Interferon regulatory factor 3 mediates Poly(I:C)-induced innate immune response and apoptosis in non‑small cell lung cancer

  • Authors:
    • Liang Yi
    • Dan Sun
    • Qian Han
    • Zhonghui Liu
    • Zeng Zeng
    • Yanping Wu
    • Xiaoyu Chai
    • Xinmin Liu
  • View Affiliations

  • Published online on: March 5, 2018     https://doi.org/10.3892/ijo.2018.4300
  • Pages: 1623-1632
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Immunotherapy is considered one of the most promising treatments for lung cancer. The cell signalling molecules melanoma differentiation-associated protein 5 (MDA5) and retinoic acid-inducible gene I protein (RIG‑I) are essential receptors that recognise intracellular pathogen-associated nucleic acids, whereas interferon regulatory factor 3 (IRF3) controls the expression of innate immunity-associated genes in macrophages. However, the innate immune response to polyinosinic:polycytidylic acid [Poly(I:C)] in lung cancer remains to be elucidated. In the present study, western blot analysis, reverse transcription-quantitative polymerase chain reaction, RNA interference, IRF3 plasmid construction, ELISA and apoptosis analysis were employed to study the innate immune response and apoptosis of non‑small cell lung cancer (NSCLC) cells. Poly(I:C) transfection in NSCLC cells triggered apoptosis via the extrinsic apoptotic pathway, and activated the innate immune response by promoting interferon-β and C-X-C motif chemokine ligand 10 expression. Treatment with the IκB kinase ε/tumour necrosis factor receptor-associated factor family member-associated nuclear factor-κB activator-binding kinase 1 inhibitor BX795, which inhibits IRF3 phosphorylation, or transfection with small interfering RNA/short hairpin RNA to downregulate MDA5, RIG‑I or IRF3, prior to Poly(I:C) transfection inhibited the innate immune response and apoptotic pathway. Conversely, IRF3 overexpression promoted activation of the apoptotic pathway, thus indicating that the MDA5/RIG‑I/IRF3 axis may mediate responses to Poly(I:C) transfection. Furthermore, phosphorylation of the transcription factor signal transducer and activator of transcription 1 (STAT1) was associated with the alterations in IRF3 phosphorylation and apoptosis, thus suggesting that STAT1 may be involved in Poly(I:C)-induced apoptosis. In NSCLC surgical samples, MDA5, RIG‑I and IRF3 were highly expressed, whereas the expression levels of phosphorylated‑IRF3 were reduced. These findings indicated that the function of the MDA5/RIG‑I/IRF3 axis may be impaired in some lung cancers. In conclusion, the present findings suggested that the MDA5/RIG‑I/IRF3 axis, which is associated with innate immunity, is intact in NSCLC cells, and IRF3 is involved in regulating the apoptotic pathway in NSCLC cells.

Introduction

Polyinosinic:polycytidylic acid [Poly(I:C)] is a synthetic analogue of double-stranded RNA (dsRNA), which is recognised by pattern recognition receptors, and elicits innate and adaptive immunity (1). Poly(I:C) exhibits potential by activating various types of immune cells, including dendritic cells, T helper (Th)1 cells and Th17 cells (2). The pathways that initiate immune activity mainly involve intact intracellular or extracellular signalling. Intracellularly, dsRNA is recognised by retinoic acid-inducible gene I protein (RIG-I)-like receptors [melanoma differentiation-associated protein 5 (MDA5) and RIG-I] that interact with the interferon (IFN)-β promoter stimulator 1 (IPS-1). Subsequently, IPS-1, which is a mitochondrial signalling adaptor protein, recruits and phosphorylates IκB kinase ε/tumour necrosis factor (TNF) receptor-associated factor family member-associated nuclear factor (NF)-κB activator-binding kinase 1 (TBK1), thus leading to the phosphorylation of interferon regulatory factor 3 (IRF3). Extracellularly, Toll-like receptors (TLRs) are the pivotal receptors. Upon Poly(I:C) binding to TLRs, the TLRs recruit adaptor proteins, including myeloid differentiation primary response 88, Toll/interlukin-1 receptor-domain-containing adapter-inducing IFN-β and translocating chain-associated membrane protein, and IRF3 is phosphorylated via NF-κB signalling followed by the activation of type I IFN signalling (3,4). A previous study revealed that Poly(I:C) exerts its antitumour properties in a type I IFN-dependent or IFN-independent manner in MCF10A, MDAMB-231, IMR32 and HEK293T cells (5).

MDA5 and RIG-I are cytoplasmic sensors of Poly(I:C), which participate in the Poly(I:C)-induced apoptosis of cancer cells by eliciting the canonical apoptotic signalling pathway, and may even boost innate immune responses (6). IRF3 is a downstream molecule of RIG-I/MDA5, which is usually located in the cytoplasm in its inactive form, and is a vital transcription factor for type I IFN. It is activated by phosphorylation at its carboxyl terminus in response to viral infection, DNA-damaging stress, ultraviolet light and chemical stimulation. Upon phosphorylation, IRF3 dimerizes and translocates into the nucleus, where its target genes are transcribed (7,8). IRF3 exerts dual roles in regulating the bioactivity of cancer. In previous studies, IRF3 has been reported to act as an antitumour factor due to its ability to control cancer cell progression (9,10). For example, IRF3 inhibits the growth of nasopharyngeal carcinoma cells, B16 melanoma cells, prostate cancer cells and malignant glioma cells via regulating the activation of natural killer cells or type I IFN signalling (9,11-13). In addition, it has previously been reported that Poly(I:C)-induced apoptosis of prostate cancer PC3 and DU145 cells is independent of IRF3 (14). Conversely, the mRNA and protein expression levels of IRF3 are upregulated in patients with acute myeloid leukaemia (AML), and IRF3 promotes the proliferation and survival of AML cells (15). In lung cancer tissues, IRF3 is aberrantly expressed in either the cytoplasm or nucleus, and its expression status is not associated with sex, histological grade, nodal metastasis, pathological stage or recurrence (16). Therefore, the present study aimed to explore the effects of Poly(I:C) on lung cancer cells and to reveal the probable role of IRF3 in these effects.

The present study demonstrated that Poly(I:C) triggered innate immunity via an intracellular signalling pathway, and induced apoptosis via the extrinsic apoptotic pathway in non-small cell lung cancer (NSCLC) cells. In addition, the results indicated that IRF3, activated by MDA5/RIG-I, may regulate the Poly(I:C)-induced apoptotic pathway. Finally, the results demonstrated that molecules in the innate immune pathway were synchronously upregulated; however, this was not accompanied by an increase in IRF3 phosphorylation, in tumour tissue samples obtained from patients with NSCLC compared with in matched adjacent tissues.

Materials and methods

Tissue specimen collection

NSCLC tissues and corresponding adjacent tissues were obtained from eight patients at Peking University First Hospital (Beijing, China). Patient characteristics are listed in Table I. The tissues were instantly frozen in liquid nitrogen once they were removed from the patients and were preserved at −80°C. All of the subjects provided informed consent for their inclusion in the present study prior to participation. The present study was conducted in accordance with the Declaration of Helsinki, and was approved by the Ethics Committee of Peking University First Hospital.

Table I

Clinical information of patients with non-small cell lung cancer.

Table I

Clinical information of patients with non-small cell lung cancer.

PatientsSex (M/F)Age (years)Pathological typeLymphatic metastasis
#1F58AdenocarcinomaNo
#2F78Squamous-cell carcinomaLobar bronchus
#3M59AdenocarcinomaNo
#4M60AdenocarcinomaNo
#5F80Combined small cell carcinoma and adenocarcinomaNo
#6F53AdenocarcinomaNo
#7M57AdenocarcinomaNo
#8F57AdenocarcinomaNo

[i] F, female; M, male.

Cell culture and drug pretreatment

The NSCLC cell line A549 was obtained from American Type Culture Collection (Manassas, VA, USA), and NCI-H1299 cells were obtained from Peking Union Medical College (Beijing, China). The A549 cells were cultured with Dulbecco’s modified Eagle’s medium (Gibco; Thermo Fisher Scientific, Inc., Waltham, MA, USA) and the H1299 cells were cultured with Roswell Park Memorial Institute 1640 medium (Gibco; Thermo Fisher Scientific, Inc.). Both media were supplemented with 10% foetal bovine serum (Corning Incorporated, Corning, NY, USA) at 37°C in a humidified incubator containing 5% CO2.

Drug pretreatment and Poly(I:C) transfection

Once cell confluence reached 70–80% in the 6-well plates, the pan-caspase inhibitor Z-VAD-FKM (50 μM; InvivoGen, San Diego, CA, USA), the caspase-8 inhibitor Z-IETD-FKM (25 μM; EMD Millipore, Billerica, MA, USA) or the TBK1 inhibitor BX795 (1 μM); (InvivoGen) were used to treat the cells for 2 h at 37°C. Subsequently, Poly(I:C) (0, 100, 200 and 400 ng/ml; Sigma-Aldrich; Merck KGaA, Darmstadt, Germany) was transfected into the cells (confluence, 70–80%) for 6 h using Lipofectamine® 3000 (Invitrogen; Thermo Fisher Scientific, Inc.) according to the manufacturer’s protocol, after which the cells were cultured for 24 h.

Western blot analysis

The cells/tissues were lysed in radioimmunoprecipitation assay buffer containing phenylmethylsulfonyl fluoride and phosphatase inhibitors (Roche Diagnostics GmbH, Mannheim, Germany) for 5 min. The protein content in the cell lysates was determined using a bicinchoninic acid protein assay kit (Thermo Fisher Scientific, Inc.). The lysates, containing 40 μg protein, were separated by 12% SDS-PAGE, and proteins were transferred to a polyvinylidene fluoride membrane. Subsequently, the membrane was blocked with 5% blocking reagent (2 g skimmed milk power dissolved in 40 ml Tris-buffered saline with 1% Tween-20) for 1 h at room temperature. Then it was incubated with the following primary antibodies (1:1,000 dilutions) overnight at 4°C: Anti-IRF3 (ab50772; Abcam, Cambridge, UK), anti-phosphorylated (p)-IRF3 (p-S386) (ab76493; Abcam), anti-MDA5 (ab126630; Abcam), anti-RIG-I (ab180675; Abcam), anti-NAK/TBK1 (ab109735; Abcam), anti-NAK/TBK1 (p-S172) (ab109272; Abcam), anti-GAPDH (ab9485; Abcam), anti-caspase-3 (622701; BioLegend, Inc., San Diego, CA, USA), anti-caspase-9 (621901; BioLegend, Inc.), anti-caspase-8 (645501; BioLegend, Inc.), anti-TNF-related apoptosis-inducing ligand (TRAIL) antibody (RLT4721; Suzhou Ruiying Biological Co., Ltd., Suzhou, China) and anti-p-STAT1 (p-Ser727) antibody (RLP0842; Suzhou Ruiying Biological Co., Ltd.). Subsequently, the membranes were incubated with goat anti-mouse immunoglobulin G (IgG)-horseradish peroxidase (HRP) (sc-2005; 1:5,000 dilution; Santa Cruz Biotechnology, Inc., Dallas, TX, USA) and goat anti-rabbit IgG-HRP (sc2004; 1:5,000 dilution; Santa Cruz Biotechnology, Inc.) secondary antibodies for 2 h at room temperature. The blots were visualised using the Luminata Forte Western HRP Substrate (EMD Millipore).

Reverse transcription-quantitative polymerase chain reaction (RT-qPCR)

Total RNA was extracted from cells or tissues using TRIzol® reagent (Invitrogen; Thermo Fisher Scientific, Inc.). Total RNA (2 μg) was reverse transcribed to cDNA using a first-strand cDNA synthesis kit (Invitrogen; Thermo Fisher Scientific, Inc.), which was conducted according to the manufacturer’s protocol. Subsequently, relative gene expression was determined by qPCR using the SYBR kit (Applied Biosystems; Thermo Fisher Scientific, Inc.) and ABI 7500 real-time PCR detection system (Applied Biosystems; Thermo Fisher Scientific, Inc.). The amplification conditions were as follows: An initial denaturation step at 95°C for 10 min, followed by 40 cycles of denaturation at 95°C for 15 sec and annealing/extending at 60°C for 1 min. The following primer sequences were used: IRF3 forward, 5′-TCAGGAGTTGGGGACTTTTC and reverse, 5′-GAATGTCTTCCTGGGTATCAG; and β-actin forward, 5′-ATATCGCCGCGCTCGTCGTC and reverse, 5′-CATGCCCACCATCACGCC CTG-3′. The MDA5 and RIG-1 primer sequences used are noted in previous studies (17,18), as are the sequences for C-X-C motif chemokine ligand 10 (CXCL-10) and IFN-β (19), and TRAIL (5). β-actin was used as an internal reference. The relative mRNA expression levels were normalised to β-actin mRNA. The results of the cell samples were calculated as 2−ΔΔCq (20), whereas the 2−ΔCq method was used to analyse the tissue samples.

ELISA

After the indicated treatments, the supernatants were collected and an ELISA analysis was performed to detect the expression levels of human CXCL-10 using an ELISA kit (555046; BD Biosciences, Franklin Lakes, NJ, USA) according to the manufacturer’s protocol.

Apoptosis analysis

An equal number of A549 cells (1.0×105/ml) were seeded into a 6-well plate. Following transfection of A549 cells with Poly(I:C) for 24 h, with or without caspase inhibitor pretreatment, the cells were harvested (1.0×106). Subsequently, the apoptotic cells were stained using the Annexin V-fluorescein isothiocyanate/propidium iodide Cell Apoptosis Detection kit (Beijing Transgen Biotech Co., Ltd., Beijing, China) according to the manufacturer’s protocol. Each sample was assessed by fluorescence-activated cell sorting using a flow cytometer (BD Biosciences) and the data were analyzed using FlowJo v10 software (FlowJo LLC, Ashland, OR, USA).

Lentivirus short hairpin (sh)RNA infection

The lentivirus shRNA pLent-U6-shIRF3-GFP-Puro and the empty vector pLent-U6-GFP-Puro were purchased from Vigene Biosciences, Inc. (Jinan, China). The sequence of IRF3 shRNA was as follows: 5′-CCGGCCCTTCATTGTAGATCTGATTCTCGAGAATCAGATCTACAATGAAGGGTTTTT-3′. A549 cells were infected with the lentiviruses once cell confluence reached 70-80%. A total of 48 h postinfection, a low concentration of puromycin (5 mg/ml) was used to select lentivirus-infected cells. After 14 days of screening, the stably infected cells were obtained.

Plasmid construction

IRF3 cDNA was obtained from peripheral blood cells collected from a healthy human volunteer, and PCR was performed with primers tagged with restriction endonuclease sequences (XhoI and AgeI; Thermo Fisher Scientific, Inc.). Subsequently, the PCR products, or the vector plasmid pEGFP-N1 (Clontech Laboratories, Inc., Mountainview, CA, USA), were incubated with XhoI or AgeI for 1.5 h and were ligated with T4 ligase for 1.5 h (Thermo Fisher Scientific, Inc.) at room temperature. Following ligation, the mix was transformed into competent Escherichia coli (Beijing Transgen Biotech Co., Ltd.) cells by heat shock treatment (42°C for 90 sec in a water bath). Finally, the transformed bacteria were cultured with Lennox L Broth Base (Invitrogen; Thermo Fisher Scientific, Inc.) containing kanamycin (30 μg/ml; Gibco; Thermo Fisher Scientific, Inc.) and the recombinant plasmid IRF3-pEGFP-N1 was identified by PCR using the IRF3 primer.

Plasmid and small interfering (si)RNA transfection

Once cell confluence reached 80–90%, the expression plasmid IRF3-pEGFP-N1 and the empty vector were transfected into A549 cells, whereas the IRF3-pEnter with His-tag (Vigene Biosciences, Inc.) and the empty vector were transfected into H1299 cells using Lipofectamine® 3000. The culture medium was changed after 6 h, and IRF3 expression was assessed after 48 h of transfection. For siRNA transfection, the RIG-I- and MDA5-specific siRNAs, and the scrambled sequence were prepared (Suzhou GenePharma, LLC, Suzhou, China). The sequences used were as follows: Scrambled, 5′-CAUAGCGUCCUUGAUCACAUU-3′; RIG-I, 5′-AACGAUUCCAUCACUAUCCAUdTdT-3′ (21); MDA5, 5′-GGUGUAAGAGAGCUACUAAtt-3′ (22). siRNA transfection was performed using Lipofectamine® 3000 once cell confluence reached 70–80%, and the concentration of siRNAs used was 50 nM. Subsequently, cells were analysed after 36 h of transfection.

Statistical analysis

Statistical analyses were performed using SPSS 20 (IBM Corporation, Armonk, NY, USA). The data were presented as the means ± standard error of the mean. Semi-quantification of western blotting was analysed using ImageJ2 software (National Institutes of Health, Bethesda, MD, USA). For comparisons between two groups, Student’s t-test was used, whereas comparisons among multiple groups were made using univariate analysis of variance, followed by least significant difference post hoc test. P<0.05 was considered to indicate a statistically significant difference. Three independent experiments were conducted.

Results

Poly(I:C) induces apoptosis via the extrinsic apoptotic pathway in A549 cells

It was hypothesised that Poly(I:C) may induce apoptosis of NSCLC cells via the canonical apoptotic pathway. The present study demonstrated that the number of apoptotic A549 cells increased in a concentration-dependent manner following transfection with Poly(I:C) for 24 h (Fig. 1A and B). The mRNA expression levels of TRAIL were significantly increased (data not shown), and the protein expression levels of TRAIL, and cleaved caspase-3 and -8, were markedly increased. Conversely, cleaved caspase-9 expression was slightly increased post-transfection (Fig. 1C). To further determine the role of caspases in Poly(I:C)-induced apoptosis of A549 cells, the cells were treated with the pancaspase inhibitor Z-VAD-FMK and the caspase-8 inhibitor Z-IETD-FMK for 2 h prior to transfection with 100 ng/ml Poly(I:C) for 24 h. As expected, apoptosis was significantly inhibited (Fig. 1D and E) and cleaved caspase-8 expression was decreased following Z-VAD-FMK pretreatment (Fig. 1F). These findings indicated that Poly(I:C) may induce apoptosis of A549 cells via activation of an extrinsic apoptotic pathway.

Poly(I:C) transfection activates innate immunity in NSCLC cells

The present study transfected Poly(I:C) into the NSCLC cell lines A549 and H1299. Subsequently, the mRNA expression levels of IFN-β and CXCL-10 were increased in both cell lines, and the secretion of CXCL-10 was increased in A549 cells in a concentration-dependent manner (Fig. 2). Furthermore, MDA5/RIG-I expression was upregulated, and their downstream molecules, TBK1 and IRF3, were activated (p-TBK1, p-IRF3) (data not shown). However, these alterations were not observed when Poly(I:C) was directly added to the medium (data not shown).

MDA5/RIG-I/IRF3 axis is associated with the effects of Poly(I:C) transfection on NSCLC cell lines

When the NSCLC A549 and H1299 cell lines were pretreated with BX795 for 2 h, and were then transfected with Poly(I:C) for 24 h, p-IRF3 was inhibited, whereas no changes in total IFR3 were detected (Fig. 3A). In addition, poly(I:C)-induced increases in IFN-β and CXCL-10 mRNA expression in both cell lines, and CXCL-10 secretion in A549 cells, were abrogated by BX795 (Fig. 3B–D). Furthermore, IRF3 was stably knocked down by shRNA in A549 cells, which inhibited the increase in p-IFR3 levels and CXCL-10 secretion following Poly(I:C) transfection (Fig. 3E and F). Furthermore, both cell lines were transfected with Poly(I:C) for 24 h following transfection with RIG-I and MDA5 siRNAs. Notably, the expression levels of p-IRF3 were decreased (Fig. 3G), which was accompanied by a decrease in IFN-β and CXCL-10 mRNA expression, and CXCL-10 secretion in A549 cells (Fig. 3F, H and I). These results indicated that responses to Poly(I:C) transfection were regulated by the MDA5/RIG-I/IRF3 axis in NSCLC cells.

Figure 3

MDA5/RIG-I/IRF3 axis mediates the responses of NSCLC cell lines to Poly(I:C) transfection. (A–D) Poly(I:C) (200 ng/ml) was transfected into A549 and H1299 cells for 24 h following 2 h pretreatment with BX795. DMSO was used as a control. (A) Protein expression levels of IRF3 and p-IRF3 were examined by western blotting. (B and C) IFN-β and CXCL-10 expression was quantified by qPCR. (D) CXCL-10 secretion was determined in A549 cell supernatants by ELISA. (E) IRF3 was stably downregulated by infection with a lentivirus containing shIFR3, the protein expression levels of p-IRF3 were detected by western blotting after Poly(I:C) transfection for 24 h. (F–I) siRNAs against RIG-I or MDA5 were transfected into cells for 36 h prior to Poly(I:C) transfection (200 ng/ml) for 24 h. Scrambled sequences were used as a negative control. (F) CXCL-10 was detected by ELISA in A549 cells with RIG-I or MDA5 or IRF3 depletion after Poly(I:C) transfection for 24 h. (G) Western blotting was used to examine the protein expression levels of RIG-I, MDA5, IRF3 and p-IRF3. (H and I) IFN-β and CXCL-10 mRNA expression was assessed by qPCR in A549 cells with RIG-I or MDA5 depletion following Poly(I:C) transfection for 24 h. *P<0.05; **P<0.005. CXCL-10, C-X-C motif chemokine ligand 10; DMSO, dimethyl sulfoxide; IFN-β, interferon-β; IRF3, interferon regulatory factor 3; MDA5, melanoma differentiation-associated protein 5; p-IRF3, phosphorylated-IRF3; Poly(I:C), polyinosinic:polycytidylic acid; qPCR, quantitative polymerase chain reaction; RIG-I, retinoic acid-inducible gene I protein; sh, short hairpin RNA; si, small interfering RNA.

IRF3 mediates apoptosis of NSCLC cells via STAT1 phosphorylation

The present study initially treated NSCLC cells with BX795 or shRNA to downregulate IRF3; the results demonstrated that inhibition of IRF3 suppressed Poly(I:C)-induced TRAIL, and cleaved caspase-3 and -8 expression, and decreased p-STAT1 (Fig. 4A and B). Conversely, IRF3 overexpression in NSCLC cells induced upregulation of TRAIL, cleaved caspase-3 and -8, and p-STAT1 (Fig. 5). Furthermore, knockdown of MDA5 or RIG-1 reduced TRAIL expression in both cell lines (data not shown). These results suggested that Poly(I:C) activated the apoptotic pathway in NSCLC cells via the MDA5/RIG-I/IRF3 pathway. In addition, STAT1 phosphorylation may also engage with IRF3-mediated apoptosis of NSCLC cells.

Activation of IRF3 is inhibited in NSCLC surgical samples

The present study collected surgical NSCLC samples from eight patients. The results of qPCR revealed that the mRNA expression levels of MDA5 were increased by ~3-fold in NSCLC tissues compared with in adjacent normal tissues (Fig. 6A). Conversely, RIG-1 mRNA expression exhibited no marked alterations in NSCLC tissues compared with in adjacent normal tissues (Fig. 6B). Western blotting demonstrated that MDA5, RIG-1 and IRF3 were increased in tumour tissues (Fig. 6C). However, p-IRF3 was decreased in the three tumour tissues compared with in the six normal samples (Fig. 6D). These results suggested that innate immunity may be impaired in NSCLC.

Discussion

Poly(I:C) is a synthetic analogue of dsRNA, which may be used as a vaccine adjuvant to boost innate and adaptive immunity (6,23,24). A vast array of genes are not only induced but also repressed by dsRNA. The inducible genes include interferon-stimulated genes, apoptosis-associated genes, cytokines and growth factor genes, and genes associated with metabolism and biosynthesis, RNA synthesis, protein synthesis and degradation, cytoskeletal components, transporters and the extracellular matrix. The repressed genes include those involved in metabolism, cell cycle regulation and cell adhesion (25). In the present study, the results indicated that the MDA5/RIG-I innate immunity pathway was activated and that type I IFN was induced when Poly(I:C) was transfected into NSCLC cells, which indicated that intracellular Poly(I:C) elicited innate immune responses by activating the MDA5/RIG-I pathway (24). However, when Poly(I:C) was added to the medium directly, it did not induce the innate immune response in NSCLC cells, which may result from a deficiency of TLR3 expression, which is the mainstay sensor for extracellular Poly(I:C) (3). However, the exact mechanism requires further investigation. The TBK1 inhibitor BX795 inhibits phosphoinositide-dependent kinase-1 and TBK1 and, subsequently blocks phosphorylation of IRF3 (26). Treatment with BX795 or knockdown of IRF3 prior to Poly(I:C) transfection led to the inhibition of innate immunity. These findings suggested that the MDA5/RIG-I/TBK1/IRF3 signalling pathway was initiated by intracellular Poly(I:C), and was intact in NSCLC cells, as in other cancer cells (14,27,28).

Previous studies have revealed that Poly(I:C) induces apoptosis of prostate cancer cells, pancreatic cancer cells and other cells (27,28). Furthermore, the canonical apoptotic pathways, including the caspase-9-dependent intrinsic pathway and the caspase-8-dependent extrinsic pathway, engage Poly(I:C)-induced apoptosis to some degree (14,27). In the present study, Poly(I:C)-induced apoptosis of A549 cells mainly via the extrinsic pathway, as determined by the slight change in activated caspase-9 expression, and this was associated with an upregulation of TRAIL. TRAIL belongs to the TNF family and induces apoptosis by binding death receptor (DR)4 or DR5, in order to activate the extrinsic apoptotic pathway (29,30). Furthermore, IRF3 serves a crucial role in TLR3-mediated apoptosis of androgen-sensitive prostate cancer LNCaP cells (12), whereas it is not implicated in androgen-resistant prostate cancer cells (14). In the present study, downregulation of IRF3 in NSCLC cells resulted in a decline in the expression of apoptosis-related proteins, and vice versa. A previous study revealed that IRF3 binds and transactivates the TRAIL promoter in the nucleus where the IRF3 response element is located (31). Therefore, active IRF3 may shuttle into the nucleus of NSCLC cells and directly target TRAIL to activate extrinsic apoptosis. However, another study elucidated the concept of RIG-I-induced IRF3-mediated apoptosis, revealing that IRF3 in the cytoplasm interacts with B-cell lymphoma 2-associated X protein and causes activation of caspase-9 and -3 (24), which mainly activate intrinsic apoptosis. Therefore, in the present study, it is likely that the results are consistent with the first mechanism. Notably, the present study demonstrated that variation of STAT1 phosphorylation was in accordance with the alterations in the apoptotic pathway. This finding suggested that activation of STAT1 by type I IFN may affect DNA damage and adaptive immunity (32), and may modulate the expression of FAS, FAS ligand and casapse-1, and the function of p53 to accelerate apoptosis (33). In addition, STAT1 suppresses the cell cycle by inducing p53, NF-κB p65, cyclin A, cyclin D1, cyclin E, F-box and WD repeat domain containing 7, Hes family BHLH transcription factor 1 and cyclin-dependent kinase 2 expression. Conversely, the Janus kinase (JAK)2/STAT1 pathway participates in the upregulation of programmed death-ligand 1 and may serve a pivotal role in inducing tumour immune escape (34). Generally, IRF3 is indeed involved in Poly(I:C)-induced apoptosis of NSCLC cells; however, the exact mechanism is ambiguous, and whether the JAK/STAT1 pathway orchestrates IRF3-mediated apoptosis in NSCLC remains to be determined.

In human tissue samples, the present study demonstrated that MDA5, RIG-I and IRF3 were increased in NSCLC samples, thus indicating that the innate immune pathway was intact in NSCLC. Notably, p-IRF3 expression was decreased, which was inconsistent with the increase of IRF3, in three samples. These findings suggested that MDA5 and RIG-I may be positively correlated with the overall survival of patients with NSCLC, similar to in patients with hepatocellular carcinoma (35). However, IRF3 may also be favourable to the survival of cancer cells via producing various inflammatory factors into the tumour microenvironment (36). In patients with herpes simplex encephalitis, a G-A mutation at the 854 base-pair region in exon 6 prevents IRF3 from forming homodimers and being phosphorylated at S286, which disrupts its transcriptional activity and the innate immune response (37). Whether the same mutation occurs in patients with NSCLC is currently unknown.

In conclusion, the present study revealed that the innate immune pathway was intact in NSCLC cells, and IRF3 was involved in regulating the apoptotic pathway. However, in NSCLC tissue samples, the innate immunity pathway may be disrupted. These findings may provide novel insights into the role of IRF3 in innate immunity and apoptosis in NSCLC. In addition, IRF3 may be considered a target to induce apoptosis in NSCLC therapy.

Acknowledgments

The authors would like to thank Professor Dingfang Bu for his expert technical guidance.

Abbreviations:

MDA5

melanoma differentiation-associated protein 5

RIG-I

retinoic acid-inducible gene I protein

IRF3

interferon regulatory factor 3

IPS-1

interferon-β promoter stimulator 1

TBK1

IκB kinase ε/tumour necrosis factor receptor-associated factor family member-associated nuclear factor-κB activator-binding kinase 1

TLRs

Toll-like receptors

Notes

[1] Funding

The present study was supported by the Beijing Natural Science Foundation (grant no. 7151011).

[2] Availability of data and materials

The datasets used and/or analyzed during the current study are available from the corresponding author on reasonable request.

[3] Authors’ contributions

XL conceived and designed the experiments; LY performed the experiments, analysed the data and wrote the paper; DS ZL and ZZ helped collect the tissue specimens; QH, YW and XC conducted statistical analysis.

[4] Ethics approval and consent to participate

The present study was approved by the Ethics Committee of Peking University First Hospital (Beijing, China). All of the subjects gave their informed consent for inclusion before they participated in the study. The peripheral blood cells were collected from a healthy volunteer and informed consent was obtained as well.

[5] Consent for publication

Not applicable.

[6] Competing interests

The authors declare that they have no competing interests.

References

1 

Jin B, Cheng LF, Wu K, Yu XH and Yeo AE: Application of dsRNA in cancer immunotherapy: Current status and future trends. Anticancer Agents Med Chem. 14:241–255. 2014. View Article : Google Scholar

2 

Jin B, Sun T, Yu XH, Liu CQ, Yang YX, Lu P, Fu SF, Qiu HB and Yeo AE: Immunomodulatory effects of dsRNA and its potential as vaccine adjuvant. J Biomed Biotechnol. 2010:6904382010. View Article : Google Scholar : PubMed/NCBI

3 

Cui J, Chen Y, Wang HY and Wang RF: Mechanisms and pathways of innate immune activation and regulation in health and cancer. Hum Vaccin Immunother. 10:3270–3285. 2014. View Article : Google Scholar

4 

Kawai T and Akira S: Signaling to NF-kappaB by Toll-like receptors. Trends Mol Med. 13:460–469. 2007. View Article : Google Scholar : PubMed/NCBI

5 

Kumar S, Ingle H, Mishra S, Mahla RS, Kumar A, Kawai T, Akira S, Takaoka A, Raut AA and Kumar H: IPS-1 differentially induces TRAIL, BCL2, BIRC3 and PRKCE in type I interferons-dependent and -independent anticancer activity. Cell Death Dis. 6:e17582015. View Article : Google Scholar : PubMed/NCBI

6 

Zhao X, Ai M, Guo Y, Zhou X, Wang L, Li X and Yao C: Poly I:C-induced tumor cell apoptosis mediated by pattern-recognition receptors. Cancer Biother Radiopharm. 27:530–534. 2012. View Article : Google Scholar : PubMed/NCBI

7 

Clément JF, Bibeau-Poirier A, Gravel SP, Grandvaux N, Bonneil E, Thibault P, Meloche S and Servant MJ: Phosphorylation of IRF-3 on Ser 339 generates a hyperactive form of IRF-3 through regulation of dimerization and CBP association. J Virol. 82:3984–3996. 2008. View Article : Google Scholar : PubMed/NCBI

8 

Escalante CR, Nistal-Villán E, Shen L, García-Sastre A and Aggarwal AK: Structure of IRF-3 bound to the PRDIII-I regulatory element of the human interferon-beta enhancer. Mol Cell. 26:703–716. 2007. View Article : Google Scholar : PubMed/NCBI

9 

Tarassishin L and Lee SC: Interferon regulatory factor 3 alters glioma inflammatory and invasive properties. J Neurooncol. 113:185–194. 2013. View Article : Google Scholar : PubMed/NCBI

10 

Romieu-Mourez R, Solis M, Nardin A, Goubau D, Baron-Bodo V, Lin R, Massie B, Salcedo M and Hiscott J: Distinct roles for IFN regulatory factor (IRF)-3 and IRF-7 in the activation of antitumor properties of human macrophages. Cancer Res. 66:10576–10585. 2006. View Article : Google Scholar : PubMed/NCBI

11 

Duan Y, Li Z, Cheng S, Chen Y, Zhang L, He J, Liao Q, Yang L, Gong Z and Sun LQ: Nasopharyngeal carcinoma progression is mediated by EBER-triggered inflammation via the RIG-I pathway. Cancer Lett. 361:67–74. 2015. View Article : Google Scholar : PubMed/NCBI

12 

Gambara G, Desideri M, Stoppacciaro A, Padula F, De Cesaris P, Starace D, Tubaro A, Del Bufalo D, Filippini A, Ziparo E, et al: TLR3 engagement induces IRF-3-dependent apoptosis in androgen-sensitive prostate cancer cells and inhibits tumour growth in vivo. J Cell Mol Med. 19:327–339. 2015. View Article : Google Scholar :

13 

Moore TC, Kumm PM, Brown DM and Petro TM: Interferon response factor 3 is crucial to poly-I:C induced NK cell activity and control of B16 melanoma growth. Cancer Lett. 346:122–128. 2014. View Article : Google Scholar :

14 

Palchetti S, Starace D, De Cesaris P, Filippini A, Ziparo E and Riccioli A: Transfected poly(I:C) activates different dsRNA receptors, leading to apoptosis or immunoadjuvant response in androgen-independent prostate cancer cells. J Biol Chem. 290:5470–5483. 2015. View Article : Google Scholar : PubMed/NCBI

15 

Tian WL, Jiang ZX, Wang F, Guo R, Tang P, Huang YM and Sun L: IRF3 is involved in human acute myeloid leukemia through regulating the expression of miR-155. Biochem Biophys Res Commun. 478:1130–1135. 2016. View Article : Google Scholar

16 

Tokunaga T, Naruke Y, Shigematsu S, Kohno T, Yasui K, Ma Y, Chua KJ, Katayama I, Nakamura T, Hishikawa Y, et al: Aberrant expression of interferon regulatory factor 3 in human lung cancer. Biochem Biophys Res Commun. 397:202–207. 2010. View Article : Google Scholar : PubMed/NCBI

17 

Tatsuta T, Imaizumi T, Shimoyama T, Sawaya M, Kunikazu T, Matsumiya T, Yoshida H, Satoh K and Fukuda S: Expression of melanoma differentiation associated gene 5 is increased in human gastric mucosa infected with Helicobacter pylori. J Clin Pathol. 65:839–843. 2012. View Article : Google Scholar : PubMed/NCBI

18 

Le Goffic R, Pothlichet J, Vitour D, Fujita T, Meurs E, Chignard M and Si-Tahar M: Cutting Edge: Influenza A virus activates TLR3-dependent inflammatory and RIG-I-dependent antiviral responses in human lung epithelial cells. J Immunol. 178:3368–3372. 2007. View Article : Google Scholar : PubMed/NCBI

19 

Guo R, Li Y, Ning J, Sun D, Lin L and Liu X: HnRNP A1/A2 and SF2/ASF regulate alternative splicing of interferon regulatory factor-3 and affect immunomodulatory functions in human non-small cell lung cancer cells. PLoS One. 8:e627292013. View Article : Google Scholar : PubMed/NCBI

20 

Livak KJ and Schmittgen TD: Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods. 25:402–408. 2001. View Article : Google Scholar

21 

Yoboua F, Martel A, Duval A, Mukawera E and Grandvaux N: Respiratory syncytial virus-mediated NF-kappa B p65 phosphorylation at serine 536 is dependent on RIG-I, TRAF6, and IKK beta. J Virol. 84:7267–7277. 2010. View Article : Google Scholar : PubMed/NCBI

22 

Parsons KS, Hsu AC and Wark PA: TLR3 and MDA5 signalling, although not expression, is impaired in asthmatic epithelial cells in response to rhinovirus infection. Clin Exp Allergy. 44:91–101. 2014. View Article : Google Scholar

23 

Cheng YS and Xu F: Anticancer function of polyinosinic-polycytidylic acid. Cancer Biol Ther. 10:1219–1223. 2010. View Article : Google Scholar : PubMed/NCBI

24 

Chattopadhyay S and Sen GC: dsRNA-activation of TLR3 and RLR signaling: Gene induction-dependent and independent effects. J Interferon Cytokine Res. 34:427–436. 2014. View Article : Google Scholar : PubMed/NCBI

25 

Geiss G, Jin G, Guo J, Bumgarner R, Katze MG and Sen GC: A comprehensive view of regulation of gene expression by double-stranded RNA-mediated cell signaling. J Biol Chem. 276:30178–30182. 2001. View Article : Google Scholar : PubMed/NCBI

26 

Clark K, Plater L, Peggie M and Cohen P: Use of the pharmaco-logical inhibitor BX795 to study the regulation and physiological roles of TBK1 and IkappaB kinase epsilon: A distinct upstream kinase mediates Ser-172 phosphorylation and activation. J Biol Chem. 284:14136–14146. 2009. View Article : Google Scholar :

27 

Duewell P, Beller E, Kirchleitner SV, Adunka T, Bourhis H, Siveke J, Mayr D, Kobold S, Endres S and Schnurr M: Targeted activation of melanoma differentiation-associated protein 5 (MDA5) for immunotherapy of pancreatic carcinoma. OncoImmunology. 4:e10296982015. View Article : Google Scholar : PubMed/NCBI

28 

Kübler K, tho Pesch C, Gehrke N, Riemann S, Dassler J, Coch C, Landsberg J, Wimmenauer V, Pölcher M, Rudlowski C, et al: Immunogenic cell death of human ovarian cancer cells induced by cytosolic poly(I:C) leads to myeloid cell maturation and activates NK cells. Eur J Immunol. 41:3028–3039. 2011. View Article : Google Scholar : PubMed/NCBI

29 

Bernardo AR, Cosgaya JM, Aranda A and Jiménez-Lara AM: Synergy between RA and TLR3 promotes type I IFN-dependent apoptosis through upregulation of TRAIL pathway in breast cancer cells. Cell Death Dis. 4:e4792013. View Article : Google Scholar : PubMed/NCBI

30 

Kiraz Y, Adan A, Kartal Yandim M and Baran Y: Major apoptotic mechanisms and genes involved in apoptosis. Tumour Biol. 37:8471–8486. 2016. View Article : Google Scholar : PubMed/NCBI

31 

Kirshner JR, Karpova AY, Kops M and Howley PM: Identification of TRAIL as an interferon regulatory factor 3 transcriptional target. J Virol. 79:9320–9324. 2005. View Article : Google Scholar : PubMed/NCBI

32 

Cheon H, Borden EC and Stark GR: Interferons and their stimulated genes in the tumor microenvironment. Semin Oncol. 41:156–173. 2014. View Article : Google Scholar : PubMed/NCBI

33 

Stephanou A and Latchman DS: STAT-1: A novel regulator of apoptosis. Int J Exp Pathol. 84:239–244. 2003. View Article : Google Scholar

34 

Doi T, Ishikawa T, Okayama T, Oka K, Mizushima K, Yasuda T, Sakamoto N, Katada K, Kamada K, Uchiyama K, et al: The JAK/STAT pathway is involved in the upregulation of PD-L1 expression in pancreatic cancer cell lines. Oncol Rep. 37:1545–1554. 2017. View Article : Google Scholar

35 

No authors listed: RIG-I is a tumor suppressor and biomarker of IFN-alpha efficacy in HCC. Cancer Discov. 4:OF122014. View Article : Google Scholar

36 

Coussens LM and Werb Z: Inflammation and cancer. Nature. 420:860–867. 2002. View Article : Google Scholar : PubMed/NCBI

37 

Andersen LL, Mørk N, Reinert LS, Kofod-Olsen E, Narita R, Jørgensen SE, Skipper KA, Höning K, Gad HH, Østergaard L, et al: Functional IRF3 deficiency in a patient with herpes simplex encephalitis. J Exp Med. 212:1371–1379. 2015. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

May-2018
Volume 52 Issue 5

Print ISSN: 1019-6439
Online ISSN:1791-2423

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Yi L, Sun D, Han Q, Liu Z, Zeng Z, Wu Y, Chai X and Liu X: Interferon regulatory factor 3 mediates Poly(I:C)-induced innate immune response and apoptosis in non‑small cell lung cancer. Int J Oncol 52: 1623-1632, 2018
APA
Yi, L., Sun, D., Han, Q., Liu, Z., Zeng, Z., Wu, Y. ... Liu, X. (2018). Interferon regulatory factor 3 mediates Poly(I:C)-induced innate immune response and apoptosis in non‑small cell lung cancer. International Journal of Oncology, 52, 1623-1632. https://doi.org/10.3892/ijo.2018.4300
MLA
Yi, L., Sun, D., Han, Q., Liu, Z., Zeng, Z., Wu, Y., Chai, X., Liu, X."Interferon regulatory factor 3 mediates Poly(I:C)-induced innate immune response and apoptosis in non‑small cell lung cancer". International Journal of Oncology 52.5 (2018): 1623-1632.
Chicago
Yi, L., Sun, D., Han, Q., Liu, Z., Zeng, Z., Wu, Y., Chai, X., Liu, X."Interferon regulatory factor 3 mediates Poly(I:C)-induced innate immune response and apoptosis in non‑small cell lung cancer". International Journal of Oncology 52, no. 5 (2018): 1623-1632. https://doi.org/10.3892/ijo.2018.4300