Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2020 Jun 3.
Published in final edited form as: Curr Opin Genet Dev. 2017 Nov 16;48:82–88. doi: 10.1016/j.gde.2017.11.001

Translation initiation factors and their relevance in cancer

Columba de la Parra 1,, Beth A Walters 1,, Phillip Geter 1, Robert J Schneider 1,*
PMCID: PMC7269109  NIHMSID: NIHMS919096  PMID: 29153484

Abstract

Deregulation of several translation initiation factors occurs in numerous types of cancers. Translation initiation factors are not merely ancillary players in cancer development and progression, but rather, they are key participants in cellular transformation and tumor development. In fact, the altered expression of translation initiation factors is involved in cancer cell survival, metastasis and tumor angiogenesis. Although the exact mechanisms remain to be fully characterized, translation initiation factors comprise novel targets for pharmacologic intervention. Here we review the most recently established roles of initiation factors in cancer development and progression, as well as unique methods used to study translational regulation.

Keywords: translational regulation, cancer, translation initiation factor, protein synthesis

Introduction

Protein synthesis is a costly biochemical process accounting for approximately 50% of the cell’s energy, which is increased in cancer cells [1]. Much of protein synthesis regulation occurs at the step of mRNA translation initiation. Briefly, 40S ribosomal subunits bind eukaryotic initiation factors (eIFs) 1,1A, 3, 5 and the eIF2/met-tRNAi/GTP ternary complex to form a 43S pre-initiation ribosome complex (PIC). The 43S ribosome, with a complex of three proteins that comprise eIF4F (eIF4E, eIF4G, eIF4A) binds the 5’ m7GpppN “cap” and scans the mRNA for an in-context AUG initiation codon (Figure 1) [2]. Importantly, overexpression or increased activity of many initiation factors, including eIF3 and eIF4F, have been implicated in the etiology of many human cancers (Table 1) [3]. However, the exact role that increased levels or activity of initiation factors play in directing cancer physiological behavior remains poorly understood.

Figure 1|. Overview of key steps in mRNA translation initiation.

Figure 1|

The 40S ribosomal subunit interacts with eIF1,eIF1A, eIF3, eIF5 and the ternary complex (eIF2/GTP/Met-tRNAi) forming the 43S pre-initiation complex. The assembly of the eIF4F complex (eIF4G, eIF4E and eIF4A) on the m7GpppN “cap” facilitates the recruitment of the 43S PIC to the mRNA via eIF4G-eIF3 interaction. The interaction of PABP with the scaffolding protein eIF4G might promote the circularization of the mRNA, at least for some part of initiation. eIF4A, a helicase, unwinds the mRNA, with single strand RNA binding protein eIF4B. The eIF4E-mRNA cap complex with eIF4G-eIF3, comprise the 48S ribosomal complex, which progressively searches for the translation initiation codon (typically an AUG). AUG recognition is promoted by hydrolysis of GTP bound to eIF2, an essential step for stable 60S ribosomal subunit recruitment and formation of an active 80S ribosome to initiate protein synthesis. eIF5A promotes peptide bond formation and translation elongation. The inactive eIF2-GDP is recycled to active eIF2-GTP by GTP recycling factor eIF2B.

Table 1|.

Differential expression of certain eIFs and regulatory factors in human tumor specimens

Protein Differential expression Cancer entity and references*
eIF4E Up Breast cancer [1]
Endometrial [2]
Head & neck [3]
Bladder [4]
Cervical [5]
Prostate [6]
eIF4G Up Nasopharyngeal carcinoma [7]
Breast [8,9]
Squamous cell lung [10]
eIF4A1 Up HCC [11]
Melanoma [12]
eIF4A2 Down Breast [13]
Lung [14]
eIF4B Up Breast [15]
eIF4H Up Lung [16]
eIF3a Up Colorectal, Gastric, Breast, Cervical, Lung, Oral cavity squamous [17]
eIF3b Up Bladder, Colon [18]
Prostate [19]
eIF3c Up Testicular seminoma [20]
Glioma [21]
Liver [22]
eIF3d Up Prostate [23]
Breast [24]
eIF3e Down Breast [25]
Colon [26]
Lung [27]
eIF3f Down Melanoma [28]
Pancreatic [28]
eIF3g Up Breast [29]
Bladder [30]
eIF3h Up Breast [31]
HCC [32]
Lung [33]
Colon [34]
eIF3i Up Head & neck [35]
HCC [35]
Colon [36]
eIF3m Up Colon [20]
eIF2α Increased phosphorylation Oropharyngeal carcinoma [37,38]
eIF2B Increased activity Melanoma [39]
eIF5A1/2 Up Increased hypusination in BCR-Abl leukemia [40]
eIF5A2 Up Urinary bladder [41]
Ovarian [42]
Colorectal [43]
HCC [40]
Lung [40]
eIF6 Up Colorectal [44]
Leukemia & lymphoma [44]
Head & neck [45]
Ovarian serous [46]
eIF1 Up HCC [47]
4E-BP1 Down Reduced survival in cancer [1]
Up Reduced tumor grade in breast cancer [1]
Locally advanced breast cancer [8]
Ovarian [48]
PABP Up Liver [49]
*

References listed in Supplementary Information

Mitogenic stimulation and signaling through the PI3K/Akt/mTORC1 pathway stimulates formation of the eIF4F complex, mRNA binding and translation initiation. Cellular stresses, such as amino acid deprivation and hypoxia, which are common in tumors, down-regulate mTORC1 activity, preventing formation of the eIF4F complex and downregulating protein synthesis. Select mRNAs containing special 5’ untranslated regions (5’UTRs) with elements known as Internal Ribosome Entry Sites (IRESs) and upstream Open Reading Frames (uORFs) are capable of maintaining translation under these stress conditions. Many of the mRNAs containing these specialized translation elements are capable of translating under stress and encode transcription factors responsible for enabling the cell to survive or resolve the stress and restore normal protein synthesis. Studies have shown that these unconventional mechanisms of mRNA translation and hyperactivation of mTORC1 signaling are essential for cancer development, resistance and metastasis [4,5].

Translation initiation factors

Commanding initiation: eIF4E and the 4E-BPs

The cap-binding protein eIF4E has been considered in classic models to be a rate-limiting factor that mRNAs compete for, particularly mRNAs with long, structured 5’UTRs that limit their efficient translation initiation. However, a recent study has challenged this view, showing that cancer cells commandeer levels of eIF4E that are typically in excess of what is required for normal cell function and animal development [6]. Tissue culture studies, animal tumor models and human tissue samples all demonstrate that overexpression of eIF4E is important for transformation and drug resistance [7]. Overexpression of eIF4E promotes translational reprogramming enabling transcription, transport and translation of select mRNAs without strongly increasing overall protein synthesis. Many of these selectively translated mRNAs are involved in angiogenesis (VEGF-A), cell proliferation (c-myc), cell survival (Bcl-2) [8] and other aspects of oncogenesis. Many studies have focused on the selective translation of eIF4E-dependent mRNAs and the roles eIF4E, eIF4A and eIF4G in reprogramming the cell’s translatome towards malignancy and the metastatic state [9,10]. For example, experimental downregulation of either eIF4E or eIF4A inhibits melanoma proliferation, survival and invasion through translational downregulation of select mRNAs [11]. Several elements, both structural and sequence-specific, have been identified within the 5’ and 3’ UTRs of mRNAs that confer a requirement for higher levels of eIF4E [6]. Genome-wide mRNA polysome translation profiling analysis, ribosome foot printing analysis and CLIP sequencing have all been used to identify cis- and trans-acting acting mRNA elements that regulate selective mRNA translation [12]. The advantages and limitations of these techniques are described in Table 2.

Table 2|.

Methods to study the translatome

Methods Advantages Disadvantages
Polysomal profiling  • Measures ribosome density on mRNAs for accurate measurement of translation efficiency
 • Enables identification of proteins associated with initiation complexes and/or ribosomes
 • Can assess global and specific changes in the translatome
 • Discovery of novel regulatory elements in mRNAs
 • Labor intensive and requires a large amount of starting material
 • High chance of RNA degradation
 • Does not provide individual nucleotide resolution
 • Requires specialized and expensive equipment
Ribosomal profiling  • Quantitatively identify mRNA associated with ribosomes
 • Capabilities of examining initiation, elongation, or termination
 • Labor intensive
 • Requires meticulous expertise in bioinformatics analysis
 • Requires a large amount of starting material
Ribosome-affinity purification  • Does not require a large amount of starting material
 • Can be used to study the translatome of quiescent cells
 • Quick and robust technique
 • Requires the tagging of ribosomes
 • Does not provide a high resolution translational map
 • Cannot discriminate between number of bound ribosomes
CLIP techniques  • Unique method that map direct RNA–protein interactions site using UV cross-linking
 • Individual-nucleotide-resolution
 • Can be performed on any tissue type
 • Labor intensive and technically challenging
 • Availability of specific antibodies
 • Require meticulous and expertise in bioinformatics analysis
 • May lead to the identification of artificial interactions

eIF4E function is regulated by its availability, which in turn is controlled by the eIF4E binding proteins (4E-BPs), which are mTORC1-regulated repressor proteins that competitively bind eIF4E, blocking its interaction with eIF4G, thereby preventing eIF4E-mediated cap-dependent mRNA translation. mTORC1 hyperphosphorylates the 4E-BPs causing the release of eIF4E for translation initiation and restoration of protein synthesis. Hyperactivated mTORC1 is a hallmark of cancer cells and promotes transformation, cell proliferation, survival and metastasis [1315]. The loss of 4E-BP1 (or overexpression of eIF4E) confers resistance to chemotherapeutic agents and accelerates tumorigenesis in mice [15], although high levels of 4E-BP1 in prostate cancer are associated with resistance to PI3K inhibitors [16].

eIF4E is phosphorylated by kinases Mnk1/2 at S209 via the MAPK/ERK/MNK pathway. eIF4E S209 phosphorylation is important for the EMT (Epithelial to Mesenchymal Transition), cancer cell survival and metastasis, likely by promoting selective mRNA translation [17]. Increased expression of eIF4E and S209 phosphorylation may also promote tumor initiation by cancer stem cells (CSCs) [18]. Higher levels of eIF4E were shown to increase selective translation of TGF-β and Wnt/β-catenin mRNAs that are critical for the oncogenic process [19,20]. Therefore, translational control via eIF4E availability acts as a convergence point for signaling of many oncogenic pathways to promote tumorigenesis.

The cap connection: eIF4G

eIF4G is a large protein, ~180 kDa, typically described as a scaffolding protein because of its interactions with eIF4E, eIF4A, the 43S PIC, MNKs and poly A binding protein (PABP), among other functions. Further, eIF4G has been associated with a switch from cap-dependent to cap-independent mRNA translation initiation during stress conditions [Supple Ref 8]. Overexpression of eIF4G is strongly associated with malignant transformation of immortalized cells in tissue culture, as well as advanced stage breast cancer and metastasis in patients. It is also highly overexpressed in inflammatory breast cancer (IBC) and cancer cells with a high proliferative index (Table 1) [Supple Ref 9] [21,22]. Genomic amplification of eIF4G occurs in ~30% of squamous cell lung cancers [Supple Ref 10].

The cap disconnection: DAP5

The eIF4G1 homolog DAP5 (NAT1, eIF4G2, p97) lacks binding sites for eIF4E and PABP, and has a number of interesting properties that underlie its role in cancer. DAP5 has been primarily studied in tissue culture where it promotes alternative translation of mRNAs that utilize cap-independent mechanisms, typically by IRES elements. Some of these mRNAs are specifically translated during cancer cell invasion and metastasis, as well as survival (anti-apoptosis), and include Bcl-2, Apaf-1, cIAP1, CDK1 and p53 [23,24]. Recent studies, including ribosomal and polysomal profiling in embryonic stem cells, have shown that DAP5 is involved in the translation of proteins required for cell differentiation [25,26]. Nevertheless, the exact mechanism of action and the role of DAP5 in vivo in cancer remains to be understood.

Unwinding complexity: RNA helicase eIF4A and ancillary factors

Overexpression of the abundant isoform eIF4A1 is observed in primary hepatocellular carcinoma (HCC), cervical cancer and some melanomas [Supple Ref 11, 12] whereas its activity is increased in some other cancer types, possibly by increased mTOR activity (Figure 1). mRNAs that translationally benefit from increased levels or activity of eIF4A are involved in cell proliferation, cell cycle progression, cell survival and angiogenesis. Studies have shown that increased levels of eIF4A are required for T-cell acute lymphoblastic leukemia (T-ALL) development and leukemia maintenance, through a translationally controlled mechanism [27].

eIF4B and eIF4H assist in eIF4A-mediated mRNA secondary structure unwinding which facilitates 40S ribosome subunit binding and scanning. The mitotic kinase MELK (Maternal Embryonic Leucine zipper Kinase) can phosphorylate eIF4B, which promotes proliferation of highly malignant breast cancer cells [Supple Ref 15]. Reduced expression of eIF4B selectively represses translation of many mRNAs, including the apoptosis inhibitor Bcl-2 [28]. eIF4H is reportedly up-regulated in lung adenocarcinomas where it may confer resistance to chemotherapy while increasing translation of mRNAs involved in cell growth and survival [29].

Bringing the pieces together: eIF3, a complex player in human cancer

Although the exact mechanism(s) by which the largest and most complex initiation factor, eIF3 (~800 kDa), affects cancer malignancy is still being elucidated, growing evidence suggests that eIF3 plays a critical role in human cancer and may be an actionable target for drug development. eIF3 is essential for recruiting translation factors and 40S ribosome subunits to the mRNA (Figure 1). Ten of the thirteen eIF3 protein subunits (eIF3a - m) have been linked to human cancers (Table 1), many by overexpression (a, b, c, h or i) [30] and possibly by direct interaction with certain cancer-related mRNAs [31]. In fact, silencing eIF3a reverses the malignant phenotype of human lung and breast cancer cell lines [32], and downregulates the CDK inhibitor p27(Kip1) [33]. A recent study suggests that eIF3b may be required for the proliferation of human osteosarcoma cells, acting by regulating tumor necrosis factor receptor superfamily member 21 (TNFRSF21), which activates the NF-kB pathway [34]. eIF3c also interacts with and impairs the function of the tumor suppressor proteins schwannomin and merlin [35], promotes cell proliferation of primary hepatocellular carcinoma (HCC) cells [Supple Ref 22] and is overexpressed in gliomas [36]. eIF3d is associated with several types of cancers (Table 1), including bladder, glioblastoma, prostate and breast [Supple Ref 23]. Silencing eIF3d impairs cancer cell proliferation and invasion by suppressing Wnt/β-catenin signaling and CDK1 [37,38] [Supple Ref 24]. It was recently demonstrated that eIF3d is a highly specialized cap binding protein, at least for the oncogenic transcription factor c-JUN [39].

Among other cancers, colorectal and HCC are associated with higher levels of eIF3h and eIF3i (Table 1). eIF3h is amplified in a variety of human carcinomas and may be a target of the anti-cancer agent harmine-derived beta-carboline, CM16 [40]. Overexpression of eIF3f was recently shown to suppress Akt and ERK signaling, increase p53 protein levels and inhibit clusterin protein expression, which promotes cancer cell proliferation and reduces chemosensitivity [41]. The remaining subunits of eIF3 (g, j, k, l and m) are the least studied in cancer development. Recent findings suggest the nuclear redistribution of eIF3g might play a role in the development and progression of breast cancer cells [Supple Ref 29], and is involved in DNA degradation during apoptosis [Supple Ref 30].

Coming together: the joining of ribosome subunits, and the roles of eIF2 and eIF2B

eIF2 is a complex of three proteins, α, β, and γ, which carry the initiating methionyl-tRNA and GTP. GTP hydrolysis on eIF2 promotes initiation codon recognition. The released eIF2-GDP must be recycled back to eIF2-GTP for initiation to occur, which is mediated by eIF2B, a multifunctional GEF (Guanine Exchange Factor). Phosphorylation of the α-subunit of eIF2 at S51 prevents GDP-to-GTP exchange by increasing the affinity of eIF2B for eIF2-GDP, thereby blocking protein synthesis. Recent studies suggest that prolonged eIF2α phosphorylation can promote cell survival, transformation and drug resistance, while in contrast, other studies suggest that eIF2α phosphorylation can trigger apoptosis [Supple Ref 37 and 38], which may be cell and tumor-type specific. A recent study showed that embryonic stem cells (ESCs) have high levels of constitutive eIF2α phosphorylation, which increases translation of uORF-containing mRNAs encoding cellular pluripotency factors [42]. This may also occur in cancer stem cells (CSCs) that promote metastasis and recurrence. Further, eIF2B is involved in the cell’s ability to adapt to prolonged activation of the integrated stress response (ISR)/unfolded protein response (UPR) [43]. A small molecule inhibitor of the ISR (ISRIB) enhances the eIF2B GEF activity, rendering cells insensitive to translational inhibition by eIF2α phosphorylation. In cancer cells, eIF2B-mediated translational reprogramming protects cancer cells from apoptosis [44,45].

Recruiting other initiator factors: eIF5A, eIF5B, eIF6, and PABP

eIF5A consists of two highly related isoforms (eIF5A1 and 2) that promote peptide bond formation by initiating 80S ribosomes [3] and translation elongation [46]. eIF5A1 is overexpressed in lung adenocarcinomas and BCR-ABL-positive leukemia [Supple Ref 40]. eIF5A1 and 2 are also the only proteins known to be hypusinated, where an amino-butyl residue is added to lysine [47]. The more poorly studied eIF5A2 protein is expressed at low levels if at all in normal cells, but is strongly overexpressed by gene amplification, often with PIK3CA, in human ovarian, brain, testicular and colorectal cancers (Table 1).

Translation initiation factor eIF6 is an anti-association factor. It binds to the 60S subunit and prevents its association with 40S ribosomal subunits in the absence of mRNA. Phosphorylation of eIF6 releases 60S ribosomal subunits, allowing for functional formation of 80S ribosomes [2]. eIF6 is up-regulated in abundance and activity in many human cancers, the latter by protein kinase C (PKC) βII phosphorylation on S235 [48]. However, the role of eIF6 in tumorigenesis is complex and incompletely understood. In a B-cell lymphoma animal model, reduced expression of eIF6 is associated with longer survival and prevention of myc-mediated lymphomagenesis, consistent with eIF6 overexpression in human lymphomas and many solid tumors [Supple Ref 61]. Over-expression of eIF6 in transformed cell lines markedly increases cell migration and invasion, possibly by regulating proteins that control tumor cell motility such as cdc42 [49].

It is debatable whether PABP should be considered a translation factor, but nevertheless, it promotes translation initiation by binding the poly (A) tail of mRNA and eIF4G and stimulates translation initiation. However, the role of PABP in cancer development is not clear. PABP is overexpressed in high grade HCC tissues, and its silencing reduces cancer cell proliferation. Independent of a direct translation factor interaction, PABP interaction with AGO2, which is involved in RNA–mediated gene silencing, has been reported to enhance miRNA inhibition of tumor suppressor genes [Supple Ref 63].

Concluding remarks and perspectives

Although detailed mechanisms by which certain translation initiation factors are involved in human cancer remain to be understood, there is no doubt they function as drivers and not just passengers in the oncogenic process. As reviewed here and by others, deregulation or altered expression of certain translation factors can promote cancer development and its progression, often acting on tumor cell survival, growth factor-independent proliferation, hypoxia responses, neo-vascularization, evasion of apoptosis and the response of tumor cells to their microenvironment, generally through selective mRNA translation. Strong progress has been made using genome-wide translation technologies to study the translatome in cancer, providing evidence for altered translational control in development, progression and metastasis of human cancers, as well as enabling identification of new targets for drug development. Understanding the detailed roles of translation initiation factors in cancer offers the opportunity for improving actual therapeutic benefit in human cancer.

Supplementary Material

dummy_label

ACKNOWLEDGMENTS

The investigators were supported by the following grants: NIH RO1CA178509, NIH RO1CA207893, BCRF-16-143 (RJS), ACS PF-16-095-01-RMC (CDLP), T32 CA9161-41 (BAW), NIH GM066704, Training in Pharmacological Sciences 5T32 GM066704, UL1 TR00038 from the National Center for Advancing Translational Sciences (NCATS), and National Institutes of Health, the Howard Hughes Medical Institute (PG). We apologize to those whose work we were not able to include due to size constraints.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Conflict of interest statement

The authors declare no conflicts of interest.

References and recommended reading

Papers of particular interest have been highlighted as:

• of special interest

•• of outstanding interest

  • 1.Lane N, Martin W: The energetics of genome complexity. Nature 2010, 467:929–934. [DOI] [PubMed] [Google Scholar]
  • 2.Chu J, Cargnello M, Topisirovic I, Pelletier J: Translation Initiation Factors: Reprogramming Protein Synthesis in Cancer. Trends Cell Biol 2016, 26:918–933. [DOI] [PubMed] [Google Scholar]
  • 3.Silvera D, Formenti SC, Schneider RJ: Translational control in cancer. Nat Rev Cancer 2010, 10:254–266. [DOI] [PubMed] [Google Scholar]
  • 4.••.Sendoel A, Dunn JG, Rodriguez EH, Naik S, Gomez NC, Hurwitz B, Levorse J, Dill BD, Schramek D, Molina H, et al. : Translation from unconventional 5′ start sites drives tumour initiation. Nature 2017, 541:494–499. [DOI] [PMC free article] [PubMed] [Google Scholar]; This article uses ribosome footprinting to identify translation from unconventional 5’ start sites as a mechanism driving tumor initiation.
  • 5.•.Demosthenous C, Han JJ, Stenson MJ, Maurer MJ, Wellik LE, Link B, Hege K, Dogan A, Sotomayor E, Witzig T, et al. : Translation initiation complex eIF4F is a therapeutic target for dual mTOR kinase inhibitors in non-Hodgkin lymphoma. Oncotarget 2015, 6:9488–9501. [DOI] [PMC free article] [PubMed] [Google Scholar]; This article highlights the importance of the regulation of the eIF4F complex in aggressive lymphoma as well as demonstrating the therapeutic potential of dual mTOR therapies.
  • 6.Truitt Morgan L, Conn Crystal s, Shi Z, Pang X, Tokuyasu T, Coady Alison M, Seo Y, Barna M, Ruggero D: Differential Requirements for eIF4E Dose in Normal Development and Cancer. Cell 2015, 162:59–71. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Fagan DH, Fettig LM, Avdulov S, Beckwith H, Peterson MS, Ho Y-Y, Wang F, Polunovsky VA, Yee D: Acquired Tamoxifen Resistance in MCF-7 Breast Cancer Cells Requires Hyperactivation of eIF4F-Mediated Translation. Hormones and Cancer 2017:1–11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Mamane Y, Petroulakis E, Rong L, Yoshida K, Ler LW, Sonenberg N: eIF4E - from translation to transformation. Oncogene 2004, 23:3172–3179. [DOI] [PubMed] [Google Scholar]
  • 9.Gandin V, Masvidal L, Hulea L, Gravel S-P, Cargnello M, McLaughlan S, Cai Y, Balanathan P, Morita M, Rajakumar A, et al. : nanoCAGE reveals 5’ UTR features that define specific modes of translation of functionally related MTOR-sensitive mRNAs. Genome Research 2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Nandagopal N, Roux PP: Regulation of global and specific mRNA translation by the mTOR signaling pathway. Translation 2015, 3:e983402. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Joyce CE, Yanez AG, Mori A, Yoda A, Carroll JS, Novina CD: Differential Regulation of the Melanoma Proteome by eIF4A1 and eIF4E. Cancer Research 2017, 77:613. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.••.King HA, Gerber AP: Translatome profiling: methods for genome-scale analysis of mRNA translation. Briefings in Functional Genomics 2016, 15:22–31. [DOI] [PubMed] [Google Scholar]; This review details the different techniques used to study the translatome and future implications of these technologies.
  • 13.Bi C, Zhang X, Lu T, Zhang X, Wang X, Meng B, Zhang H, Wang P, Vose JM, Chan WC, et al. : Inhibition of 4EBP phosphorylation mediates the cytotoxic effect of mechanistic target of rapamycin kinase inhibitors in aggressive B-cell lymphomas. Haematologica 2017, 102:755–764. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.•.Hsieh AC, Liu Y, Edlind MP, Ingolia NT, Janes MR, Sher A, Shi EY, Stumpf CR, Christensen C, Bonham MJ, et al. : The translational landscape of mTOR signalling steers cancer initiation and metastasis. Nature 2012, 485:55–61. [DOI] [PMC free article] [PubMed] [Google Scholar]; This article describes a control that couples cell type-specific 4E-BP1 abundance with the regulation of global protein synthesis rates. This coupling renders each epithelial cell type of the prostate sensitive or resistant to inhibitors of the PI3K-AKT-mTOR signaling pathway.
  • 15.Qu Y, Zhao R, Wang H, Chang K, Yang X, Zhou X, Dai B, Zhu Y, Shi G, Zhang H, et al. : Phosphorylated 4EBP1 is associated with tumor progression and poor prognosis in Xp11.2 translocation renal cell carcinoma. 2016, 6:23594. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Hsieh AC, Nguyen HG, Wen L, Edlind MP, Carroll PR, Kim W, Ruggero D: Cell type–specific abundance of 4EBP1 primes prostate cancer sensitivity or resistance to PI3K pathway inhibitors. Science Signaling 2015, 8:ra116–ra116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Kumar K, Chow CR, Ebine K, Arslan AD, Kwok B, Bentrem DJ, Eckerdt FD, Platanias LC, Munshi HG: Differential Regulation of ZEB1 and EMT by MAPK-Interacting Protein Kinases (MNK) and eIF4E in Pancreatic Cancer. Molecular Cancer Research 2016, 14:216. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Robichaud N, del Rincon SV, Huor B, Alain T, Petruccelli LA, Hearnden J, Goncalves C, Grotegut S, Spruck CH, Furic L, et al. : Phosphorylation of eIF4E promotes EMT and metastasis via translational control of SNAIL and MMP-3. Oncogene 2015, 34:2032–2042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.••.Decarlo L, Mestel C, Barcellos-Hoff MH, Schneider RJ: Eukaryotic Translation Initiation Factor 4E Is a Feed-Forward Translational Coactivator of Transforming Growth Factor beta Early Protransforming Events in Breast Epithelial Cells. Mol Cell Biol 2015, 35:2597–2609. [DOI] [PMC free article] [PubMed] [Google Scholar]; This article demonstrates the important relationship of elevated expression of eIF4E and the activation of the transforming growth factor β (TGF-β) pathway, which further promotes many hallmarks of cancer such as increased cell invasion and cell survival. As shown in a breast cancer model, eIF4E confers selective mRNA translation and reprogramming of nonmalignant cells to an invasive phenotype.
  • 20.Pettersson F, del Rincon SV, Emond A, Huor B, Ngan E, Ng J, Dobocan MC, Siegel PM, Miller WH: Genetic and Pharmacologic Inhibition of eIF4E Reduces Breast Cancer Cell Migration, Invasion, and Metastasis. Cancer Research 2015. [DOI] [PubMed] [Google Scholar]
  • 21.Fukuchi-Shimogori T, Ishii I, Kashiwagi K, Mashiba H, Ekimoto H, Igarashi K: Malignant Transformation by Overproduction of Translation Initiation Factor eIF4G. Cancer Research 1997, 57:5041–5044. [PubMed] [Google Scholar]
  • 22.Silvera D, Arju R, Darvishian F, Levine PH, Zolfaghari L, Goldberg J, Hochman T, Formenti SC, Schneider RJ: Essential role for eIF4GI overexpression in the pathogenesis of inflammatory breast cancer. Nat. Cell Biol 2009, 11:903–908. [DOI] [PubMed] [Google Scholar]
  • 23.Hundsdoerfer P, Thoma C, Hentze MW: Eukaryotic translation initiation factor 4GI and p97 promote cellular internal ribosome entry sequence-driven translation. Proc Natl Acad Sci U S A 2005, 102:13421–13426. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Weingarten-Gabbay S, Khan D, Liberman N, Yoffe Y, Bialik S, Das S, Oren M, Kimchi A: The translation initiation factor DAP5 promotes IRES-driven translation of p53 mRNA. Oncogene 2014, 33:611–618. [DOI] [PubMed] [Google Scholar]
  • 25.•.Sugiyama H, Takahashi K, Yamamoto T, Iwasaki M, Narita M, Nakamura M, Rand TA, Nakagawa M, Watanabe A, Yamanaka S: Nat1 promotes translation of specific proteins that induce differentiation of mouse embryonic stem cells. Proceedings of the National Academy of Sciences of the United States of America 2017, 114:340–345. [DOI] [PMC free article] [PubMed] [Google Scholar]; The authors demonstrate through ribosomal profiling that Dap5 is involved in the translation of proteins required for cell differentiation in embryonic stem cells.
  • 26.Yoffe Y, David M, Kalaora R, Povodovski L, Friedlander G, Feldmesser E, Ainbinder E, Saada A, Bialik S, Kimchi A: Cap-independent translation by DAP5 controls cell fate decisions in human embryonic stem cells. Genes & Development 2016, 30:1991–2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Wolfe AL, Singh K, Zhong Y, Drewe P, Rajasekhar VK, Sanghvi VR, Mavrakis KJ, Jiang M, Roderick JE, Van der Meulen J, et al. : RNA G-quadruplexes cause eIF4A-dependent oncogene translation in cancer. Nature 2014, 513:65–70. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Shahbazian D, Parsyan A, Petroulakis E, Topisirovic I, Martineau Y, Gibbs BF, Svitkin Y, Sonenberg N: Control of Cell Survival and Proliferation by Mammalian Eukaryotic Initiation Factor 4B. Molecular and Cellular Biology 2010, 30:1478–1485. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.•.Vaysse C, Philippe C, Martineau Y, Quelen C, Hieblot C, Renaud C, Nicaise Y, Desquesnes A, Pannese M, Filleron T, et al. : Key contribution of eIF4H-mediated translational control in tumor promotion. Oncotarget 2015, 6:39924–39940. [DOI] [PMC free article] [PubMed] [Google Scholar]; This article reveals that the mRNA levels of eIF4H are up-regulated in lung adenocarcinoma tissues resistant to chemotherapy. The mechanism of action of eIF4H in cancer could be associated with the translation control of mRNA involved in growth and survival.
  • 30.Zhang L, Pan X, Hershey JWB: Individual Overexpression of Five Subunits of Human Translation Initiation Factor eIF3 Promotes Malignant Transformation of Immortal Fibroblast Cells. Journal of Biological Chemistry 2007, 282:5790–5800. [DOI] [PubMed] [Google Scholar]
  • 31.••.Lee ASY, Kranzusch PJ, Cate JHD: eIF3 targets cell proliferation mRNAs for translational activation or repression. Nature 2015, 522:111–114. [DOI] [PMC free article] [PubMed] [Google Scholar]; In this study, the authors use a PAR-CLIP approach to identify the mRNAs regulated through direct interaction with eIF3. The study demonstrates that specific subunits of eIF3 bind to mRNAs involved in cell cycling, differentiation and apoptosis.
  • 32.Saletta F, Rahmanto YS, Siafakas AR, Richardson DR: Cellular Iron Depletion and the Mechanisms Involved in the Iron-dependent Regulation of the Growth Arrest and DNA Damage Family of Genes. The Journal of Biological Chemistry 2011,286:35396–35406. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Zhang Y, Yu J-J, Tian Y, Li Z-Z, Zhang C-Y, Zhang S-F, Cao L-Q, Zhang Y, Qian C-Y, Zhang W, et al. : eIF3a improve cisplatin sensitivity in ovarian cancer by regulating XPC and p27(Kip1) translation. Oncotarget 2015, 6:25441–25451. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Choi YJ, Lee YS, Lee HW, Shim DM, Seo SW: Silencing of translation initiation factor eIF3b promotes apoptosis in osteosarcoma cells. Bone & Joint Research 2017, 6:186–193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Scoles DR, Yong WH, Qin Y, Wawrowsky K, Pulst SM: Schwannomin inhibits tumorigenesis through direct interaction with the eukaryotic initiation factor subunit c (eIF3c). Human Molecular Genetics 2006, 15:1059–1070. [DOI] [PubMed] [Google Scholar]
  • 36.Hao J, Liang C, Jiao B: Eukaryotic translation initiation factor 3, subunit C is overexpressed and promotes cell proliferation in human glioma U-87 MG cells. Oncology Letters 2015, 9:2525–2533. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Gao Y, Teng J, Hong Y, Qu F, Ren J, Li L, Pan X, Chen L, Yin L, Xu D, et al. : The oncogenic role of EIF3D is associated with increased cell cycle progression and motility in prostate cancer. Medical Oncology 2015, 32:196. [DOI] [PubMed] [Google Scholar]
  • 38.Fan Y, Guo Y: Knockdown of eIF3D inhibits breast cancer cell proliferation and invasion through suppressing the Wnt/β-catenin signaling pathway. International Journal of Clinical and Experimental Pathology 2015, 8:10420–10427. [PMC free article] [PubMed] [Google Scholar]
  • 39.••.Lee AS, Kranzusch PJ, Doudna JA, Cate JH: eIF3d is an mRNA cap-binding protein that is required for specialized translation initiation. Nature 2016, 536:96–99. [DOI] [PMC free article] [PubMed] [Google Scholar]; This study shows for the first time the cap-binding capacity of eIF3d for the oncogenic transcription factor c-JUN and exposes a new model on cap-dependent translation.
  • 40.Carvalho A, Chu J, Meinguet C, Kiss R, Vandenbussche G, Masereel B, Wouters J, Kornienko A, Pelletier J, Mathieu V: A harmine-derived beta-carboline displays anti-cancer effects in vitro by targeting protein synthesis. European Journal of Pharmacology 2017, 805:25–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Lee J-Y, Kim H-J, Rho SB, Lee S-H: eIF3f reduces tumor growth by directly interrupting clusterin with anti-apoptotic property in cancer cells. Oncotarget 2016, 7:18541–18557. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.•.Friend K, Brooks HA, Propson NE, Thomson JA, Kimble J: Embryonic Stem Cell Growth Factors Regulate eIF2α Phosphorylation. PLOS ONE 2015, 10:e0139076. [DOI] [PMC free article] [PubMed] [Google Scholar]; This article suggests a role of constitutive eIF2α phosphorylation in promoting the selective translation of embryonic stem cell factors encoded by mRNAs with uORFs.
  • 43.Martin L, Kimball SR, Gardner LB: Regulation of the Unfolded Protein Response by eIF2Bδ Isoforms. Journal of Biological Chemistry 2010, 285:31944–31953. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Sidrauski C, McGeachy AM, Ingolia NT, Walter P: The small molecule ISRIB reverses the effects of eIF2α phosphorylation on translation and stress granule assembly. eLife 2015, 4:e05033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.••.Falletta P, Sanchez-del-Campo L, Chauhan J, Effern M, Kenyon A, Kershaw CJ, Siddaway R, Lisle R, Freter R, Daniels MJ, et al. : Translation reprogramming is an evolutionarily conserved driver of phenotypic plasticity and therapeutic resistance in melanoma. Genes & Development 2017. [DOI] [PMC free article] [PubMed] [Google Scholar]; This article underscores the significance of translational reprogramming as a mediator of cancer developemtn and therapeutic resistance.
  • 46.Caraglia M, Park MH, Wolff EC, Marra M, Abbruzzese A: eIF5A isoforms and cancer: two brothers for two functions? Amino acids 2013, 44:103–109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Park MH, Nishimura K, Zanelli CF, Valentini SR: Functional significance of eIF5A and its hypusine modification in eukaryotes. Amino acids 2010, 38:491–500. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Grosso S, Volta V, Sala Leonardo A, Vietri M, Marchisio Pier C, Ron D, Biffo S: PKCβII modulates translation independently from mTOR and through RACK1. Biochemical Journal 2008, 415:77. [DOI] [PubMed] [Google Scholar]
  • 49.Pinzaglia M, Montaldo C, Polinari D, Simone M, La Teana A, Tripodi M, Mancone C, Londei P, Benelli D: eIF6 over-expression increases the motility and invasiveness of cancer cells by modulating the expression of a critical subset of membrane-bound proteins. BMC Cancer 2015, 15:131. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

dummy_label

RESOURCES