Skip to main content

RNA interferences targeting the Fanconi anemia/BRCA pathway upstream genes reverse cisplatin resistance in drug-resistant lung cancer cells

Abstract

Background

Cisplatin is one of the most commonly used chemotherapy agent for lung cancer. The therapeutic efficacy of cisplatin is limited by the development of resistance.

In this study, we test the effect of RNA interference (RNAi) targeting Fanconi anemia (FA)/BRCA pathway upstream genes on the sensitivity of cisplatin-sensitive (A549 and SK-MES-1) and -resistant (A549/DDP) lung cancer cells to cisplatin.

Result

Using small interfering RNA (siRNA), knockdown of FANCF, FANCL, or FANCD2 inhibited function of the FA/BRCA pathway in A549, A549/DDP and SK-MES-1 cells, and potentiated sensitivity of the three cells to cisplatin. The extent of proliferation inhibition induced by cisplatin after knockdown of FANCF and/or FANCL in A549/DDP cells was significantly greater than in A549 and SK-MES-1 cells, suggesting that depletion of FANCF and/or FANCL can reverse resistance of cisplatin-resistant lung cancer cells to cisplatin. Furthermore, knockdown of FANCL resulted in higher cisplatin sensitivity and dramatically elevated apoptosis rates compared with knockdown of FANCF in A549/DDP cells, indicating that FANCL play an important role in the repair of cisplatin-induced DNA damage.

Conclusion

Knockdown of FANCF, FANCL, or FANCD2 by RNAi could synergize the effect of cisplatin on suppressing cell proliferation in cisplatin-resistant lung cancer cells through inhibition of FA/BRCA pathway.

Background

Cisplatin, a DNA cross-linking agent, is one of the most commonly used drugs in the treatment of lung cancer, including non-small cell lung cancer (NSCLC) and small cell lung cancer (SCLC) [1, 2]. Although cisplatin-based chemotherapy is used as the standard treatment for patients with advanced NSCLC and SCLC, patients respond to the therapy had been more variable, and essentially all patients relapse due to acquired drug resistance [3, 4]. Thus, a great understanding of mechanisms of cisplatin resistance are essential to provide a strategy for sensitizing some drug-resistant tumor cells and improve treatment of patients with lung cancer.

Cisplatin reacts with DNA to form intrastrand cross-links, which constitutes about 85–90 % of all lesions, and interstrand cross-links (ICLs) contributing about 1–3 % to the total lesion burden [5–7]. It is ordinarily thought that the intrastrand cross-links may be the major type of DNA lesions responsible for the toxic effects of cisplatin. However, due to its severe suppressing effect on replication and transcription and the complicated character of its repair, the lower yield-forming ICLs may greatly contribute to the toxicity of cisplatin [6, 7]. There is accumulating evidence that cellular response to cisplatin exposure include the activation of multiple DNA repair pathway and enhancement of DNA damage repair process [7, 8]. It has been reported that cisplatin resistance can arise from changes that increase a cell’s capacity to repair DNA damage [9, 10]. Therefore, DNA damage repair is one of main cisplatin resistant mechanisms. The DNA repair pathway that resolve DNA ICLs, such as homologous recombination (HR) and translesion synthesis (TLS), are coordinated by a DNA damage response pathway termed the Fanconi anemia (FA)/BRCA pathway [11].

FA is a rare human genetic disease featured with severe bone marrow failure, many congenital defects, and an extremely high cancer incidence [12]. Cells from patients with FA display abnormally high sensitivity to DNA ICL-inducing agents, such as cisplatin and mitomycin. Currently there are 16 verified FA genes that make up the FA/BRCA pathway [13–17]. The FA proteins encoded by these genes are thought to work cooperatively in a common signaling pathway to repair ICLs [8, 11]. Upon DNA damage or upon entry into the S phase of the cell cycle, the core complex consists of eight-protein (FANC-A, B, C, E, F, G, L, M) act as an E3 ubiquitin ligase to monoubiquitinate FANCD2 and its paralog FANCI [18]. Monoubiquitinated FANCD2/I complex is targeted to site of chromatin damage where it recruits the downstream FA proteins (FANC-D1, N, J, O, P) and additional DNA repair proteins such as BRCA1. These downstream proteins participate in DNA repair by HR or TLC [8, 11, 17–20]. Monoubiquitination and nuclear foci formation of FANCD2 are central steps in the activation of FA/BRCA pathway [8, 11]. Cells defective in the upstream core complex are hypersensitive to certain therapeutics, particularly ICL-inducing agents. The extent of hypersensitivity depends on the specific ICL-inducing agent used [21, 22] and the defective FA protein [23].

It is realized that the integrity of the FA/BRCA pathway is required for tumor resistance to cisplatin [8, 24]. Consequently, the FA gene knockdown by RNA interference (RNAi) may sensitize tumor cells to cisplatin. Although inhibition of FANCF is known to result in FANCD2 ubiquitination inactivation and dysfunction of the FA/BRCA pathway, which enhance the sensitivity of cancer cells to DNA cross-linking agents, such as cisplatin, mitomycin and melphalan in ovarian, breast cancer, and myelomas [25–27], little is known about which gene depletion in the upstream core complex is of a stronger sensitizing effect to cisplatin in drug-resistance lung cancer cells. Therefore, we hypothesize that the knockdown of the different genes in the upstream core complex led to various degree sensitization to cisplatin in lung cancer cells. In this study, we test the impact of FANCF, FANCL and/or FANCD2 knockdown by small interfering RNA (siRNA) on the FA/BRCA pathway function and the sensitivity to cisplatin in human lung cancer cells A549 (an adenocarcinoma cell line), A549/DDP (a cisplatin-resistant A549 cell line), and SK-MES-1 (a squamous cancer cell line). Our data show that the sensitization effect to cisplatin by depleting FANCF, FANCL and FANCD2 in A549/DDP cells was significantly greater than in A549 cells. The knockdown of FANCL display more significant sensitivity to cisplatin in A549/DDP cells, as opposed to the depletion of FANCF, comparable to that of the FANCD2, suggesting that the combination of the gene therapy by depleting certain FA/BRCA pathway upstream gene and cisplatin has synergistic anti-tumor activity against cisplatin-resistant lung cancer cells.

Methods

Cell lines and culture conditions

Human lung caner cells A549 (an adenocarcinoma cell line), A549/DDP (a cisplatin-resistant A549 cell line), and SK-MES-1 (a squamous cancer cell line) were purchased from the Shanghai Institute for Biological Science (China). The cisplatin-resistant A549/DDP variant cell line was established by continuous exposure to increasing concentration of cisplatin for a 10 month period following the methodology reported previously [28]. The characteristics of A549/DDP cells have been described previously [29]. The cells were grown in RPMI-1640 complete medium supplemented with 10 % fetal calf serum (Gibco) in a humidified atmosphere containing 5 % CO2 at 37 Â°C. A549/DDP cells medium additionally contained 0.5 μg/ml cisplatin to maintain cells’ drug-resistant phenotype. Cells in the logarithmic phase of growth were used for all experiments. A549/DDP cells were cultured in complete media without cisplatin for 5 days before performing experiments.

Small interfering RNA transfection

All siRNA used in this study were purchased from Guangzhou Riboio Co., Ltd (China). The siRNA sequences used in this study are as follows: siRNAs against FANCF were siRNA-1 5′-GGUCAACGUUUGCACUAUG-3′, siRNA-2: 5′-GUACCUGGUCUUAGCAUCU-3′ and siRNA-3: 5′-CUUCGUAGUGGUGCAUUUA-3′; siRNAs against FANCL were siRNA-1 5′-GACAAGAGCTGTATGCTCT-3′, siRNA-2: 5′-CTTGCTGTGTGACTGTCAC-3′ and siRNA-3: 5′-CCAGGAAGCAACCACTTTC-3′; siRNAs against FANCD2 were siRNA-1 5′-CCTCGTAGAAGCTTATCAA-3′, siRNA-2 5′-GGTTGGCTTAATAAATTCA-3′ and siRNA-3 5′-GCTCTTTCAACGTAAATTA-3′; and non-targeting control siRNA, 5′-UAGCGACUAAACACAUCAA-3′. The indicated siRNAs were synthesized by Guangzhou Riboio Co., Ltd (China). Cell transfection was performed using Lipofectamin 2000 according to the manufacturer’s instructions. Cells were seeded in free-antibiotics media at a concentration of 1 × 105/well and cultured for 24 h before siRNA transfection. Black control, non-targeting control, siRNA-1, siRNA-2, and siRNA-3 groups were set up. Then cells were mixed with 5 Î¼/ Lipofectamin TM 2000 and 5 Î¼/ siRNA, slightly waved and then incubated in 50 % CO2 at 37 Â°C for 6 h, after which the medium was changed to complete medium and incubation was continued for 48 h before harvesting cellular proteins. Efficacy of siRNA transfection was evaluated by western blot and real-time quantitative RT–PCR. The target siRNA was chosen for further investigation as it had the highest intereference efficiency.

All experiments in this study were performed in triplicate.

Cell proliferation detection

The cell Counting-kit (CCK-8) assays were performed to measure cell proliferation. A549 and A549/DDP cells were seed at a density of 3 × 104 cells per well in 96-well plates, and allowed to grow in the growth medium for 24 h. Cells were then treated with 20 Î¼g/ml cisplatin (He pharmaceutica Co., Ltd, China) with concentrations ordinally diluted by 50 %. There was a negative control with 0 Î¼g/ml cisplatin. A blank group had no added cells. The total medium volume of each well was 100 Î¼l. Two day after drugs addition, 10 Î¼l of cell proliferation reagent CCK-8 (DOJNDO, Japan) was added into media in each well and the cell were incubated for 2 h at 37 Â°C. The absorbance of each well was determined with a spectrophotometer reading at a wavelength of 450 nm. Absorbance is assumed to be directly proportional to the number of viable cells. The IC50 was calculated as the cisplatin concentration that kills 50 % of cells of untreated control.

Real-time quantitative-PCR analysis

Total RNA was extracted using a Trizol kit (Invitrogen, CA, USA), and the concentration and purity of the recovered RNA were measured by UV-visible spectrophotometer (Bio-Tek). Then cDNA was synthesized from total RNA using MML-RT reverse transcriptase in accordance with the manufacturr’s instructions.

The real-time quantitative (RTQ)-PCR of FANCF, FANCL, and FANCD2 mRNA and β-actin as internal control were carried out on an ABI 7500 Thermal Cycler Real-time PCR system (Applied Biosystems, CA, USA), using the SYBR-Green I chemistry. Amplification primers of the three genes were synthesized by BioAsia Corporation (Shanghai, China) as follows: primer sequences for FANCF were 5′-TGCTAACAGACTGGGGTCAAC-3′ and 5′-TACAGGTCTCCAGGGCAGTTA-3′, and for FANCL, 5′-AAAGGCATTCTGGGATGTTTATG-3′ and 5′-GAAAGAAGCACTCAGGAAGCAT-3′, and for FANCD2, 5′-CCUCGUAGAAGCUUAUCAADTDT-3′ and 5′-DTDTGGAGCAUCUUCGAAUAGUU-3′ for β-actin, 5′-TGACGTGGACATCCGCAAAG-3′ and 5′-CTGGAAGGTGGACAGCGAGG-3′. The cycling conditions have been described in detail in previous report [30]. Detection of PCR products was accomplished by measuring the emitting fluorescence (Rn) at the end of each reaction step. Threshold cycle (Ct) corresponds with the cycle number required to detect a fluorescence signal above the baseline. Relative quantification was calculated with the Ct (2—△△Ct) method [31].

Western blot assay

Western blot analyses were performed to detect FANCD2 monoubiquitination and FANCF, FAMCL, FANCD2, cleaved-caspase-3, caspase-9, and poly ADP-ribose polymerase (PARP) proteins. Total cell extracts were obtained by direct lysis of cells in lysis buffer (50 mM Tris HCl, 150 mM NaCl, 10 mM β-glycerophosphate, 1 % NP-40, 1 Î¼g/ml aprotinine, 1 Î¼g/ml leupeptine, 1 Î¼g/ml pepstatine, 0.5 mg/ml pefabloc, and 1 mM Na3VO4) boiled for 5 min, and subjected to polyacrylamide SDS gel electrophoresis (Sigma Co., USA). After electrophoresis, proteins were transfer to Immobilonp-FL transfer membranes (Sigma Co., USA). The membranes were then blocked with 5 % nonfat dried milk in PBS-T overnight at 4 Â°C and then incubated at room temperature for 1 h or overnight at 4°Cin appropriate primary antibody diluted in PBS-T. After three times 5 min washes in PBS-T, membranes were then incubated for 1 h at room temperature in appropriate secondary antibody at a dilution of 1:2000 in PBS-T + 0.5 % milk. Finally, blots were washed five times for 5 min in PBS-T, and films were developed by using the ECL Western Blot Analysis System (Beijing Sage Creation Technology Co., Ltd, China). Immunoblots were probed with the following primary antibodies: FANCF, FANCL, FANCD2 (Santa Cruz and Rockland, USA), and cleaved-caspase-3, caspase-9, and PARP (Cell Signaling Technology, USA).

Immunofluorescence analysis

Immunofluorescence techniques were used to analyze FANCD2 nuclear foci formation. Cells were seed on glass coverslips and incubated overnight before or after treatment with different concentration of cisplatin for 24 h. Cells were then rinsed with PBS three times and fixed with 4 % paraformaldehyde in PBS for 15 min at 4 Â°C. The fixed cells were permeabilized with 0.3 % Triton-x-100 (Sigma, USA) in PBS for 20 min on ice. After blocking in 30 % BAS for 1 h at room temperature, anti-FANCD2 antibodies (Santa Cruz, USA) were added at dilution of 1:200. After overnight incubation at 4 Â°C, cells were washed three times with PBS and then incubated with fluorescent secondary antibodies (Sigma, USA) for 1 h at 37 Â°C. Then cells were rinsed with PBS three times. Glass coverslips were mounted in Vectashied containing DAPI. Images were captured with a Nikon fluorescence microscope and analyzed using Adobe Photoshop software (Axanoae, USA).

Flow cytometry analysis of apoptosis

The percentage of apoptosis cells induced following cisplatin treatment was determined by flow cytometry with Annexin V-FITC/PI staining. Exponential growing cells were seeded in drug-free media at a concentration of 1 × 105 cells/well overnight before experiment and then treated with different concentration of cisplatin for 48 h. After treatment, the cells were harvested, washed twice with PBS and resuspended in the provided binding buffer and subsequently staining with Annexin V-FITC (5-μg/ml) and incubated for 15 min in the dark at room temperature. Afterwards, the cells were analyzed by flow cytometry (BD Co., USA).

Statistical analysis

Statistical analyses were performed using SPSS 16.0 version. All values were expressed as means ± standard error of the mean (SEM). Comparisons between two groups were conducted by unpaired t test. Group differences resulting in P < 0.05 were considered to be statistically significant.

Results

Cisplatin resistant A549/DDP cells have enhanced function of FA/BRCA pathway

Using CCK-8 assay, we compared cisplatin sensitivity in the drug-sensitive and drug-resistant lung cancer cells. As expected, A549/DDP cells had increased proliferation rates compared with their drug-sensitive parental A549 cells 24 and 48 h after cisplatin treatment in a concentration-dependent manner (Fig. 1a). The full integrity of the FA/BRCA pathway is needed for tumor cell resistance to ICL-inducing agents and activation extent of FA/BRCA pathway depend on the level of FANCD2 monoubiquitination [11]. To determine whether the activation degree of the FA/BRCA pathway is associated with cisplatin resistance in A549/DDP cells, we tested FANCD2 monoubiquitination after treatment with various concentration of cisplatin in the two cells. The results showed that the levels of FANCD2 monoubiquitination induced by cisplatin in A549/DDP cells, as indicated by the ratio of FANCD2-L (monoubiquitinated form of FANCD2) to FANCD2-S (unubiquitinated form of FANCD2), were much higher than those in A549 cells (Fig. 1b-d). Immunofluorescent techniques were used to analyze FANCD2 nuclear foci formation, a hallmark of Fanconi anemia pathway activation. Cisplatin-induced FANCD2 foci formation, as measured by percentage of cells with greater than five foci, was found to be significantly increased in A549/DDP cells than in A549 cells (Fig. 2). These results indicate that A549/DDP cells have more marked activation of FA/BRCA pathway and stronger capability of DNA damage repair as compared with A549 cells, and also imply that the FA upstream protein participate in cisplatin resistance and upregulatory function of the FA/BRCA pathway may contribute to enhanced ICL repair capacity in A549/DDP cells.

Fig. 1
figure 1

The proliferation rates and FANCD2 monoubiquitination in A549 and A549/DDP cells treated with DDP. a The proliferation rates in A549/DDP cells 24 and 48 h after cisplatin treatment were significantly higher than those in A549 cells. b FANCD2 monubiquitination in A549 cells induced by cisplatin with different concentrations. c FANCD2 monubiquitination in A549/DDP cells induced by cisplatin with different concentrations. d The levels of FANCD2 monoubiquitination in A549/DDP cells were markedly higher than those in A549 cells after cisplatin treatment

Fig. 2
figure 2

Immunofluorescence and quantification of cells exhibiting 5 or more FANCD2 foci in lung cancer cells. Cisplatin-induced FANCD2 foci formation in A549 (a) and A549/DDP cells (b) were inhibited markedly by knockdown of FANCF, FANCL, and FANCD2. c The graph below show quantified data of FANCD2 foci in the two cells. cisplatin-induced FANCD2 foci expressions were stronger in A549/DDP than in A549 cells

Efficiency of FANCF, FANCL and FANCD2 gene knockdown by siRNA

In order to assess the role of FANCF, FANCL, and FANCD2 in the sensitivity of lung cancer cells to cisplatin, we designed three different siRNA sequences (siRNA-1, siRNA-2, siRNA-3) for each gene to knockdown the three genes respectively in A549 cells. The expression levels of both protein and mRNA of them were determined after the trasfection of different siRNAs for 48 h. As shown in Fig. 3a, b, the FANCF-siRNA-3 and FANCL-siRNA-3 significantly down-regurated both mRNA and protein expression levels of FANCF and FANCL, compared with FANCF-siRNA-1 or -siRNA-2 and FANCL-siRNA-1 or -siRNA-2. Therefore, FANCF-siRNA-3 and FANCL-siRNA-3 were selected for followed investigation. Meantime, western blot and QRT-PCR analysis showed that the interference efficiency of FANCD2-siRNA-2 was the highest among three siRNAs against FANCD2 (Fig. 3c), thus FANCD2-siRNA-2 was chosen as the target siRNA for further experiments. In addition, there was no significant difference in expression levels of the three genes between the black control group and non-targeting control group.

Fig. 3
figure 3

Expression of FANCF, FANCL and FANCD2 in A549 cells after siRNA transfection. The protein and mRNA relative expression levels of FANCF (a) and FANCL (b) were the lowest by using siRNA-3 transfection among three-different siRNA transfections. c The protein and mRNA expression inhibitions of FANCD2 were the strongest by using siRNA-2 transfection among three-different siRNA transfections. BC: black control; NC: non-target control

Knockdown of FANCF, FANCL, or FANCD2 enhance sensitivity of A549 and SK-MES-1 cells to cisplatin

Using RNAi approach, we knocked down the expression of FANCF, FANCL, and FANCD2 in A549 cells. As shown in Fig. 4a, depletion of FANCF or FANCL individually sensitized A549 cells to cisplatin. Although cells depleted of FANCL appear to be more sensitive to cisplatin than cells depleted of FANCF, difference in the IC50 between the two cells did not reach statistical significance (Table 1). Moreover, individual depletion of FANCF or FANCL from these cells and the double depletion of both FANCF and FANCL resulted in comparable hypersensitivity to cisplatin (Fig. 4a and Table 1). These data indicate that FANCF and FANCL function in a common pathway, or at least a partially overlopping pathway in the repair of DNA ICLs. The notion was demonstrated by the results of Western blot analysis, which shown that the level of FANCD2 monoubiquitination induced by cisplatin was dramatically and similarly reduced in A549 cells individually depleted of FANCF and FANCL (Fig. 4b, c). Furthermore, the double depletion of FANCF and FANCL did not further reduce FANCD2 monoubiquitination (Fig. 4b, c).

Fig. 4
figure 4

The effect of FA upstream gene knockdowns on proliferation and FANCD2 monoubiquitination in A549 cells. a Knockdowns of FANCF, FANCL, or both them inhibited the proliferation rates, and (b, c) reduced FANCD2 monoubiquitination levels after cisplatin treatment. d Knockdown of FANCD2, or double knockdown of FANCD2 and FANCF, or and FANCL suppressed the proliferation rate, and (e, f) dropped FANCD2 monoubiquitination levels after cisplatin treatment

Table 1 The 50 % inhibitory concentration (IC50) of cisplatin in A549 and A549/DDP cells 48 h after drug treatment

Meanwhile, we assessed the impact of FANCD2 status on the ability of DNA ICLs repair and the cell proliferation after treatment with cisplatin. As seen in Fig. 4d, A549 cells lacking FANCD2 were remarkably sensitive to the antiproliferative effects of this ICLs agent. Co-depletion of FANCD2 and FANCF, or and FANCL did not cause more profound sensitization to cisplatin (Fig. 4d, Table 1). Given that these gene knockdowns were highly effective, the results confirm that FANCF and FANCL are epistatic of FANCD2 and are involved in the same pathway to confer tolerance of cisplatin-induced DNA damage. Consistently, co-depletion of both FANCD2 and FANCF, or FANCL produced no additional decrease of FANCD2 monoubiquitination levels upon DNA damage induced by cisplatin, when compared with depletion of FANCD2 alone (Fig. 4e, f). Additionally, cisplatin-induced FANCD2 foci formation was inhibited by depleting FANCF, FANCL, or FANCD2 in similar degree (Fig. 2a, c), which were identical with results of FANCD2 monoubiquitination analyses.

SK-MES-1 cell is a lung squamous cancer cell which display slight resistance to cisplatin (IC = 9.92 ± 0.35, Table 1) compared with A549 cell (IC = 4.97, Table 1). We examined the change of sensitivity in SK-MES-1 cells by knocking down FANCF, FANCL, and FANCD2 using siRNA transfection, and observed FANCD2 monoubiquitination levels. The results showed that individual and double knockdown of FANCF, or FANCL, or FANCD2 markedly enhanced the sensitivity of SK-MES-1 cells to cisplatin, and reduced the levels of FANCD2 monoubiquitination (Fig. 5a-f), which are in line with the findings in A549 cells.

Fig. 5
figure 5

The effect of FA upstream gene knockdowns on proliferation and FANCD2 monoubiquitination in SK-MES-1 cells. a Knockdowns of FANCF, FANCL, or both them inhibited the proliferation rates, and (b, c) reduced FANCD2 monoubiquitination levels after cisplatin treatment. (d) Knockdown of FANCD2, or double knockdown of FANCD2 and FANCF, or and FANCL suppressed the proliferation rate, and (e, f) dropped FANCD2 monoubiquitination levels after cisplatin treatment

Knockdown of FANCL results in higher cisplatin sensitivity compared with knockdown of FANCF in A549/DDP cells

The above-mentioned experimental results have shown that FA/BRCA pathway involved the sensitivity to cisplatin in drug-resistant A549/DDP cells. We next explored relative roles of FANCF and FANCL in the sensitivity to cisplatin by depleting FANCF only, FANCL only, or both the two core complex components, and testing cisplatin-induced cell proliferation in A549/DDP cells. As shown in Fig. 6a, siRNA mediated depletion of FANCF or FANCL significantly sensitized A549/DDP cells to cisplatin. It is worth noting that cells depleted of FANCL were more sensitive to cisplatin than those depleted of FANCF. As seen in Table 1, the IC50 value in FANCF depleted cells was significantly higher than that in FANCL depleted cells (P < 0.001). But co-depletion of FANCF and FANCL did not produce any additional sensitivity over that of FANCL. The differential sensitization suggests that FANCL has an epistatic yet more important role than FANCF in cellular resistance to cisplatin. Moreover, western blot and immunofluorescence assay revealed that individual or double depletion of FANCF and FANCL resulted in significant decrease of FANCD2 monoubiquitination levels (Fig. 6b, c) and inhibition of FANCD2 foci formation (Fig. 2b, c).

Fig. 6
figure 6

The effect of FA upstream gene knockdowns on proliferation and FANCD2 monoubiquitination in A549/DDP cells. a Knockdown of FANCF, FANCL, or both them markedly inhibited the proliferation rates, and (b, c) reduced FANCD2 monoubiquitination levels after treatment with cisplatin. (d) Knockdown of FANCD2, or double knockdown of FANCD2 and FANCF, or and FANCL significantly suppressed the proliferation rates, and (e, f) reduced FANCD2 monoubiquitination levels after treatment with cisplatin

Additionally, we found that reduced extent of proliferation rate induced by cisplatin treatment in A549/DDP cells was significantly greater than in A549 cells after knockdown of FANCF or FANCL (Fig. 4a and 6a, Table 1), suggesting that knockdown of FANCF or FANCL can reverse resistance of cisplatin-resistant lung cancer cells to cisplatin, whereas the knockdown of the two genes in cisplatin-sensitive lung cancer cells only moderately increase sensitivity to cisplatin.

Knockdown of FANCD2 lead to a similar cell sensitivity to that of FANCL in A549/DDP cells

We also evaluated the sensitizing effect of FACD2 knockdown on cisplatin in A549/DDP cells. As shown in Fig. 6d, depletion of FANCD2 remarkably reduced cell proliferation rate after cisplatin treatment. Nevertheless, depletion of FANCD2 alone, or co-depletion of FANCD2 and FANCF, or and FANCL exhibited a similar reduction in cell proliferation following cisplatin treatment. Consistent with the sensitization effects of these gene knockdowns on cisplatin, cells depleted of FANCD2 alone, and cells co-depleted of both FANCD2 and FANCF, or FANCL, had nearly identical decrease of FANCD2 monoubiquitination (Fig. 6e, f). Moreover, the IC50 value of FANCD2 knockdown cells was similar to that of FANCL depleted cells, but significantly lower than that of FANCF depleted cells (P < 0.001) (Table 1). Taken together, these results further support the idea that FANCF and FANCL are epistatic to FANCD2, and FANCF and FANCL function in a common pathway, but FANCL may play a more important role in facilitating FANCD2 monoubiquitination.

Knockdown of FA upstream proteins promotes apoptosis of A549/DDP cells

To explore whether effect of proliferative inhibition in cisplatin-resistant lung cancer cells by depleting FANCF or FANCL was associated with apoptosis, we detected the apoptosis rate of the A549/DDP cells induced by cisplatin following depletion of the two genes. As shown in Fig. 7, compared with the control, depletion of FANCF or FANCL in A549/DDP cells significantly increased the percentage of both early and late apoptosis cells after cisplatin treatment in a concentration-dependent manner (Fig. 7a-d, P = 0.001 and P < 0.001). Furthermore, the percentages of apoptosis cells in FANCL depleted cells were markedly higher than those in FANCF depleted cells, when treatment with 10 Î¼g/ml and 20 Î¼g/ml cisplatin (Fig. 7d, P = 0.001 and P < 0.001), indicating that knockdown of FANCF or FANCL by siRNA could synergize the effect of cisplatin on inducing apoptosis of cisplatin-resistant lung cancer cells through suppressing FA/BRCA pathway. To investigate if the apoptosis caused by depleting FANCF or FANCL is caspase-dependent, we analyzed caspase-dependent apoptpsis-associated proteins by western blot. The results showed that depletion of FANCF and FANCL clearly increased the levels of cleaved caspase-3, caspase-9, and PARP proteins in A549/DDP cells (Fig. 8a, b), indicating that knockdown of FANCF and FANCL activate caspase-dependent apoptosis.

Fig. 7
figure 7

Knockdown of FANCF and FANCL promote the apoptosis induced by cisplatin in A549/DDP cells. a The apoptosis rates of control cells, and (b) cells transfected with FANCF-siRNA, and (c) cells transfected with FANCL-siRNA after cisplatin treatment. Early apoptotic cells are in the lower right quadrant; late apoptotic cells are in the upper right quadrant. d The apoptosis rates in cells transfected with FANCF-siRNA or FANCL-siRNA were significantly higher than in control cells in a concentration-dependent manner. In addition, the apoptosis rates of cells transfected with FANCL-siRNA have dramatic increase as compared with cells transfected with FANCF-siRNA following treatment of 10 Î¼g/ml and 20 Î¼g/ml cisplatin

Fig. 8
figure 8

Caspase-3, caspase-9, and PARP expressions were detected by Western blot in A549/DDP cells. a The expressions of cleaved caspase-3, caspase-9, and PARP induced by cisplatin were elevated after siFANCF transfection. b The expressions of cleaved caspase-3, caspase-9, and PARP induced by cisplatin were increased after siFANCL transfection

Discussion

Increasing evidences indicate that cancer cell resistance to cross-links agents, such as cisplatin, is linked to the functional status of the FA/BRCA pathway [8, 11, 25–27], albeit the exact mechanism of this resistance to the chemical crosslinker remains to be fully elucidated. In this study, we confirm that upregulation of FANCD2 monoubiquitination and activation of FA/BACA pathway contributes to acquired resistance to cisplatin for lung cancer A549/DDP cells. Moreover, we show that knockdown of FANCF and/or FANCL by siRNA potentiates sensitivity to cisplatin in cisplatin-sensitive and -resistant lung cancer cells.

Within the FA/BRCA pathway, eight upstream FA protein and some FA-associated proteins form a ligase core complex that monoubiquitinates both components of the ID complex, FANCD2 and FANCI, in response to DNA damage and replication stress during S phase. The FA core complex mainly comprises three subunits, including FANCB-FANCL-FAAP100 (B-L-100), FANCC-FANCE-FANCF (C-E-F) and FANCA-FANCG-FAAP20 (A-G-20) [32–37]. It is generally thought that a single mutation in any of the FA upstream genes that make up the core complex prevents the FANCD2 monoubiquitination event from occurring. FANCF had been implicated in the physical stability of the majority of other core complex proteins, FANCC, FANCE, FANCG and FANCA by several groups and functions as a adaptor that is essential for the integrity of the core complex and the monoubiquitination of FANCD2 [34, 38]. FANCL is considered a key component of the FA core complex and represent the E3 ligase necessary for ubiquitin conjugation to lysine 561 of FANCD2 and lysine 523 of FANCI [39, 40]. The depletion of FANCF gene results in the inhibition of FANCD2 monoubiquitination and the enhancement of sensitivity to DNA cross-linking agents (melphalan, cisplatin and mitomycin) in myeloma, ovarian and breast cancer cells [25–27]. In line with these reports, we results show that knockdown of FANCF sensitized drug-sensitive and -resistant lung cancer cells to cisplatin and diminished the levels of FANCD2 monoubiquitination. Similarly, knockdown of FANCL increase sensitivity to cisplatin in the lung cancer cells. However, the sensitization effect to cisplatin by depleting FANCF or FANCL in A549/DDP cells was more significant as compared with A549 and SK-MES-1 cells. For instance, the IC50 values of A549/DDP cells transfected with FANCL-siRNA were reduced from 34.15 Î¼g/ml pre-transfection to 2.51 Î¼g/ml post-transfection, whereas the IC50 values of A549 cells pre- and post-transfection were 4.97 Î¼g/ml and 2.46 Î¼g/ml, respectively (Table 1). Also, the dropped extent of FANCD2 monoubiquitination levels in A549/DDP cells was greater than in A549 cells following depletion of FANCF and/or FANCL. The results suggest that enhanced E3 ligase activity of the core complex in A549/DDP cells might be responsible for cisplatin resistance, and the suppression of the core complex ligase activity and downregulation of FANCD2 monoubiquitnation by siRNA targeting FANCF or FANCL can reverse resistance of cisplatin-resistant lung cancer cells to cisplatin.

In this study, another interesting finding is that knockdown of FANCL led to more dramatic sensitization effect on cisplatin compared with knockdown of FANCF in A549/DDP cells. The IC50 value of depleted FANCF A549/DDP cells was markedly higher than that of depleted-FANCL cells 48 h after cisplatin treatment. Moreover, apoptosis rates of A549/DDP cells depleted of FANCL were significantly elevated following cisplatin treatment, not only compared with those of control, but also compared with those depleted of FANCF. But co-depletion of FANCF and FANCL did not have a significantly additive impact on sensitivity to cisplatin. This phenomenon was different from A549 cells, in which individual knockdown of FANCF and FANCL, or in combination, resulted in comparable cisplatin sensitivity and levels of FANCD2 monoubiquitination. These finding suggest that FANCL may play a notably important role for acquired resistance to cisplatin in A549/DDP cells, although FANCF and FANCL were epistatic and act in the common FA/BRCA pathway.

A recent study shown that not all the components of the core complex contribute equally to cellular resistance against DNA damage [23], even though the study results deviates from a general idea that losing any one of the core comlex components results in complete elimination of FANCD2 activation and disintegration of the core complex. The study reported that the subunits in core complex are of different function and the overall integrity of the core complex is sustained when certain FA proteins are removed. Among these subunits, FANCB, L and FAAP100 constitute a catalytic subunit that is absolutely required for the E3 ligase activity of the core complex, whereas the A-G-20 and C-E-F subunits provide non-redundant and ancillary functions that help the catalytic subunit bind chromatin or site of DNA damage [23]. Also, other recent studies suggest a key role of FANCL as the E3 ligase in the monoubiquitination of FANCD2 and activation of FA/BRCA pathway. For instance, DT40 cells (a chicken somatic cell) lacking FANCL, B, or FAAP100 completely lost the ability to monoubiquitinate FANCD2, implying that subunit is indispensable for FANCD2 monubiquitination. However, FANCD2 monoubiquitination activity in cells depleted of FANCC and FANCF was comparable to the wild–type cells [37]. Furthermore, the study found that in comparison to wild-type cells, residual chromatin-associated FANCD2 monoubiquitination was still present in the cells lacking FANCA, FANCG, FANCF, or FANCC [37]. Therefore, loss of different FA core components produces variable impact on the activation of the FA/BRCA pathway and correspondingly variable cellular sensitivity to cross-linking agents. Perhaps these findings may explain our results that depletion of FANCL had a more potent sensitization on A549/DDP cells than knockdown of FANCF. However, it is unclear why there was no significant difference for the sensitivity of A549 cells to cisplatin between knockdown of FANCF and FANCL. But it is known that cisplatin-induced DNA cross-links might be repaired by FANCF and FANCL which are components of FA core complex. FANCF might play a minor role in the repair pathway for cisplatin-induced damage, whereas FANCL play a crucial role as mentioned above. The difference for repair capacity of DNA damage between FANCF and FANCL might be more significant in cisplatin-resistant lung cancer cells. Further study is required to ascertain the precise mechanism that DNA repair activities of FA/BRCA pathway in cisplatin-resistant cells differ from cisplatin-sensitive cells.

It is known that monoubiquitinated FANCD2-I (ID) complex is at the heart of the FA/BRCA pathway. The monoubiquitination of FANCD2 at lysine 561 is essential for the downstream function of the pathway in DNA damage repair [41]. Cells expressing k561R FANCD2, which is a mutant form of FANCD2 and cannot be monoubiquitinated, have increased sensitivity to cross-linking agent compared with wild-type FANCD2 [41]. Consistent with previous reports using other cells [42, 43], siRNA-mediated depletion of FANCD2 potently sensitized A549, A549/DDP and SK-MES-1 cells to cisplatin, which is similar to the sensitization of the three cells to cisplatin by depleting FANCL. As expected, we did not observe further sensitization of the three cells to cisplatin upon treatment with double depletion of FANCD2 and FANCF, or FANCL, further supporting the notion that FANCD2, FANCF and FANCL act in a common DNA repair pathway.

Cisplatin-based combination chemotherapy is the first-line therapy for SCLC and advanced NSCLC. Despite considerable success in treatment of lung cancer, the effectiveness of cisplatin in clinics has been limited by the resistance that cancer cells developed during the course of chemotherapy. Thus, overcoming tumor cells resistance to cisplatin using RNAi targeting the FA upstream genes in combination with cisplatin is a promising strategy for treating lung cancer. With recent advance in gene therapy techniques, improvements of in vivo RNAi delivery methodologies indicate that adjuvant siRNA therapy may be achievable in accessible tumor sites [44]. Further investigations are warranted to determine whether the gene therapy by targeting the FA/BRCA pathway combined with chemotherapy can used as an effective novel therapeutic regimen for patients with lung cancer, particular those with tumor relapse or progression after chemotherapy.

Conclusion

In summary, we herein report that depletion of FANCF, FANCL, or FANCD2 by using RNAi could reverse cisplatin resistance in cisplatin-resistant lung cancer cells through inhibition of FA/BRCA pathway. Moreovre, knockdown of FANCL resulted in higher cisplatin sensitivity and dramatically elevated apoptosis rates compared with knockdown of FANCF, indicating that FANCL play an important role in the repair of cisplatin-induced DNA damage. Targeting of the FA/BRCA pathway may constitute a potential treatment modality for lung cancer.

References

  1. Boulikas T, Vougiouka M. Recent clinical trials using cisplatin, carboplatin and their combination chemotherapy drugs. Oncol Rep. 2004;11:559–95.

    CAS  PubMed  Google Scholar 

  2. Chua YJ, Steer C, Yip D. Recent advances in management of small-cell lung cancer. Cancer Treat Rev. 2004;30:521–43.

    Article  CAS  PubMed  Google Scholar 

  3. Schiller JH, Harington D, Balani CP, Langer C, Sandler A, Krook J, et al. Comparison of four chemotherapy regimens for advanced non-small cell lung cancer. N Engl J Med. 2002;346:92–8.

  4. Li J, Dai C-H, Chen P, Wu J-N, Bao Q-L, Qiu H, et al. Survival and prognostic factors in small cell lung cancer. Med Oncol. 2010;27:73–81.

  5. Jung Y, Lipapard SJ. Direct cellular response to platinum-induced DNA damage. Chem Rev. 2007;107:1387–407.

    Article  CAS  PubMed  Google Scholar 

  6. Wang D, Lippard SJ. Cellular processing of platinum anticancer drugs. Nat Rev Drug Discov. 2005;4:307–20.

    Article  CAS  PubMed  Google Scholar 

  7. Ahmad S. Platinum-DNA interactions and subsequent cellular processes controlling sensitivity to anticancer platinum complexes. Chem Biodivers. 2010;7:543–66.

    Article  CAS  PubMed  Google Scholar 

  8. Kim H, D’Andrea AD. Regulation of DNA cross-link repair by the fanconi anemia/BRCA pathway. Genes Develop. 2012;26:1393–408.

  9. Li J, Li Z-N, Ou Y-J, Li X-Q, Bao Q-L, Chen P. Expression of MRP1, BCRP, LRP, and ERCC1 in advanced non-small-cell lung cancer: correlation with response to chemotherapy and survival. Clin Lung Cancer. 2009;6:414–21.

    Article  Google Scholar 

  10. Wang H, Zhang S-Y, Wang S, Lu J, Wu W, Weng L, et al. REV3 L confers chemoresistance to cisplatin in human gliomas: The potential of its RNAi for synergistic therapy. Neuro-Oncol. 2009;11:790–802.

  11. Kee Y, D’Andrea AD. Expanded roles of the Fanconi anemia pathway in preserving genomic stability. Genes Dev. 2010;24:1680–94.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  12. Wang W. Emergence of a DNA-damage response network consisting of Fanconi anemia and BRCA proteins. Nat Rev Genet. 2007;8:735–48.

    Article  CAS  PubMed  Google Scholar 

  13. Levran O, Attwooll C, Henry RT, Miton KL, Neveling K, Rio P, et al. The BRCA-interacting helicase BRIP1 is deficient in Fanconi anemia. Nature Genet. 2005;37:931–3.

  14. Smogorzewska A, Matsuoka S, Vinciguerra P, Mc Donald ER, Hurov KE, Luo J, et al. Identification of the FANCI protein, a monoubiquitinated FANCD2 paralog required for DNA repair. Cell. 2007;129:289–301.

  15. Kim Y, Lach FP, Desetty R, Hanenbery H, Auerbach AD, Smogorzewska A. Mutations of the SLX4 gene in Fanconi anemia. Nature Genet. 2011;43:142–6.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  16. Vaz F, Hanenberg H, Schuster B, Barker K, Wiek C, Erven V, et al. Mutation of the RAD51C gene in a Fanconi anemia-like disorder. Nature Genet. 2010;42:406–9.

  17. Bogliolo M, Schuster B, Stoepker C, Derkunt B, Su Y, Raams A, et al. Mutations in ERCC4, encoding the DNA-repair endonuclease XPF, cause Fanconi anemia. Am J Hum Genet. 2013;92:800–6.

  18. Garner E, Smogorzewska A. Ubiquitylation and the Fanconi anemia pathway. FEBS Lett. 2011;585:2853–60.

    Article  CAS  PubMed  Google Scholar 

  19. Knipscheer P, Raschle M, Smogorzewska A, Enoiu M, Ho TV, Scharer OD, et al. The Fanconi anemia pathway promotes replication-dependent DNA interstrand cross-link repair. Science. 2009;326:1698–701.

  20. Kottemann MC, Smogorzewska A. Fanconi anemia and the repair of Watson and Crick DNA crosslinks. Nature. 2013;493:356–63.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  21. Gallmeier E, Calhoun ES, Rago C, Brody JR, Cunninghan SC, Hucl T, et al. Targeted disruption of FANCC and FANCG in human cancer provides a preclinical model for specific therapeutic options. Gastroenterology. 2006;130:2145–54.

  22. Van der Heijden MS, Brody JR, Gallmeier E, Cunningham SC, Dezentje DA, Shen D, et al. Functional defects in the fanconi anemia pathway in pancreatic cancer cells. Am J Pathol. 2004;165:651–7.

  23. Huang Y, Leung JWC, Lowery M, Matsushita N, Wang Y, Shen X, et al. Modulatized functions of the Fanconi anemia core complex. Cell Rep. 2014;7:1–19.

  24. Taniguchi T, D’Andrea AD. Molecular pathogenesis of Fanconi anemia: recent progress. Blood. 2006;107:4223–33.

    Article  CAS  PubMed  Google Scholar 

  25. Chirnomas D, Taniguchi T, de la Vega M, Vaidya AP, Vasserman M, Hartman A-R, et al. Chemosensitization to cisplatin by inhibitors of the Fanconi anemia/BRCA pathway. Mol Cancer Ther. 2006;5:952–6.

  26. Yu J, Zhao L, Li Y, Li N, He M, Bai X, et al. Silencing of Fanconi anemia complementation group F exhibits potent chemosenstization of mitomycin C activity in breast cancer cells. J Breast Cancer. 2013;16:291–9.

  27. Chen Q, Van der Sluis PC, Boulware D, Hazlehurst LA, Dalton WS. The FA/BRCA pathway is involved in melphalan-induced DNA interstrand cross-link repair and accounts for melphalan resistance in multiple myeloma cells. Blood. 2005;106:698–705.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  28. Plasencia C, Martinez-Balibrea E, Martinez-Cardus A, Quinn DI, Abad A, Neamati N. Expression analysis of genes involved in oxaliplatin response and development of oxaliplatin-resistant HT 29 colon cancer cells. Int J Oncol. 2006;29:225–35.

    CAS  PubMed  Google Scholar 

  29. Huanq Z, Tong Y, Wang J, Huang Y. NMR studies of the relationship between the changes of membrane lipids and the cisplatin-resistance ofA549/DDP cells. Cancer Cell Int. 2003;8:5.

    Article  Google Scholar 

  30. Tang X-P, Li J, Yu L-C, Chen Y-C, Shi S-B, Zhu L-R, et al. Clinical significance of survivin and VEGF mRNA detection in the cell fraction of the peripheral blood in non-small cell lung cancer patients before and after surgery. Lung Cancer. 2013;81:273–9.

  31. Livak KJ, Schnittgen TD. Analysis of relative gene expression data using real-time quantitative PCR and the 2-△△Ct method. Methods. 2001;25:402–8.

    Article  CAS  PubMed  Google Scholar 

  32. Medhurst AL, Laghmani H, Steltenpool J, Ferrer M, Fontaine C, de Groot J, et al. Evidence for subcomplexes in the Fanconi anemia pathway. Blood. 2006;108:2072–80.

  33. Ling C, Ishiai M, Ali AM, Medhurst AL, Neveling K, Kalb R, et al. FAAP100 is essential for activation of the Fanconi anemia-associated DNA damage response pathway. EMBO J. 2007;26:2104–14.

  34. Leveille F, Blom E, Medhurst AL, Bier P, Lanhmani H, Johnson M, et al. The Fanconi anemia gene product FANCF is a flexible adaptor protein. J Biol Chem. 2004;279:39421–30.

  35. Ali AM, Pradhan A, Singh TR, Du C, Li J, Wahengbam K, et al. FAAP 20: a novel ubiquitin-binding FA nuclear core-complex protein required for functional integrity of the FA-BRCA DNA repair pathway. Blood. 2012;119:3285–94.

  36. Reuter T, Herterich S, Bernhard Q, Hoehn H, Gross HJ. Strong FANCA/FANCG but week FANCA/FANCC interaction in the yeast z-hybrid system. Blood. 2000;95:719–20.

    CAS  PubMed  Google Scholar 

  37. Rajendra E, Oestergaard VH, Langevin F, Wang M, Dornan GL, Patel KJ, et al. The genetic and biochemical basis of FANCD2 monoubiquitination. Mol Cell. 2014;54:858–69.

  38. Kowl P, Gurtan AM, Stuckert P, D’Andrea AD, Ellenberger T. Structural determinants of human FANCF protein that function in the assembly of a DNA damage signaling complex. J Biol Chem. 2007;282:2047–55.

    Article  Google Scholar 

  39. Cole AR, Lewis LP, Walden H. The structure of the catalytic subunit FANCL of the Fanconi anemia core complex. Nat Struct Mol Biol. 2012;17:294–8.

    Article  Google Scholar 

  40. Alpl AF, Pace PE, Babu MM, Patel KJ. Mechanistic insight into site-restricted monoubiguitination of FANCD2 by Ubezt, FANCL, and FANCI. Mol Cell. 2008;32:767–77.

    Article  Google Scholar 

  41. Gatcia-Higuera I, Taniguchi T, Ganesan S, Meyn MS, Timmers C, Hejna J, et al. Interaction of Fanconi anemia proteins and BRCA1 in a common pathway. Mol cell. 2001;7:249–62.

  42. Palle K, Vaziri C. Rad 18 E3 ubiquitin ligase activity mediates Fanconi anemia pathway activation and cell survival following DNA topoisomerase I inhibition. Cell Cycle. 2011;10:1625–38.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  43. Song IY, Ralle K, Gurkar A, Tateishi S, Kupfer GM, Vaziri C. Rad 18-mediated translesion sythesis of bulky DNA adducts in coupled to activation of the Fanconi anemia DNA repair pathway. J Biol Chem. 2010;285:31525–36.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  44. Wullner U, Neef I, Tur M, Bath S. Targeted delivery of short interfering RNAs-strategies for in vivo delivery. Recent Pat Anticancer Drug Discov. 2009;4:1–8.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This research was supported by Center of Experimental Medicine, Affiliated Hospital of Jiangsu University.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jian Li.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

C-HD and JL contributed to the conception of design of the study, analysis of data, and wrote the manuscript. PC, H-GJ, and MW performed the experiments and the statistical analysis. Y-CC participated in the design of the study, analysis of data, and revised the manuscript. All authors read and approved the final manuscript.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dai, CH., Li, J., Chen, P. et al. RNA interferences targeting the Fanconi anemia/BRCA pathway upstream genes reverse cisplatin resistance in drug-resistant lung cancer cells. J Biomed Sci 22, 77 (2015). https://doi.org/10.1186/s12929-015-0185-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12929-015-0185-4

Keywords