

ATM: 

ataxia telangiectasia mutated

ATR: 

ATM and Rad3-related protein kinase

BAX: 

B-cell lymphoma-2-related protein X

Bcl2: 

B-cell lymphoma-2

cAIMP: 

cyclic-adenine monophosphate-inosine monophosphate

c-di-AMP: 

cyclic dimeric adenosine monophosphate

c-di-GMP: 

cyclic dimeric guanosine monophosphate

cGAMP: 

cyclic-guanosine monophosphate-adenosine monophosphate

cGAS: 

cGAMP synthase

Casp: 

caspase

CCL: 

chemokine ligand

CHK1: 

cell cycle checkpoint kinase 1

CMA: 

10-carboxymethyl-9-acridone

di-ABZI: 

di-amidobenzimidazole

DMXAA: 

5,6-dimethylxanthone-4-acetic acid

ER: 

endoplasmic reticulum

FAA: 

flavone acetic acid

FFA: 

free fatty acid

HCC: 

hepatocellular carcinoma

HSC: 

hepatic stellate cell

IFI16: 

interferon inducible protein 16

IFN-1: 

interferon-1

IKKε: 

inhibitor of NF-κB kinase epsilon

IL-6: 

interleukin-6

IR: 

insulin resistance

IRF3: 

interferon regulatory factor 3

mtDNA: 

mitochondrial DNA

NAFLD: 

nonalcoholic fatty liver disease

NASH: 

nonalcoholic steatohepatitis

NF-κB: 

nuclear factor-kappa B

NK: 

natural killer

PARP-1: 

poly (ADP-ribose) polymerase 1

PDL: 

programmed cell death ligand

SCAP: 

sterol regulatory element binding protein cleavage-activated protein

STING: 

stimulator of interferon gene

TBK1: 

TANK binding kinase 1

TNF-α: 

tumor necrosis factor-alpha

TRAF6: 

tumor necrosis factor receptor-associated factor 6

Declarations

Acknowledgement

The authors are grateful to the dedicated and committed participants in the study.

Funding

This work was supported by the Nature Science Foundation of Jiangsu province (BK20211388), Science and technology development plan project of Jiangsu Provincial Bureau of Traditional Chinese Medicine (ZT202207) (LC), and National Nature Science Foundation of China (NNSFC) 82274445 (YF).

Conflict of interest

The authors have no conflict of interest related to this publication.

Authors’ contributions

Contributed equally to this work (WL, ZZ), and provided intellectual input to the Editorial (All authors).

References

  1. Huang TD, Behary J, Zekry A. Non-alcoholic fatty liver disease: a review of epidemiology, risk factors, diagnosis and management. Intern Med J 2020;50(9):1038-1047 View Article PubMed/NCBI
  2. Buzzetti E, Pinzani M, Tsochatzis EA. The multiple-hit pathogenesis of non-alcoholic fatty liver disease (NAFLD). Metabolism 2016;65(8):1038-1048 View Article PubMed/NCBI
  3. Nati M, Chung KJ, Chavakis T. The Role of Innate Immune Cells in Nonalcoholic Fatty Liver Disease. J Innate Immun 2022;14(1):31-41 View Article PubMed/NCBI
  4. Wang X, Rao H, Zhao J, Wee A, Li X, Fei R, et al. STING expression in monocyte-derived macrophages is associated with the progression of liver inflammation and fibrosis in patients with nonalcoholic fatty liver disease. Lab Invest 2020;100(4):542-552 View Article PubMed/NCBI
  5. Thomsen MK, Skouboe MK, Boularan C, Vernejoul F, Lioux T, Leknes SL, et al. The cGAS-STING pathway is a therapeutic target in a preclinical model of hepatocellular carcinoma. Oncogene 2020;39(8):1652-1664 View Article PubMed/NCBI
  6. Carranza-Trejo AM, Vetvicka V, Vistejnova L, Kralickova M, Montufar EB. Hepatocyte and immune cell crosstalk in non-alcoholic fatty liver disease. Expert Rev Gastroenterol Hepatol 2021;15(7):783-796 View Article PubMed/NCBI
  7. Rada P, González-Rodríguez Á, García-Monzón C, Valverde ÁM. Understanding lipotoxicity in NAFLD pathogenesis: is CD36 a key driver?. Cell Death Dis 2020;11(9):802 View Article PubMed/NCBI
  8. Arrese M, Cabrera D, Kalergis AM, Feldstein AE. Innate Immunity and Inflammation in NAFLD/NASH. Dig Dis Sci 2016;61(5):1294-1303 View Article PubMed/NCBI
  9. Chen J, Vitetta L. Gut Microbiota Metabolites in NAFLD Pathogenesis and Therapeutic Implications. Int J Mol Sci 2020;21(15):E5214 View Article PubMed/NCBI
  10. Luo Z, Ji Y, Zhang D, Gao H, Jin Z, Yang M, et al. Microbial DNA enrichment promotes liver steatosis and fibrosis in the course of non-alcoholic steatohepatitis. Acta Physiol (Oxf) 2022;235(3):e13827 View Article PubMed/NCBI
  11. Pirola CJ, Garaycoechea M, Flichman D, Castaño GO, Sookoian S. Liver mitochondrial DNA damage and genetic variability of Cytochrome b - a key component of the respirasome - drive the severity of fatty liver disease. J Intern Med 2021;289(1):84-96 View Article PubMed/NCBI
  12. Luo Z, Ji Y, Gao H, Gomes Dos Reis FC, Bandyopadhyay G, Jin Z, et al. CRIg(+) Macrophages Prevent Gut Microbial DNA-Containing Extracellular Vesicle-Induced Tissue Inflammation and Insulin Resistance. Gastroenterology 2021;160(3):863-874 View Article PubMed/NCBI
  13. Zhong W, Rao Z, Xu J, Sun Y, Hu H, Wang P, et al. Defective mitophagy in aged macrophages promotes mitochondrial DNA cytosolic leakage to activate STING signaling during liver sterile inflammation. Aging Cell 2022;21(6):e13622 View Article PubMed/NCBI
  14. Wang Z, Chen N, Li Z, Xu G, Zhan X, Tang J, et al. The Cytosolic DNA-Sensing cGAS-STING Pathway in Liver Diseases. Front Cell Dev Biol 2021;9:717610 View Article PubMed/NCBI
  15. Bai J, Liu F. The cGAS-cGAMP-STING Pathway: A Molecular Link Between Immunity and Metabolism. Diabetes 2019;68(6):1099-1108 View Article PubMed/NCBI
  16. Ablasser A, Schmid-Burgk JL, Hemmerling I, Horvath GL, Schmidt T, Latz E, et al. Cell intrinsic immunity spreads to bystander cells via the intercellular transfer of cGAMP. Nature 2013;503(7477):530-534 View Article PubMed/NCBI
  17. Mao Y, Luo W, Zhang L, Wu W, Yuan L, Xu H, 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(5):920-929 View Article PubMed/NCBI
  18. Yoon SH, Waters CM. The ever-expanding world of bacterial cyclic oligonucleotide second messengers. Curr Opin Microbiol 2021;60:96-103 View Article PubMed/NCBI
  19. Wu J, Sun L, Chen X, Du F, Shi H, Chen C, et al. Cyclic GMP-AMP is an endogenous second messenger in innate immune signaling by cytosolic DNA. Science 2013;339(6121):826-830 View Article PubMed/NCBI
  20. Zhang C, Shang G, Gui X, Zhang X, Bai XC, Chen ZJ. Structural basis of STING binding with and phosphorylation by TBK1. Nature 2019;567(7748):394-398 View Article PubMed/NCBI
  21. Balka KR, Louis C, Saunders TL, Smith AM, Calleja DJ, D’Silva DB, et al. TBK1 and IKKε Act Redundantly to Mediate STING-Induced NF-κB Responses in Myeloid Cells. Cell Rep 2020;31(1):107492 View Article PubMed/NCBI
  22. Balka KR, De Nardo D. Molecular and spatial mechanisms governing STING signalling. FEBS J 2021;288(19):5504-5529 View Article PubMed/NCBI
  23. Chen H, Sun H, You F, Sun W, Zhou X, Chen L, et al. Activation of STAT6 by STING is critical for antiviral innate immunity. Cell 2011;147(2):436-446 View Article PubMed/NCBI
  24. Gui X, Yang H, Li T, Tan X, Shi P, Li M, et al. Autophagy induction via STING trafficking is a primordial function of the cGAS pathway. Nature 2019;567(7747):262-266 View Article PubMed/NCBI
  25. Moretti J, Roy S, Bozec D, Martinez J, Chapman JR, Ueberheide B, et al. STING Senses Microbial Viability to Orchestrate Stress-Mediated Autophagy of the Endoplasmic Reticulum. Cell 2017;171(4):809-823.e13 View Article PubMed/NCBI
  26. Gulen MF, Koch U, Haag SM, Schuler F, Apetoh L, Villunger A, et al. Signalling strength determines proapoptotic functions of STING. Nat Commun 2017;8(1):427 View Article PubMed/NCBI
  27. Sarhan J, Liu BC, Muendlein HI, Weindel CG, Smirnova I, Tang AY, et al. Constitutive interferon signaling maintains critical threshold of MLKL expression to license necroptosis. Cell Death Differ 2019;26(2):332-347 View Article PubMed/NCBI
  28. Dunphy G, Flannery SM, Almine JF, Connolly DJ, Paulus C, Jønsson KL, et al. Non-canonical Activation of the DNA Sensing Adaptor STING by ATM and IFI16 Mediates NF-κB Signaling after Nuclear DNA Damage. Mol Cell 2018;71(5):745-760.e5 View Article PubMed/NCBI
  29. Motwani M, Pesiridis S, Fitzgerald KA. DNA sensing by the cGAS-STING pathway in health and disease. Nat Rev Genet 2019;20(11):657-674 View Article PubMed/NCBI
  30. Bertran L, Adalid L, Vilaró-Blay M, Barrientos-Riosalido A, Aguilar C, Martínez S, et al. Expression of STING in Women with Morbid Obesity and Nonalcoholic Fatty Liver Disease. Metabolites 2023;13(4):496 View Article PubMed/NCBI
  31. Chen L, Xia S, Wang S, Zhou Y, Wang F, Li Z, et al. Naringenin is a Potential Immunomodulator for Inhibiting Liver Fibrosis by Inhibiting the cGAS-STING Pathway. J Clin Transl Hepatol 2023;11(1):26-37 View Article PubMed/NCBI
  32. 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 Invest 2019;129(2):546-555 View Article PubMed/NCBI
  33. Donne R, Saroul-Ainama M, Cordier P, Hammoutene A, Kabore C, Stadler M, et al. Replication stress triggered by nucleotide pool imbalance drives DNA damage and cGAS-STING pathway activation in NAFLD. Dev Cell 2022;57(14):1728-1741.e6 View Article PubMed/NCBI
  34. Li H, Hu L, Wang L, Wang Y, Shao M, Chen Y, et al. Iron Activates cGAS-STING Signaling and Promotes Hepatic Inflammation. J Agric Food Chem 2022;70(7):2211-2220 View Article PubMed/NCBI
  35. Chen C, Yang RX, Xu HG. STING and liver disease. J Gastroenterol 2021;56(8):704-712 View Article PubMed/NCBI
  36. Thomsen MK, Nandakumar R, Stadler D, Malo A, Valls RM, Wang F, et al. Lack of immunological DNA sensing in hepatocytes facilitates hepatitis B virus infection. Hepatology 2016;64(3):746-759 View Article PubMed/NCBI
  37. Lei Z, Deng M, Yi Z, Sun Q, Shapiro RA, Xu H, et al. cGAS-mediated autophagy protects the liver from ischemia-reperfusion injury independently of STING. Am J Physiol Gastrointest Liver Physiol 2018;314(6):G655-G667 View Article PubMed/NCBI
  38. Luo X, Li H, Ma L, Zhou J, Guo X, Woo SL, et al. Expression of STING Is Increased in Liver Tissues From Patients With NAFLD and Promotes Macrophage-Mediated Hepatic Inflammation and Fibrosis in Mice. Gastroenterology 2018;155(6):1971-1984.e4 View Article PubMed/NCBI
  39. Ma L. STING promotes the progression of macrophage-mediated nonalcoholic fatty liver disease [Dissertation]. Chongqing: Chongqing Medical University; 2019 View Article PubMed/NCBI
  40. Iracheta-Vellve A, Petrasek J, Gyongyosi B, Satishchandran A, Lowe P, Kodys K, et al. Endoplasmic Reticulum Stress-induced Hepatocellular Death Pathways Mediate Liver Injury and Fibrosis via Stimulator of Interferon Genes. J Biol Chem 2016;291(52):26794-26805 View Article PubMed/NCBI
  41. Qiao JT, Cui C, Qing L, Wang LS, He TY, Yan F, et al. Activation of the STING-IRF3 pathway promotes hepatocyte inflammation, apoptosis and induces metabolic disorders in nonalcoholic fatty liver disease. Metabolism 2018;81:13-24 View Article PubMed/NCBI
  42. Petrasek J, Iracheta-Vellve A, Csak T, Satishchandran A, Kodys K, Kurt-Jones EA, et al. STING-IRF3 pathway links endoplasmic reticulum stress with hepatocyte apoptosis in early alcoholic liver disease. Proc Natl Acad Sci U S A 2013;110(41):16544-16549 View Article PubMed/NCBI
  43. Miyao M, Kotani H, Ishida T, Kawai C, Manabe S, Abiru H, et al. Pivotal role of liver sinusoidal endothelial cells in NAFLD/NASH progression. Lab Invest 2015;95(10):1130-1144 View Article PubMed/NCBI
  44. Cho CS, Park HW, Ho A, Semple IA, Kim B, Jang I, et al. Lipotoxicity induces hepatic protein inclusions through TANK binding kinase 1-mediated p62/sequestosome 1 phosphorylation. Hepatology 2018;68(4):1331-1346 View Article PubMed/NCBI
  45. Huang X, Yao Y, Hou X, Wei L, Rao Y, Su Y, et al. Macrophage SCAP Contributes to Metaflammation and Lean NAFLD by Activating STING-NF-κB Signaling Pathway. Cell Mol Gastroenterol Hepatol 2022;14(1):1-26 View Article PubMed/NCBI
  46. Liu K, Qiu D, Liang X, Huang Y, Wang Y, Jia X, et al. Lipotoxicity-induced STING1 activation stimulates MTORC1 and restricts hepatic lipophagy. Autophagy 2022;18(4):860-876 View Article PubMed/NCBI
  47. Akhmetova K, Balasov M, Chesnokov I. Drosophila STING protein has a role in lipid metabolism. Elife 2021;10:e67358 View Article PubMed/NCBI
  48. Zhang W, Tang Z, Zhu X, Xia N, Zhao Y, Wang S, et al. TRAF1 knockdown alleviates palmitate-induced insulin resistance in HepG2 cells through NF-κB pathway. Biochem Biophys Res Commun 2015;467(3):527-533 View Article PubMed/NCBI
  49. Hu H, Zhao R, He Q, Cui C, Song J, Guo X, et al. cGAS-STING mediates cytoplasmic mitochondrial-DNA-induced inflammatory signal transduction during accelerated senescence of pancreatic β-cells induced by metabolic stress. FASEB J 2022;36(5):e22266 View Article PubMed/NCBI
  50. Hu HQ, Qiao JT, Liu FQ, Wang JB, Sha S, He Q, et al. The STING-IRF3 pathway is involved in lipotoxic injury of pancreatic β cells in type 2 diabetes. Mol Cell Endocrinol 2020;518:110890 View Article PubMed/NCBI
  51. Yong H, Wang S, Song F. Activation of cGAS/STING pathway upon TDP-43-mediated mitochondrial injury may be involved in the pathogenesis of liver fibrosis. Liver Int 2021;41(8):1969-1971 View Article PubMed/NCBI
  52. Wang Q, Bu Q, Liu M, Zhang R, Gu J, Li L, et al. XBP1-mediated activation of the STING signalling pathway in macrophages contributes to liver fibrosis progression. JHEP Rep 2022;4(11):100555 View Article PubMed/NCBI
  53. Li A, Yi M, Qin S, Song Y, Chu Q, Wu K. Activating cGAS-STING pathway for the optimal effect of cancer immunotherapy. J Hematol Oncol 2019;12(1):35 View Article PubMed/NCBI
  54. Zhang Y, Zhai Q, Feng X, Chen D, Lu Y, Hu J, et al. Cancer cell-intrinsic STING is associated with CD8 + T-cell infiltration and might serve as a potential immunotherapeutic target in hepatocellular carcinoma. Clin Transl Oncol 2021;23(7):1314-1324 View Article PubMed/NCBI
  55. Garcia-Diaz A, Shin DS, Moreno BH, Saco J, Escuin-Ordinas H, Rodriguez GA, et al. Interferon Receptor Signaling Pathways Regulating PD-L1 and PD-L2 Expression. Cell Rep 2017;19(6):1189-1201 View Article PubMed/NCBI
  56. Lemos H, Mohamed E, Huang L, Ou R, Pacholczyk G, Arbab AS, et al. STING Promotes the Growth of Tumors Characterized by Low Antigenicity via IDO Activation. Cancer Res 2016;76(8):2076-2081 View Article PubMed/NCBI
  57. Munn DH, Mellor AL. IDO in the Tumor Microenvironment: Inflammation, Counter-Regulation, and Tolerance. Trends Immunol 2016;37(3):193-207 View Article PubMed/NCBI
  58. Wu J, Dobbs N, Yang K, Yan N. Interferon-Independent Activities of Mammalian STING Mediate Antiviral Response and Tumor Immune Evasion. Immunity 2020;53(1):115-126.e5 View Article PubMed/NCBI
  59. Dou Z, Ghosh K, Vizioli MG, Zhu J, Sen P, Wangensteen KJ, et al. Cytoplasmic chromatin triggers inflammation in senescence and cancer. Nature 2017;550(7676):402-406 View Article PubMed/NCBI
  60. Haag SM, Gulen MF, Reymond L, Gibelin A, Abrami L, Decout A, et al. Targeting STING with covalent small-molecule inhibitors. Nature 2018;559(7713):269-273 View Article PubMed/NCBI
  61. Hansen AL, Buchan GJ, Rühl M, Mukai K, Salvatore SR, Ogawa E, 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 U S A 2018;115(33):E7768-E7775 View Article PubMed/NCBI
  62. Li S, Hong Z, Wang Z, Li F, Mei J, Huang L, et al. The Cyclopeptide Astin C Specifically Inhibits the Innate Immune CDN Sensor STING. Cell Rep 2018;25(12):3405-3421.e7 View Article PubMed/NCBI
  63. Hong Z, Mei J, Li C, Bai G, Maimaiti M, Hu H, et al. STING inhibitors target the cyclic dinucleotide binding pocket. Proc Natl Acad Sci U S A 2021;118(24):e2105465118 View Article PubMed/NCBI
  64. Siu T, Altman MD, Baltus GA, Childers M, Ellis JM, Gunaydin H, et al. Discovery of a Novel cGAMP Competitive Ligand of the Inactive Form of STING. ACS Med Chem Lett 2019;10(1):92-97 View Article PubMed/NCBI
  65. Feng M, Kong D, Guo H, Xing C, Lv J, Bian H, et al. Gelsevirine improves age-related and surgically induced osteoarthritis in mice by reducing STING availability and local inflammation. Biochem Pharmacol 2022;198:114975 View Article PubMed/NCBI
  66. Gao J, Zheng M, Wu X, Zhang H, Su H, Dang Y, et al. CDK inhibitor Palbociclib targets STING to alleviate autoinflammation. EMBO Rep 2022;23(6):e53932 View Article PubMed/NCBI
  67. Li YN, Su Y. Remdesivir attenuates high fat diet (HFD)-induced NAFLD by regulating hepatocyte dyslipidemia and inflammation via the suppression of STING. Biochem Biophys Res Commun 2020;526(2):381-388 View Article PubMed/NCBI
  68. Cao L, Xu E, Zheng R, Zhangchen Z, Zhong R, Huang F, et al. Traditional Chinese medicine Lingguizhugan decoction ameliorate HFD-induced hepatic-lipid deposition in mice by inhibiting STING-mediated inflammation in macrophages. Chin Med 2022;17(1):7 View Article PubMed/NCBI
  69. Huang Y, Qiu D, Qin Y. 14th National Conference on Immunology; 2021 Oct 21-24. Chengdu, China: ; 2021, 74 View Article PubMed/NCBI
  70. Huang Y, Liang W, Li K, Liao X, Chen J, Qiu X, et al. Sorafenib suppresses the activation of type I interferon pathway induced by RLR-MAVS and cGAS-STING signaling. Biochem Biophys Res Commun 2022;623:181-188 View Article PubMed/NCBI
  71. Wu R, Chen F, Chen L. Effects of probiotics on intestinal microbiota and STING signaling pathway in mice with nonalcoholic fatty liver disease[J]. Chinese Journal of Microecology 2022;34(11):1292-1298 View Article PubMed/NCBI
  72. Wu JJ, Zhao L, Hu HG, Li WH, Li YM. Agonists and inhibitors of the STING pathway: Potential agents for immunotherapy. Med Res Rev 2020;40(3):1117-1141 View Article PubMed/NCBI
  73. Gobbi S, Belluti F, Rampa A, Bisi A. Flavonoid-Inspired Vascular Disrupting Agents: Exploring Flavone-8-Acetic Acid and Derivatives in the New Century. Molecules 2021;26(14):4228 View Article PubMed/NCBI
  74. Cavlar T, Deimling T, Ablasser A, Hopfner KP, Hornung V. Species-specific detection of the antiviral small-molecule compound CMA by STING. EMBO J 2013;32(10):1440-1450 View Article PubMed/NCBI
  75. Ramanjulu JM, Pesiridis GS, Yang J, Concha N, Singhaus R, Zhang SY, et al. Design of amidobenzimidazole STING receptor agonists with systemic activity. Nature 2018;564(7736):439-443 View Article PubMed/NCBI
  76. Wei Z. cGAS-STING agonist enhances the phagocytosis of hepatoma cells by macrophages [Dissertation]. Jinan: Shandong University; 2021 View Article PubMed/NCBI
  77. Lasarte-Cia A, Lozano T, Cano D, Martín-Otal C, Navarro F, Gorraiz M, et al. Intratumoral STING Agonist Injection Combined with Irreversible Electroporation Delays Tumor Growth in a Model of Hepatocarcinoma. Biomed Res Int 2021;2021:8852233 View Article PubMed/NCBI
  78. Fu J, Kanne DB, Leong M, Glickman LH, McWhirter SM, Lemmens E, et al. STING agonist formulated cancer vaccines can cure established tumors resistant to PD-1 blockade. Sci Transl Med 2015;7(283):283ra52 View Article PubMed/NCBI
  79. Lee SE, Jang GY, Lee JW, Park SH, Han HD, Park YM, et al. Improvement of STING-mediated cancer immunotherapy using immune checkpoint inhibitors as a game-changer. Cancer Immunol Immunother 2022;71(12):3029-3042 View Article PubMed/NCBI
  80. Zhao Q, Wang Y, Zhao B, Chen H, Cai Z, Zheng Y, et al. Neoantigen Immunotherapeutic-Gel Combined with TIM-3 Blockade Effectively Restrains Orthotopic Hepatocellular Carcinoma Progression. Nano Lett 2022;22(5):2048-2058 View Article PubMed/NCBI
  81. Zierhut C, Yamaguchi N, Paredes M, Luo JD, Carroll T, Funabiki H. The Cytoplasmic DNA Sensor cGAS Promotes Mitotic Cell Death. Cell 2019;178(2):302-315.e23 View Article PubMed/NCBI
  82. Wang C, Sun Z, Zhao C, Zhang Z, Wang H, Liu Y, et al. Maintaining manganese in tumor to activate cGAS-STING pathway evokes a robust abscopal anti-tumor effect. J Control Release 2021;331:480-490 View Article PubMed/NCBI
  83. Bai J, Liu F. The cGAS-cGAMP-STING Pathway: A Molecular Link Between Immunity and Metabolism. Diabetes 2019;68(6):1099-1108 View Article PubMed/NCBI
  • Journal of Clinical and Translational Hepatology
  • pISSN 2225-0719
  • eISSN 2310-8819
  • Copyright © 2022 JCTH. All Rights Reserved.
  • Published by Xia & He Publishing Inc.
  • Address: 14090 Southwest Freeway, Suite 300, Sugar Land, Texas 77478, USA
  • Email: service@xiahepublishing.com