Geranylgeranyl transferase 1 inhibitor GGTI‑298 enhances the anticancer effect of gefitinib

  • Authors:
    • Bi‑Sheng Liu
    • Xin‑Yu Dai
    • Hong‑Wei Xia
    • Huan‑Ji Xu
    • Qiu‑Lin Tang
    • Qi‑Yong Gong
    • Yong‑Zhan Nie
    • Feng Bi
  • View Affiliations

  • Published online on: August 9, 2018     https://doi.org/10.3892/mmr.2018.9371
  • Pages: 4023-4029
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Dysregulation of epidermal growth factor receptor (EGFR) signaling is responsible for the resistance to EGFR tyrosine kinase inhibitors (TKIs), such as gefitinib and erlotinib, and is thereby associated with the progression of tumors in non‑small cell lung cancers (NSCLCs). Immunoblotting results revealed that geranylgeranyl transferase 1 inhibitor (GGTI)‑298, a geranylgeranyl transferase 1 inhibitor with potential antitumor effects, effectively inhibited the phosphorylation of EGFR and its downstream target protein kinase B (AKT). A combination of gefitinib and GGTI‑298 amplified the inhibition of the EGFR‑AKT signaling pathway. In addition, GGTI‑298 treatment produced a synergistic effect on the inhibition of proliferation as indicated by the combination index values of <1 when combined with gefitinib in the NSCLC cell lines HCC827 and A549. These synergistic effects were also observed to induce apoptosis and migration inhibition. Further mechanistic studies demonstrated that GGTI‑298 inhibited the activity of Ras homolog family member A (RhoA), and downregulation of RhoA with small interfering RNA impaired the phosphorylation of EGFR, which suggested that EGFR inhibition by GGTI‑298 may be exerted mainly through RhoA mediation. These results presented a novel, promising therapeutic strategy involving a combination of two drugs for targeting EGFR signaling in lung cancer.

Introduction

Tyrosine kinase inhibitors (TKIs), such as gefitinib and erlotinib, target epidermal growth factor receptors (EGFRs) and have been demonstrated to be efficacious in treating non-small cell lung cancers (NSCLCs) (13). The dysregulation of EGFR signaling is closely associated with cell proliferation and EGFR TKI resistance (4,5), which highlights the importance of complete EGFR signaling inhibition in lung cancer treatment.

Geranylgeranylation is an important biochemical modification for cellular localization and the function of small G proteins (6,7). Previous studies have reported that geranylgeranylation is critical for cancer cell proliferation and invasion (8,9), which indicates the potential anticancer function of inhibiting geranylgeranylation. The geranylgeranyl transferase 1 (GGT1) inhibitor (GGTI)-298 inhibits protein geranylgeranylation. Recent studies in vitro and in animal models have revealed that GGTI-298 inhibited cell proliferation and induced apoptosis when applied as a single agent or in combination with chemotherapeutics (1012). In addition, GGTIs inhibited EGF-mediated EGFR tyrosine phosphorylation, indicating that the GGT1 substrate may serve a vital role in the process of EGFR phosphorylation (13). However, the molecular mechanism underlying the GGTI-induced inhibition of EGFR phosphorylation has not been well studied.

Ras homolog family member A (RhoA), a widely studied protein in the Rho family of guanosine triphosphate (GTP)-ases, participates in various physiological processes, including cytoskeleton organization, cell cycle progression, cell survival and cell migration (14). A growing number of studies have revealed that RhoA interacts with various signaling molecules to exert effects on cell proliferation and apoptosis (1517). Thus, a deeper understanding of RhoA-associated signaling regulation will support the identification of novel strategies and modalities in cancer treatment.

In the present study, a potentially favorable interaction between GGTI-298 and the EGFR TKI gefitinib was identified. The two combined stimulated cytotoxicity and induced apoptosis in lung cancer cells. The synergistic mechanism was also studied.

Materials and methods

Cell culture and agents

The NSCLC cell lines (HCC827, A549, H1975, SPCA1 and H1299) were purchased from American Type Culture Collection (Manassas, VA, USA). All cell lines were cultured in Dulbecco's modified Eagle's medium (Gibco; Thermo Fisher Scientific, Inc., Waltham, MA, USA) supplemented with or without 10% heat inactivated fetal bovine serum (Gibco; Thermo Fisher Scientific, Inc.) depending on the assay. Cell cultures were maintained at 37°C in a humidified (5% CO2, 95% air) incubator. GGTI-298 and gefitinib were purchased from Selleck Chemicals (Houston, TX, USA), and dimethylsulfoxide (DMSO) and EGF were purchased from Sigma-Aldrich; Merck KGaA (Darmstadt, Germany).

Immunoblotting analysis

A549 cells were treated with 5 µM GGTI-298 in serum-free medium for 48 h. Prior to protein extraction, the cultures were exposed to 30 ng EGF for 30 min. Cells (HCC827, A549, H1975, SPCA1 and H1299) were lysed with ice-cold Radioimmunoprecipitation Assay buffer [150 mM NaCl, 1% NP-40, 50 mM Tris-HCl (pH 7.4), 1 mM phenylmethylsulfonyl fluoride, 1 µg/ml leupeptin, 1 mM deoxycholic acid and 1 mM EDTA] containing a protease inhibitor cocktail (Sigma-Aldrich; Merck KGaA). The protein concentration was detected using the BCA assay (Thermo Fisher Scientific, Inc., Waltham, MA, USA). In total, ~20 µg total protein was loaded on to a 12% SDS-PAGE gel, separated electrophoretically and then transferred to nitrocellulose membranes (EMD Millipore, Billerica, MA, USA). The membranes were blocked with 5% nonfat milk at room temperature for 30 min and were washed three times with phosphate buffer saline (PBS). The membranes were incubated with the following primary antibodies: Phosphorylated (p)-EGFR (cat. no. 3777), cyclin D1 (cat. no. 2922), phosphatase and tensin homolog (PTEN; cat. no. 9552), GAPDH (cat. no. 97166), glutathione S-transferase (GST; all Cell Signaling Technology, MA, USA), p-extracellular signal-regulated kinase (ERK; cat. no. EP197Y), ERK (cat. no. ab 17942), p-protein kinase B (AKT; cat. no. EP2109Y), AKT (cat. no. EPR16798; all Abcam, Cambridge, UK), EGFR (cat. no. sc-03-G), 3-hydroxy-3-methylglutaryl-co-enzyme A reductase (HMGCR; cat. no. sc-271595) and RhoA (cat. no. sc-418; all Santa Cruz Biotechnology, Inc., Dallas, TX, USA). All of the primary antibodies were added at a 1:1,000 dilution overnight at 4°C. Subsequently, the membranes were incubated with the horseradish peroxidase conjugated anti-mouse (cat. no. HA1022) or anti-rabbit (cat. no. HA1001) secondary antibodies with dilution of 1:5,000 (Huabio, Hangzhou, China) at room temperature for 3 h and were washed with PBS. Protein expression was detected using the ODYSSEY Infrared Imaging System (LI-COR Biosciences, Lincoln, NE, USA) and quantified using Image J 1.48 (National institute of Health, Bethesda, MA, USA).

Cell proliferation assay

HCC827 and A549 cells were seeded (~5×103 cells per well in 100 µl DMEM with 10% FBS) in a 96-well flat-bottomed plate (Corning Incorporated, Corning, NY, USA) at 24 h prior to treatment, namely, HCC827 and A549 cells were treated with gefitinib (HCC827, 10 nM; A549, 5 µM) and/or GGTI-298 (concentration indicated) for 48 h at 37°C in a humidified (5% CO2, 95% air) incubator. Following incubation at 37°C, cell growth was measured using a Cell Counting Kit-8 (Dojindo Molecular Technologies, Inc., Kumamoto, Japan). Synergistic activity between gefitinib and GGTI-298 was assessed using combination index (CI) analysis, where a CI <1 indicated synergism, CI=1 indicated additive effects and CI >1 indicated antagonism (18).

Cell apoptosis assay

HCC827 and A549 cells were seeded (~5×103 cells per well in 100 µl medium) in a 6-well flat-bottomed plate 24 h prior to treatment, namely, HCC827 and A549 cells were treated with DMSO, gefitinib (HCC827, 10 nM; A549, 5 µM), GGTI-298 (HCC827, 5 µM; A549, 7.5 µM) or a combination of gefitinib and GGTI-298 for 48 h at 37°C in a humidified (5% CO2, 95% air) incubator. Apoptotic cells were detected using Hoechst 33342 (Sigma-Aldrich; Merck KGaA) DNA staining (room temperature for 10 min), according to the kit instructions. Apoptotic cells were characterized by nuclear condensation and DNA fragmentation, which were observed by fluorescence microscopy (Nikon Corporation, Tokyo, Japan).

Wound healing assay

HCC827 and A549 cells (~5×105) were seeded in a 6-well culture plate for 24 h at 37°C. Wounding was performed by scraping through the cell monolayer with a 200 µl pipette tip. Medium and nonadherent cells were removed and replaced with 2 ml of fresh serum-free medium containing the indicated reagents (DMSO, gefitinib, GGTI-298 or a combination of gefitinib and GGTI-298). Cells were permitted to migrate for 48 h at 37°C. Wound healing was photographed microscopically (light) at 0, 24 and 48 h. Three fields of view for each wound were randomly selected, and the average wound width was calculated. The relative wound width was calculated as follows: wound width at the indicated time (0, 24 or 48 h)/wound width at 0 h.

Pull-down assays

Briefly, to detect the levels of active RhoA, equal amounts (~8×106 cells) of A549 cell lysates [20 mM Hepes, 100 mM NaCl, 0.5% NP-40, 10% glycerol, 0.2% deoxycholic acid, and 10 mM MgCl2, (pH 7.5)] were incubated with GST-Rho binding domain proteins immobilized on glutathione-Sepharose beads at 4°C overnight. The beads were then washed three times with ice-cold phosphate buffered solution. Bound proteins were eluted from the beads with SDS-PAGE sample buffer. RhoA protein was then analyzed by immunoblotting with anti-RhoA antibody, which was performed as previously described.

Transfection

Small interfering (si)-RNAs against RhoA and the control were synthesized by Shanghai GenePharma Co., Ltd. (Shanghai, China). The sequence of the RhoA siRNA was 5′-GACAUGCUUGCUCAUAGUCTT-3′, and the sequence of the control siRNA was 5′-AAUUCUCCGAACGUGUCACGUUU-3′. A549 cells were transfected with 50 nM control siRNA and RhoA siRNA for 72 h using Lipofectamine 2000 (Invitrogen; Thermo Fisher Scientific, Inc.) according to the manufacturer's instructions.

Statistical analysis

All statistical analyses were performed with SPSS 14.0 for Windows software (SPSS, Inc., Chicago, IL, USA). The results were analyzed using the Student's t-test when two groups were compared, or by one-way analysis of variance with Tukey's post hoc test for multiple comparisons. P<0.05 was considered to indicated a statistically significant difference. Values are presented as the mean ± standard error of the mean.

Results

Synergic antiproliferation effect of gefitinib and GGTI-298 in NSCLC cell lines A549 and HCC827

As GGTI-298 is an inhibitor of a branch of the mevalonate pathway (19), the present study first detected the relative protein expression of HMGCR (the rate-limiting enzyme of mevalonate pathway) in five cell lines (H1975, A549, HCC827, SPCA1 and H1299) and then the two cell lines with a relatively high expression of HMGCR, HCC827 and A549, were selected for further experimentation (Fig. 1A). To determine the effect of GGTI-298 on gefitinib sensitivity, lung cancer cells were treated with gefitinib, GGTI-298 and combination of gefitinib and GGTI-298. The results revealed that the combined treatment of gefitinib and GGTI-298 inhibited cell proliferation more thoroughly than either of the single drug treatments, in a dose-dependent manner (Fig. 1B and C). The CI values indicated synergy, with values of 0.48 and 0.58 for HCC827 and A549 cells, respectively.

Enhanced effects of gefitinib and GGTI-298 on the apoptosis and migration of lung cancer cells

To investigate the effect of GGTI-298 on gefitinib-induced apoptosis, Hoechst assays were conducted. The results demonstrated that a combination of GGTI-298 and gefitinib markedly promoted cell apoptosis when compared with each monotherapy group (Fig. 2A and B). Addition of GGTI-298 increased apoptosis by >25% when compared with gefitinib alone (percentage of apoptotic cells: HCC827, gefitinib vs. gefitinib+GGTI-298=18.3 vs. 48.3%, P<0.01; A549, gefitinib vs. gefitinib+GGTI-298=12 vs. 39.7%, P<0.01; Fig. 2B). Then, wound healing assays were conducted to investigate the influence of GGTI-298 and gefitinib on cell migration. The results revealed that migration ability was significantly inhibited with a combination of gefitinib and GGTI-298 when compared with either single drug treatment following 48 h (Fig. 2C and D). The relative wound widths (wound width at the indicated time/wound width at 0 h) at 48 h of gefitinib and gefitinib+GGTI-298 were 57.5 and 89.0% in HCC827 cells, and 53.5 and 83.6% in A549 cells, respectively (P<0.05; Fig. 2D).

Effects of gefitinib and GGTI-298 on the EGFR signaling pathway and cyclin D1

To gain a greater understanding of the mechanism underlying the combined effect of GGTI-298 and gefitinib, the present study evaluated protein expression by performing immunoblotting. Protein expression was examined following 48 h of treatment with gefitinib and/or GGTI-298. As shown in Fig. 3, GGTI-298 treatment effectively inhibited the expression of pEGFR and pAKT. Furthermore, inhibition of the EGFR-AKT signaling pathway was substantially amplified when GGTI-298 and gefitinib were combined (Fig. 3). However, the present study did not observe marked changes in pERK expression with GGTI-298 treatment alone (Fig. 3). The protein expression of cyclin D1 was also analyzed, which is a proliferation-associated protein and a downstream target of AKT (20). Consistent with the pEGFR and pAKT results, treatment with gefitinib and GGTI-298 induced additional cyclin D1 downregulation when compared with either monotherapy (Fig. 3).

GGTI-298 inhibits the EGFR-AKT signaling pathway through RhoA

The present study then examined the specific mechanism underlying pAKT downregulation. Immunoblotting was employed to analyze the expression of pEGFR (which is expressed upstream of pAKT), PTEN (a negative regulator of pAKT) and pAKT. The results revealed that pEGFR and pAKT expression increased markedly following treatment with EGF, while GGTI-298 treatment reversed the activation of pEGFR and pAKT signaling induced by EGF (Fig. 4A). However, PTEN expression remained essentially unchanged (Fig. 4B). These results revealed that GGTI-298 may have inhibited pAKT signaling primarily through regulating pEGFR expression. Further studies were then conducted to elucidate how GGTI-298 mediates the downregulation of pEGFR. GGTI-298, a GGT1 inhibitor, prevents the geranylgeranylation of small G proteins such as RhoA (19), one of the most intensely studied proteins within the Rho family of GTPases, which is involved in various signaling regulations. Thus, it was hypothesized that GGTI-298 may have inhibited pEGFR through RhoA. Firstly, the present study detected the expression of RhoA. As shown in Fig. 4C, total RhoA increased in A549 cells following treatment with 5 µM GGTI-298 when compared with the untreated (DMSO) group. Then, pull-down assays were conducted to detect active RhoA, which is responsible for the bioactivity of RhoA (21). Compared with control cells, GGTI-298 markedly impaired RhoA activation (Fig. 4C). These results indicated that GGTI-298 inhibited RhoA activation in A549 cells. To detect the involvement of the RhoA protein in EGFR signaling, cells were transfected with siRNAs for RNA interference and the effects were analyzed. As detected by immunoblotting, the transfection of RhoA-specific siRNA resulted in the marked inhibition of total and active RhoA expression, and the downregulation of pEGFR (Fig. 4D). Taken together, these results validated the critical role of RhoA in the regulation of pEGFR and suggested that GGTI-298 may inhibit pEGFR signaling primarily through RhoA in A549 cells.

Discussion

Treatment of advanced NSCLC was revolutionized by EGFR TKIs such as gefitinib and erlotinib. Inadequate inhibition of EGFR signaling is considered to significantly contribute to TKI resistance and disease progression (22). Thus, effective inhibition of EGFR function through multiple mechanisms may augment EGFR TKI activity and overcome TKI resistance. The combination therapy consisting of afatinib and cetuximab previously presented promising clinical efficacy (22), indicating that the co-inhibition of EGFR may a potential method for enhancing the antitumor response. The present study revealed that GGTI-298, a GGT1 inhibitor, inhibited the EGFR signaling pathway mainly through RhoA, and a combined treatment of GGTI-298 and gefitinib impaired the phosphorylation of EGFR more thoroughly than either compound alone, significantly inhibiting cell proliferation and migration, and promoting apoptosis.

EGFR signaling is regulated by multiple protein molecules (2326). In the present study, the EGFR-AKT signaling pathway was inhibited by the GGT1 inhibitor GGTI-298. Further experiments demonstrated that there was a slight upregulation of total RhoA expression following GGTI-298 treatment. However, the expression of active RhoA was markedly inhibited by GGTI-298. These results indicated that GGTI-298 inhibited RhoA activation. Crosstalk involving RhoA and other proteins has been widely reported (14,27,28). In the present study, transfection experiments with siRNA against RhoA decreased the level of pEGFR. A study has previously demonstrated that actin organization served an important role in EGFR localization and activation (29), and the bioactivity of RhoA was critical for actin organization (30), which further supports the present study's results regarding the effect of RhoA on EGFR signaling. Taken together, these results indicated that GGTI-298 may inhibit the EGFR-AKT signaling pathway mainly through RhoA in A549 cells.

AKT and ERK are two main downstream signaling proteins of the EGFR signaling pathway (31). In the present study, the expression of pAKT was markedly decreased following GGTI-298 treatment. Mechanistic studies further revealed that GGTI-298 inhibited the EGFR activation induced by EGF; however, it had little effect on PTEN expression, which negatively regulates pAKT expression. These results suggested that GGTI-298 inhibited pAKT through EGFR, though not PTEN. However, the expression of pERK was essentially unchanged following treatment with GGTI-298. It was hypothesized that Ras, which is upstream of pERK, may mediate activity mainly through farnesylation, but not geranylgeranylation, which is consistent with the results of Sebti et al (32).

Recent studies have demonstrated the potential for mevalonate metabolite depletion to affect EGFR activity (33,34). Targeting the mevalonate pathway inhibited the function of EGFR (35). Combinations of statins, inhibitors of the mevalonate pathway, and an EGFR inhibitor, such as gefitinib and cetuximab, induced a potent synthetical cytotoxicity (31,36). However, it should be noted that statins inhibit the geranylgeranylation of RhoA as well as the synthesis of many other types of bioactive substances such as cholesterol and ubiquinone (19). Cholesterol is a critical component for the structure of animal cell membranes (37), and ubiquinone participates in aerobic cellular respiration (38). Thus, the use of statins may cause multiple side effects associated with these bioactive substances. In the present study, the results revealed that the inhibition of RhoA geranylgeranylation by GGTI-298 also suppressed the EGFR-AKT signaling pathway, and it was not involved in the synthesis of cholesterol and ubiquinone (19). Upon combination treatment with GGTI-298 and gefitinib, the synthetical loss of EGFR-AKT activation in A549 and HCC827 cells resulted in a marked increase in the inhibition of cell proliferation and migration inhibition, and increased levels of apoptosis (Fig. 5). The effect of the GGTI alone was not satisfactory (39), which indicated that a combination-based therapy with a GGTI and other agents may be a potential option for augmenting the antitumor efficiency of GGTIs. Thus, these results were suggestive of the potential of combining these two drugs targeting EGFR, through distinct mechanisms; they therefore represent a novel, promising therapeutic strategy in lung cancer.

In conclusion, the results of the present study suggested that the antitumor efficacy could be markedly increased by combining gefitinib with GGTI-298, a specific inhibitor of protein geranylgeranylation. However, further investigation is required to evaluate the clinical efficacy of this combinatorial treatment in lung cancer.

Acknowledgements

Not applicable.

Funding

The present study was supported by the National Natural Science Foundation of China (grant no. 81621003), the National Key R&D Program of China (grant nos. 2016YFC1303200 and 2016YFC1303203) and the National Natural Science Foundation of China (grant no. 81572731).

Availability of data and materials

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

Authors' contributions

BSL and XYD designed and performed experiments, analyzed data and prepared the manuscript. HWX, HJX and QLT conducted the experiments associated with Figs. 1 and 2. QYG and YZN analyzed the data and contributed to the manuscript. FB supervised these experiments, analyzed data, and prepared the manuscript. All authors reviewed the final manuscript.

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

References

1 

Mok TS, Wu YL, Thongprasert S, Yang CH, Chu DT, Saijo N, Sunpaweravong P, Han B, Margono B, Ichinose Y, et al: Gefitinib or carboplatin-paclitaxel in pulmonary adenocarcinoma. N Engl J Med. 361:947–957. 2009. View Article : Google Scholar : PubMed/NCBI

2 

Maemondo M, Inoue A, Kobayashi K, Sugawara S, Oizumi S, Isobe H, Gemma A, Harada M, Yoshizawa H, Kinoshita I, et al: Gefitinib or chemotherapy for non-small-cell lung cancer with mutated EGFR. N Engl J Med. 362:2380–2388. 2010. View Article : Google Scholar : PubMed/NCBI

3 

Zhou C, Wu YL, Chen G, Feng J, Liu XQ, Wang C, Zhang S, Wang J, Zhou S, Ren S, et al: Erlotinib versus chemotherapy as first-line treatment for patients with advanced EGFR mutation-positive non-small-cell lung cancer (OPTIMAL, CTONG-0802): A multicentre, open-label, randomised, phase 3 study. Lancet Oncol. 12:735–742. 2011. View Article : Google Scholar : PubMed/NCBI

4 

Nicholson RI, Gee JM and Harper ME: EGFR and cancer prognosis. Eur J Cancer. 37 Suppl 4:S9–S15. 2001. View Article : Google Scholar : PubMed/NCBI

5 

Kobayashi S, Boggon TJ, Dayaram T, Jänne PA, Kocher O, Meyerson M, Johnson BE, Eck MJ, Tenen DG and Halmos B: EGFR mutation and resistance of non-small-cell lung cancer to gefitinib. N Engl J Med. 352:786–792. 2005. View Article : Google Scholar : PubMed/NCBI

6 

Yokoyama K, Goodwin GW, Ghomashchi F, Glomset JA and Gelb MH: A protein geranylgeranyltransferase from bovine brain: Implications for protein prenylation specificity. Proc Natl Acad Sci USA. 88:pp. 5302–5306. 1991; View Article : Google Scholar : PubMed/NCBI

7 

Reid TS, Terry KL, Casey PJ and Beese LS: Crystallographic analysis of CaaX prenyltransferases complexed with substrates defines rules of protein substrate selectivity. J Mol Biol. 343:417–433. 2004. View Article : Google Scholar : PubMed/NCBI

8 

Kusama T, Mukai M, Tatsuta M, Matsumoto Y, Nakamura H and Inoue M: Selective inhibition of cancer cell invasion by a geranylgeranyltransferase-I inhibitor. Clin Exp Metastasis. 20:561–567. 2003. View Article : Google Scholar : PubMed/NCBI

9 

Sun J, Qian Y, Chen Z, Marfurt J, Hamilton AD and Sebti SM: The geranylgeranyltransferase I inhibitor GGTI-298 induces hypophosphorylation of retinoblastoma and partner switching of cyclin-dependent kinase inhibitors. A potential mechanism for GGTI-298 antitumor activity. J Biol Chem. 274:6930–6934. 1999. View Article : Google Scholar : PubMed/NCBI

10 

Dan HC, Jiang K, Coppola D, Hamilton A, Nicosia SV, Sebti SM and Cheng JQ: Phosphatidylinositol-3-OH kinase/AKT and survivin pathways as critical targets for geranylgeranyltransferase I inhibitor-induced apoptosis. Oncogene. 23:706–715. 2004. View Article : Google Scholar : PubMed/NCBI

11 

Sonnemann J, Bumbul B and Beck JF: Synergistic activity of the histone deacetylase inhibitor suberoylanilide hydroxamic acid and the bisphosphonate zoledronic acid against prostate cancer cells in vitro. Mol Cancer Ther. 6:2976–2984. 2007. View Article : Google Scholar : PubMed/NCBI

12 

Berndt N, Hamilton AD and Sebti SM: Targeting protein prenylation for cancer therapy. Nat Rev Cancer. 11:775–791. 2011. View Article : Google Scholar : PubMed/NCBI

13 

McGuire TF, Qian Y, Vogt A, Hamilton AD and Sebti SM: Platelet-derived growth factor receptor tyrosine phosphorylation requires protein geranylgeranylation but not farnesylation. J Biol Chem. 271:27402–27407. 1996. View Article : Google Scholar : PubMed/NCBI

14 

Li F, Jiang Q, Shi KJ, Luo H, Yang Y and Xu CM: RhoA modulates functional and physical interaction between ROCK1 and Erk1/2 in selenite-induced apoptosis of leukaemia cells. Cell Death Dis. 4:e7082013. View Article : Google Scholar : PubMed/NCBI

15 

Del Re DP, Miyamoto S and Brown JH: Focal adhesion kinase as a RhoA-activable signaling scaffold mediating Akt activation and cardiomyocyte protection. J Biol Chem. 283:35622–35629. 2008. View Article : Google Scholar : PubMed/NCBI

16 

Basile JR, Gavard J and Gutkind JS: Plexin-B1 utilizes RhoA and Rho kinase to promote the integrin-dependent activation of Akt and ERK and endothelial cell motility. J Biol Chem. 282:34888–34895. 2007. View Article : Google Scholar : PubMed/NCBI

17 

Hebert M, Potin S, Sebbagh M, Bertoglio J, Breard J and Hamelin J: Rho-ROCK-dependent ezrin-radixin-moesin phosphorylation regulates Fas-mediated apoptosis in Jurkat cells. J Immunol. 181:5963–5973. 2008. View Article : Google Scholar : PubMed/NCBI

18 

Chou TC, Motzer RJ, Tong Y and Bosl GJ: Computerized quantitation of synergism and antagonism of taxol, topotecan, and cisplatin against human teratocarcinoma cell growth: A rational approach to clinical protocol design. J Natl Cancer Inst. 86:1517–1524. 1994. View Article : Google Scholar : PubMed/NCBI

19 

Sorrentino G, Ruggeri N, Specchia V, Cordenonsi M, Mano M, Dupont S, Manfrin A, Ingallina E, Sommaggio R, Piazza S, et al: Metabolic control of YAP and TAZ by the mevalonate pathway. Nat Cell Biol. 16:357–366. 2014. View Article : Google Scholar : PubMed/NCBI

20 

Li P, Wei J and Gao X: Insulin promotes the proliferation of human umbilical cord matrix-derived mesenchymal stem cells by activating the Akt-Cyclin D1 axis. Stem Cells Int. 2017:73716152017. View Article : Google Scholar : PubMed/NCBI

21 

Deng W, Gu L, Li X, Zheng J, Zhang Y, Duan B, Cui J, Dong J and Du J: CD24 associates with EGFR and supports EGF/EGFR signaling via RhoA in gastric cancer cells. J Transl Med. 14:322016. View Article : Google Scholar : PubMed/NCBI

22 

Janjigian YY, Smit EF, Groen HJ, Horn L, Gettinger S, Camidge DR, Riely GJ, Wang B, Fu Y, Chand VK, et al: Dual inhibition of EGFR with afatinib and cetuximab in kinase inhibitor-resistant EGFR-mutant lung cancer with and without T790M mutations. Cancer Discov. 4:1036–1045. 2014. View Article : Google Scholar : PubMed/NCBI

23 

Fischer OM, Hart S, Gschwind A and Ullrich A: EGFR signal transactivation in cancer cells. Biochem Soc Trans. 31:1203–1208. 2003. View Article : Google Scholar : PubMed/NCBI

24 

Song S, Honjo S, Jin J, Chang SS, Scott AW, Chen Q, Kalhor N, Correa AM, Hofstetter WL, Albarracin CT, et al: The Hippo coactivator YAP1 mediates EGFR overexpression and confers chemoresistance in esophageal cancer. Clin Cancer Res. 21:2580–2590. 2015. View Article : Google Scholar : PubMed/NCBI

25 

He C, Mao D, Hua G, Lv X, Chen X, Angeletti PC, Dong J, Remmenga SW, Rodabaugh KJ, Zhou J, et al: The Hippo/YAP pathway interacts with EGFR signaling and HPV oncoproteins to regulate cervical cancer progression. EMBO Mol Med. 7:1426–1449. 2015. View Article : Google Scholar : PubMed/NCBI

26 

Chrysogelos SA and Dickson RB: EGF receptor expression, regulation, and function in breast cancer. Breast Cancer Res Treat. 29:29–40. 1994. View Article : Google Scholar : PubMed/NCBI

27 

Zhang Y, Xia H, Ge X, Chen Q, Yuan D, Chen Q, Leng W, Chen L, Tang Q and Bi F: CD44 acts through RhoA to regulate YAP signaling. Cell Signal. 26:2504–2513. 2014. View Article : Google Scholar : PubMed/NCBI

28 

Li H, Ung CY, Ma XH, Li BW, Low BC, Cao ZW and Chen YZ: Simulation of crosstalk between small GTPase RhoA and EGFR-ERK signaling pathway via MEKK1. Bioinformatics. 25:358–364. 2009. View Article : Google Scholar : PubMed/NCBI

29 

Lunn JA, Wong H, Rozengurt E and Walsh JH: Requirement of cortical actin organization for bombesin, endothelin, and EGF receptor internalization. Am J Physiol Cell Physiol. 279:C2019–C2027. 2000. View Article : Google Scholar : PubMed/NCBI

30 

Asanuma K, Yanagida-Asanuma E, Faul C, Tomino Y, Kim K and Mundel P: Synaptopodin orchestrates actin organization and cell motility via regulation of RhoA signalling. Nat Cell Biol. 8:485–491. 2006. View Article : Google Scholar : PubMed/NCBI

31 

Lee J, Lee I, Han B, Park JO, Jang J, Park C and Kang WK: Effect of simvastatin on cetuximab resistance in human colorectal cancer with KRAS mutations. J Natl Cancer Inst. 103:674–688. 2011. View Article : Google Scholar : PubMed/NCBI

32 

Sebti SM and Hamilton AD: Farnesyltransferase and geranylgeranyltransferase I inhibitors and cancer therapy: Lessons from mechanism and bench-to-bedside translational studies. Oncogene. 19:6584–6593. 2000. View Article : Google Scholar : PubMed/NCBI

33 

Mantha AJ, Hanson JE, Goss G, Lagarde AE, Lorimer IA and Dimitroulakos J: Targeting the mevalonate pathway inhibits the function of the epidermal growth factor receptor. Clin Cancer Res. 11:2398–2407. 2005. View Article : Google Scholar : PubMed/NCBI

34 

Niknejad N, Morley M and Dimitroulakos J: Activation of the integrated stress response regulates lovastatin-induced apoptosis. J Biol Chem. 282:29748–29756. 2007. View Article : Google Scholar : PubMed/NCBI

35 

Dimitroulakos J, Lorimer IA and Goss G: Strategies to enhance epidermal growth factor inhibition: Targeting the mevalonate pathway. Clin Cancer Res. 12:4426s–4431s. 2006. View Article : Google Scholar : PubMed/NCBI

36 

Hwang KE, Kwon SJ, Kim YS, Park DS, Kim BR, Yoon KH, Jeong ET and Kim HR: Effect of simvastatin on the resistance to EGFR tyrosine kinase inhibitors in a non-small cell lung cancer with the T790M mutation of EGFR. Exp Cell Res. 323:288–296. 2014. View Article : Google Scholar : PubMed/NCBI

37 

Simons K and Sampaio JL: Membrane organization and lipid rafts. Cold Spring Harb Perspect Biol. 3:a0046972011. View Article : Google Scholar : PubMed/NCBI

38 

Ernster L and Dallner G: Biochemical, physiological and medical aspects of ubiquinone function. Biochim Biophys Acta. 1271:195–204. 1995. View Article : Google Scholar : PubMed/NCBI

39 

Lobell RB, Liu D, Buser CA, Davide JP, DePuy E, Hamilton K, Koblan KS, Lee Y, Mosser S, Motzel SL, et al: Preclinical and clinical pharmacodynamic assessment of L-778,123, a dual inhibitor of farnesyl:protein transferase and geranylgeranyl:protein transferase type-I. Mol Cancer Ther. 1:747–758. 2002.PubMed/NCBI

Related Articles

Journal Cover

October-2018
Volume 18 Issue 4

Print ISSN: 1791-2997
Online ISSN:1791-3004

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Liu BS, Dai XY, Xia HW, Xu HJ, Tang QL, Gong QY, Nie YZ and Bi F: Geranylgeranyl transferase 1 inhibitor GGTI‑298 enhances the anticancer effect of gefitinib. Mol Med Rep 18: 4023-4029, 2018
APA
Liu, B., Dai, X., Xia, H., Xu, H., Tang, Q., Gong, Q. ... Bi, F. (2018). Geranylgeranyl transferase 1 inhibitor GGTI‑298 enhances the anticancer effect of gefitinib. Molecular Medicine Reports, 18, 4023-4029. https://doi.org/10.3892/mmr.2018.9371
MLA
Liu, B., Dai, X., Xia, H., Xu, H., Tang, Q., Gong, Q., Nie, Y., Bi, F."Geranylgeranyl transferase 1 inhibitor GGTI‑298 enhances the anticancer effect of gefitinib". Molecular Medicine Reports 18.4 (2018): 4023-4029.
Chicago
Liu, B., Dai, X., Xia, H., Xu, H., Tang, Q., Gong, Q., Nie, Y., Bi, F."Geranylgeranyl transferase 1 inhibitor GGTI‑298 enhances the anticancer effect of gefitinib". Molecular Medicine Reports 18, no. 4 (2018): 4023-4029. https://doi.org/10.3892/mmr.2018.9371