Next Article in Journal
Pregabalin–Tolperisone Combination to Treat Neuropathic Pain: Improved Analgesia and Reduced Side Effects in Rats
Next Article in Special Issue
New Antimicrobials for Gram-Positive Sustained Infections: A Comprehensive Guide for Clinicians
Previous Article in Journal
Synthesis of 1,2,3-Triazole-Containing Methoxylated Cinnamides and Their Antileishmanial Activity against the Leishmania braziliensis Species
Previous Article in Special Issue
Collaterally Sensitive β-Lactam Drugs as an Effective Therapy against the Pre-Existing Methicillin Resistant Staphylococcus aureus (MRSA) Biofilms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Surmounting Cancer Drug Resistance: New Perspective on RNA-Binding Proteins

School of Basic Medical Sciences and State Key Laboratory of Southwestern Chinese Medicine Resources, Chengdu University of Traditional Chinese Medicine, Chengdu 611137, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this paper.
Pharmaceuticals 2023, 16(8), 1114; https://doi.org/10.3390/ph16081114
Submission received: 30 June 2023 / Revised: 20 July 2023 / Accepted: 2 August 2023 / Published: 7 August 2023
(This article belongs to the Special Issue Multi-Drug Resistance)

Abstract

:
RNA-binding proteins (RBPs), being pivotal elements in both physiological and pathological processes, possess the ability to directly impact RNA, thereby exerting a profound influence on cellular life. Furthermore, the dysregulation of RBPs not only induces alterations in the expression levels of genes associated with cancer but also impairs the occurrence of post-transcriptional regulatory mechanisms. Consequently, these circumstances can give rise to aberrations in cellular processes, ultimately resulting in alterations within the proteome. An aberrant proteome can disrupt the equilibrium between oncogenes and tumor suppressor genes, promoting cancer progression. Given their significant role in modulating gene expression and post-transcriptional regulation, directing therapeutic interventions towards RBPs represents a viable strategy for combating drug resistance in cancer treatment. RBPs possess significant potential as diagnostic and prognostic markers for diverse cancer types. Gaining comprehensive insights into the structure and functionality of RBPs, along with delving deeper into the molecular mechanisms underlying RBPs in tumor drug resistance, can enhance cancer treatment strategies and augment the prognostic outcomes for individuals afflicted with cancer.

1. Introduction

Alterations in gene regulatory networks cause changes in the cellular proteome, leading to altered phenotypes and abnormal cell proliferation. Abnormal gene expression and abnormal regulatory functions in cells are responsible for the development of cancer. Typically, this is thought to be associated with mutations and translocations of cancer-related genes. However, recent studies have shown that a group of cancer-associated RBPs can regulate the expression of specific genes in cells by interfering with post-transcriptional mechanisms such as mRNA stability and translation, thereby contributing to the development and progression of cancer [1,2,3] (Figure 1).
The mediators that influence the occurrence of post-transcriptional regulatory mechanisms are cis-acting elements, mainly found in the 5′- and 3′-untranslated regions (UTR) of mRNA, and trans-acting transcription factors that can specifically recognize and interact with cis-acting elements [4,5]. Among them, trans-acting transcription factors mainly include RBPs and non-coding RNAs (ncRNAs), such as microRNAs (miRNAs) and long-stranded non-coding RNAs (LncRNAs) [6]. Different types of RBP regulatory networks can regulate post-transcriptional mechanisms such as mRNA stability and translation by interfering with regular sequence motifs and RNA secondary structure, thereby regulating the expression levels of cancer-related genes [7]. Various RBPs have been found to regulate a wide range of genes associated with cancer development.
RBPs are highly conserved and abundant proteins that play a key role in maintaining normal and stable gene expression and are key players in gene expression and a variety of post-transcriptional regulatory mechanisms involved in various aspects of RNA regulation [8,9], including transcription, translation [10], splicing [11], polyadenylation [12], stability [13], and localization [14]. RBPs can form ribonucleoprotein complexes (RNPs) through direct interactions with other proteins or as scaffolds for coding or ncRNAs, and the life cycle of RNAs can be affected by the interaction of RBPs with other proteins, resulting in abnormal protein phenotypes that contribute to tumorigenesis and progression [15].
There are approximately 1914 RBP genes in the human genome, accounting for 7.5% of protein-coding genes [15,16]. When cancer occurs, many RBP confinements are opened, and the expression levels and localization of RBPs are altered, resulting in changes in the expression levels of proto-oncogenes, oncogenes, and genomic stability-related genes. For example, the expression levels of cancer-related genes such as CD44 and VEGF are affected by changes in RBPs as splicing factors. Also, based on genome-wide analysis, as they largely affect cell growth and proliferation, many RBPs are considered to be key factors in cancer development and progression. The majority of RBPs are comprised of one or more RNA-binding domains (RBDs), such as RRMs, KH structural domains, dsRBDs, znFs domains, and PAZ structural domains [17]. For instance, hnRNPA1 possesses RRMs and hnRNP, IGF2BPS contains RRMs and KHs, and ZEB1 is composed of ZnF-CCCHs and ZnF-CCHCs. The diverse arrangements of these RBD combinations enable RBPs to regulate various mRNA functions [18].
Overall, the broad function of RBPs confers a central role in the RBP-RNA regulatory network during disease progression, particularly in cancer [19]. Thus, exploring the mechanisms of drug resistance induced by RBP abnormalities may provide potential approaches to improving cancer prognosis. This paper reviews the function and structure of RBPs, highlights the regulatory role of RBPs in therapeutic resistance, and discusses RBPs as potential targets for cancer therapy.

2. Definition of RBPs and Their Structural Features

RBPs are proteins that bind directly to the RBD in RNA to form RNPs, which are directly involved in cellular life activities such as synthesis, processing, translocation, translation, degradation, and function of coding and non-coding RNAs [20]. RBP ensures that genetic information can be expressed as protein through DNA and RNA, which means that RBP is essential for all physiological and pathological processes [21].
The RBD is the functional unit that forms the RBP. The presence of one or more typical RBDs in a specific order ensures that RBPs can bind properly to RNA. RBPs can be classified as canonical or non-canonical according to the presence or absence of an RBD, and we refer to proteins containing RBDs as canonical RBPs. It has been shown that many RBPs lack a classical RBD. The absence of a classical RBD leads to structural instability of the RBP, which affects the ability of the RBP to bind to RNA and, thus, the stability and translation of mRNA [22]. Many RBDs are very small (<100 residues) and only use a small number of residues to interact directly with RNA. A typical and well-structured RBP consists of multiple repetitive sequences that contain only a few specific basic domains. These repetitive sequences can be arranged in different combinations with specific RNAs, thus allowing for a diversity of RBP functions [18]. Furthermore, precise recognition of proteins can be achieved by rearranging these typical structural domains [23]. Here we present several classical RBDs, including RRM, KH structural domains, dsRBDs, and znFs domains (Figure 2).

2.1. RNA Recognition Motifs (RRM)

RRM, also known as the RBD or ribonucleoprotein motif RNP, is by far the most common and well-studied RNA binding module. A typical RRM consists of 80–90 amino acids and has a βαββββαβ topology, forming two alpha helices against a four-stranded anti-parallel beta-sheet. Over 10,000 RRMs have been identified, and approximately 0.5–1% of all human genes contain RRMs. Most act during post-transcriptional gene expression, where RNA recognition occurs [23]. In most cases, recognition usually occurs on the surface of the β-sheet, mediating post-recognition binding through three conserved residues: an Arg/Lys salt bridge forming the phosphodiester backbone and two aromatic residues that create stacking interactions with the nucleobase. RRM-containing proteins bind to different RNAs by recognizing two to eight nucleotides of single-stranded RNA; a single RRM can recognize four to eight nucleotides by using exposed loops and other secondary structure elements not present in the canonical structure [24,25,26], and this general recognition mechanism is present in many but not all RRMs [27,28], with some of these structural domains even interacting with proteins rather than RNA [29,30,31,32,33,34]. Some individual RRMs can bind to RNA with high specificity, but in many cases, multiple structural domains are required to define specificity because the number of nucleotides recognized by a single RRM is usually too small to define a unique binding sequence.

2.2. The K Homology (KH) Domain

The K-Homology domain is a single-strand, sequence-specific nucleic acid binding domain first identified in heterogeneous ribonucleoprotein K (hnRNPK) and commonly found in proteins that regulate gene expression in eukaryotes and prokaryotes [35]. Compared with RRM, the KH structural domain is smaller, consisting of 70 amino acids, and is located at the center of a very functionally important structural domain. Typically, this structural domain has a conserved GXXG loop, which is able to link the two α-helices and the β-strand [18], recognizing four nucleotides in ssRNA or ssDNA. All KH structural domains form a three-stranded β-sheet stacked on three α-helices. Based on their topology, they can be divided into two subfamilies (type I: βααβββα topology; type II: αβββααβ topology [36]). As with RRMs, individual RBPs with repeating KH structural domains can increase the chance of binding. These KH structural domains are also able to increase binding specificity independently or in concert [37].

2.3. Double-Stranded RNA-Binding Structural Domains (DsRBDs)

DsRBDs or motifs (dsRBMs) are the third most common RBDs, consisting of 65–70 amino acids [38], with amino acids folding into an α1β1β2β3α2 structure, forming an antiparallel β-sheet and laterally attached to an α-helix on one face. DsRBDs play a role in viral protection, RNAi, and cellular transport of proteins found in proteins such as ribonuclease III and RNA editing enzymes. DsRBD has a central role in binding double-stranded RNA (dsRNA) or highly structured RNA, but it does not make specific contact with nucleobases [38]. DsRBD usually occurs as a tandem repeat sequence or in combination with other functional RNA-binding structural domains. Unlike the RRM and KH structural domains, most intermolecular contacts of the dsRBD are sequence independent, forming hydrogen bond contacts with the 2′-OH group and the phosphate backbone 46 [39]. The presence of multiple dsRBDs can confer specificity to specific structures as they are able to recognize certain arrangements of the RNA helix [40,41,42]. Stacking interactions are rarely seen in dsRBDs, a phenomenon that provides some evidence to explain the low affinity of dsRBDs for RNA targets [43,44].

2.4. Zinc Finger Structural Domains (ZnFs)

ZnFs are classical DNA-binding proteins that can also bind to RNA, proteins, and small molecules [45]. ZnFs comprise a large family of proteins, with members averaging 30 amino acids in size. ZnFs have the ability to form a simple ββα topology, and Zn2+ is able to coordinate residues in the β hairpin of this structure with the α-helix for alignment [46]. ZnF isoforms that interact with RNA include the CCHC (zinc finger joints), CCCH, CCCC (RanBP2), and CCHH isoforms, where the cysteine (C) and histidine (H) arrangement determines the RNA binding preferences of the ZnFs, displaying a range of sequence and structural specificities [25,47]. ZnFs can use some of the same residues to recognize both nucleic acids, but different DNA and RNA structures dictate different structural arrangements of ZnFs on nucleic acid templates. Zinc finger joints (CCHC) recognize stem-loop elements in RNA (or ssDNA) by contacting bases in the loop and the phosphate backbone of the stem. Subtypes with multiple CCCH and CCCC tend to bind the 3 nucleotide repeats of RNA, whereas subtypes with abundant CCHH interact with single- and double-stranded RNA and dsRNA. CCHH ZnFs interact with DNA primarily by forming direct hydrogen bonds with Watson-Crick base pairs, using them to recognize residues in α-helix 60 [48]. In contrast, TFIIIA binds RNA by specific contact with two RNA loops through the recognition helices of fingers 4 and 6.

3. The Function of RBP and Its Aberrant Expression in Cancer

3.1. Function of RBP

3.1.1. Pre-mRNA Alternative Splicing (AS)

The core spliceosome is a key regulator of selective splicing and is composed of small nuclear proteins (snRNPs), peptides, and RBPs. RBPs can form trans-acting splicing factors with cis-regulatory precursors to regulate splicing events [49]. The most representative splicing regulators are the serine/arginine-rich (SR) family of proteins and the hnRNP proteins, which affect the recruitment of spliceosomal components and the selection of splice sites by recognizing and binding exons for enhancer splicing and by participating in RNA-protein interactions [50]. HnRNP proteins act as another splicing regulator, mainly by inhibiting the splicing process by disrupting exon recognition [51] (Figure 3).

3.1.2. Alternative Polyadenylation of mRNA (APA)

APA is a key component in the generation of mature RNA transcripts, and the mRNA 3′-UTR is an important site where APA occurs. In addition, APA can mediate 3′-end cleavage and polyadenylation (CPA), resulting in 3′-UTRs of variable length [17]. RBP is a major component of the APA machinery, which includes cleavage stimulating factors, cleavage factor I, and cleavage factor II, among others. In addition to direct involvement in the APA pathway, RBPs can recruit or compete with polyadenylation-associated proteins to regulate the CPA process in target mRNAs [52]. HnRNP and SR family proteins can also regulate the polyadenylation process in mRNAs [53] (Figure 4).

3.1.3. RNA Stability

The 5′ terminal 7-methylguanosine (m7G) cap and the 3′ poly (A) tail of RNA determine the stability of RNA, which is regulated by RBPs [54,55]. ARE-associated RBPs such as AUF1 (AU-rich elemental RNA-binding protein 1), HuR (Human antigen R), TTP (trite-traproline), IGF2BP (Insulin-like growth factor 2 mRNA-binding protein) family proteins, and Wig1 can recognize and bind to AREs, thereby regulating the stability of target mRNAs [56] (Figure 5a).

3.1.4. RNA Localization

RBP binding to RNA coordinates the localization of target RNA and its translation location, which is essential for the regulation of RNA stability and translation. RBPs recognize and bind to specific sequences in the target RNA, and the bound RBPs subsequently assemble into multiunit complexes that bind the RNA to cytoskeletal molecular motors and transport it to the target site (Figure 6).

3.1.5. Translation

RBPs are dynamic components of ribonucleoproteins, which are key factors in the process of cellular translation. Different translation-associated RBPs bind differently to mRNAs, resulting in different translation efficiencies [57]. RBPs promote translation by recognizing relevant elements in the structure of the target RNA [58,59,60,61] (Figure 5b).

3.2. Aberrant Expression in Cancer

Over the past few decades, studies have shown that dysregulation of RBPs contributes to cancer treatment resistance. RBPs can target many target RNAs to influence the characteristics of cancer cells. RBPs can bind not only to mRNA exons, introns, and untranslated regions (UTRs), but also to ncRNAs such as miRNAs, siRNAs (small interfering RNAs), telomerase RNAs, tRNA, small nucleolar RNA (snoRNA), and spliced small nucleolar RNA (snRNA). Upon binding to secondary structures formed by ncRNAs, RBPs are involved in regulating several processes of gene expression, such as RNA modification, protein localization and secretion, and chromosome remodeling [21]. Dysregulation of RBPs affects the expression levels of target RNAs associated with the cancer phenotype, such as proliferation, apoptosis, angiogenesis, senescence, and EMT/invasion/metastasis [17]. Dysfunctional RBPs lead to abnormalities in cellular processes, and these abnormalities cause alterations in the proteome. An abnormal proteome can cause an imbalance in the checks and balances between oncogenes, leading to cancer. For example, in human CRC (colorectal cancer) tissues, the expression of the long non-coding RNA (lncRNA) NEAT1 is upregulated, which is associated with poor prognosis in CRC patients. NEAT1 interacts with the RBP DDX5 and activates the Wnt/β-linked protein signaling pathway, which further promotes tumorigenesis [62]; the molecular chaperones required for DICER1 to function are the RBP TRBP, and mutations in TRBP can also lead to aberrant miRNA expression as well as proliferation and differentiation of cancer cells [63,64]; in addition, the epithelial-mesenchymal transition (EMT)-specific transcription factor ZEB1 protein directly inhibits the mRNA levels of epithelial splicing regulatory protein 1 (ESRP1), leading to cell surface antigen Increased expression of various spliceosomes of CD44 and induction of stem cell-like and aggressive cells in lung, breast, and pancreatic cancers. A single RBP can bind multiple target mRNAs, and different RBPs can also regulate cellular processes by sharing the same specificity on mRNAs bound to the same mRNA [65]. This complex, overlapping post-transcriptional RNA regulation works together with other cellular process-participating factors to form a more complex regulatory network that plays an irreplaceable role in the regulation of cancer-related gene expression [66].

4. Molecular Mechanisms of RBPs in Tumor Drug Resistance

4.1. Selective Splicing

The processing of pre-mRNAs interacting with RBPs, particularly selective splicing, is affected when the function of RBPs is dysfunctional [67]. Selective splicing is an important mechanism that drives transcriptome diversity, and AS alters gene expression by generating multiple mRNA transcripts from a single precursor gene, through which a variety of unique transcripts are generated and subsequently translated during cellular processes [68,69,70]. Dysregulation of AS is extremely common in the development and progression of cancer. Most of the proteins involved in the AS process are RBPs, which bind to target RNAs and thus influence the splicing process [71,72,73]. Different species of RBPs, such as SRSF, hnRNPs, PTB, ESRP, and QKI, can exert positive or negative regulatory effects on selective splicing. Through selective splicing processing, pre-mRNAs can be transcribed into a variety of mRNA variants with different stability and protein-coding potential. Different mRNA splice variants contribute to increased protein diversity in cancer [74]. In cancer cells, aberrant RBP expression and aberrant RNA splicing are observed more frequently compared with normal cells [75], and these alterations lead to changes in the expression levels of proto-oncogenes and oncogenes, thus promoting cancer development and progression [49,74,76].
In a variety of cancers (including HCC), RBP abnormalities usually lead to a number of abnormal AS events. Examples include increased or decreased exons and dysregulation of splicing factors [68,77,78]. Also, dysfunction of hnRNP and SR can be observed in a variety of cancers, suggesting a key role for selective splicing in cancer progression. The complexes that determine tissue- and tumor-specific splicing events are formed by the association of core proteins of the splicing machinery with RBPs, and these complexes can collaborate with or antagonize spliceosomal activity depending on the position of the RBPs binding site in relation to the regulated exon.

4.2. Polyadenylation

Selective polyadenylation is an extremely critical step in the generation of mature RNA transcripts [79]. APA occurs mainly within the 3′-UTR of mRNA and produces 3′-UTRs of different lengths through 3′-end CPA [80]. APA can regulate the stability, subcellular localization, and translation efficiency of target mRNAs by altering the 3′-UTR length [81]. This tissue-specific mechanism can convert the coding sequence of transcripts to produce different isoforms, affecting their metabolism in various ways, including stability, nuclear export, and cellular localization [82]. RBPs can regulate the CPA of target mRNAs by recruiting or competing with APA machinery proteins [52], thereby affecting APA progression. Similar to selective splicing, selective polyadenylation enables a single gene to have encoded sequences or different transcripts in the untranslated region and can affect the function of the protein encoded by that gene by altering the coding sequence.
The C-terminal region of CPEB family proteins contains two RRMs and two zinc finger-like motifs, while CPEB family proteins also possess a variable N-terminal region [83]. CPEB 1–4 can regulate the length of the poly (A) tail of mRNA using cytoplasmic polyadenylation elements (CPE), resulting in a shorter 3′-UTR for transcripts associated with cell proliferation and tumorigenesis. CPEB 1 regulates nuclear transcription-specific APA and can modulate cytoplasmic polyadenylation-mediated translation [84]. If CPEB1 is absent, the poly (A) tail of mRNA is prolonged, and mRNA translation of MMP9 is enhanced in breast cancer cells [85].

4.3. Stability

The stability of eukaryotic RNA is ensured by two stability determinants: the m7G cap at the 5′ end and the 3′ poly (A) tail. The normal process of RNA degradation is an essential part of the gene expression process. The 3′-poly (A) tail and the adulterated 5′ cap protect the structure and stability of eukaryotic RNA well [86]. These two components protect the mRNA from decay and facilitate the initiation of translation [54]. In order to initiate decay, either of these structures must be disrupted, including deacetylation of the poly (A) tail, removal of the 5′ end m7G cap (decamp), or the mRNA, which must be cleaved internally by a nuclease-mediated reaction. The ARE is a structure that has been extensively studied as being closely related to mRNA stability [87] and is present in approximately 16% of transcripts [22,88]. Typically, cellular transcripts containing AREs are inherently unstable. The most common cis-unstable element in the 3′-UTR is the ARE, which achieves decay of mRNA mainly through deacetylase-mediated shortening of the poly (A) tail. The ARE-binding protein (AUBP) is actively involved in this destabilization process. Based on different cellular stimuli, AUBP is involved in destabilization or participates in stabilization [89].
In cancer, these ARE-containing transcripts appear to overexpress oncogenes, growth factors and their receptors, inflammatory mediators, and cell cycle genes, among others. This phenomenon suggests that mRNA stability plays an important role in cancer production and metastasis [90,91]. For mRNA stability, RBPs have a bidirectional regulatory role. For example, AUF1, HuR, PTBP, IGF2BP family proteins, and MCPIP1 can enhance mRNA stability in cancer, regulate the protein expression of their target genes, or accelerate the degradation of lncRNA in cancer [56].
The AU-rich elemental RNA-binding protein 1 (AUF1, also known as hnRNPD) contains a family of four splice isoforms that become p37AUF1, p40AUF1, p42AUF1, and p45AUF1 depending on their molecular weight [92]. AUF1 has a bidirectional role in maintaining mRNA stability in different systems. Transgenic mice overexpressing the p37AUF1 isoform develop spontaneous sarcomas, while cancer-associated transcripts such as CCND1, FOS, and MYC mRNA are significantly increased. In addition, AUF1 can trigger an anti-tumor response by destabilizing mRNAs encoding the anti-apoptotic protein BCL2 and the pro-inflammatory factors GM-CS, IL-6, IL-10, and TNF-α.
HuR expression has been reported to be extremely elevated in several cancer types [93]. Overexpression of HuR in cancer can stabilize several ARE-containing mRNAs encoding cell cycle regulators, such as CCNA 1, CCNB 1, CCND 1, and CCNE 1, which contribute to the proliferation of cancer cells. In addition to promoting HCC by stabilizing mRNA transcripts to promote cancer cell proliferation, HuR can also intervene in cancer progression by regulating RNA stability through IncRNA [94,95]. HuR can promote the growth of glioma cells through post-transcriptional regulation while enhancing their resistance to a variety of drugs, such as etoposide, topotecan, and cisplatin [96]. This process occurs mainly through HuR binding to the 2′UTR of MCL3, BCL1, and BLxL to stabilize their mRNAs, thereby regulating some genes involved in apoptosis. In contrast, mRNA levels of C-Myc, Wnt5a, and P27 were down-regulated by HuR.
The three main members of the PTBP family are PTBP1, PTBP2, and PTBP3, of which PTBP3 promotes the growth and metastasis of cancer cells and also prevents the degradation of mRNAs. This is achieved by promoting EMT in breast tumor cells and by regulating the expression of the transcription factor ZEB3 by binding to the 1′UTR of their mRNA, respectively [97].
In addition, the IGF2BP family can play an important role in the tumor development process by regulating mRNA stability. On the one hand, IGF2BP1 protects the degradation of PTEN mRNA and thus promotes the migration of tumor cells [98], while on the other hand, IGF2BP1 promotes the expression of C-Myc and MKI67 mRNA to regulate the proliferation and apoptosis of hepatocellular carcinoma cells [99]. In addition, IGF2BP1 reduces the stability of HULC mRNA, which is specifically and highly expressed in hepatocellular carcinoma. IGF2BP2 has the ability to stabilize HMGA1 mRNA and RAF196 mRNA, thereby enhancing cancer cell viability and promoting cancer cell proliferation [100]. IGF2BP3 promotes the degradation of EIF4E-BP2 mRNA, thereby promoting cervical cancer cell proliferation [101].

4.4. Subcellular Localization

It is well known that the subcellular localization of mRNAs or lncRNAs is closely related to their stability and translation. The biological function of lncRNAs is determined by their subcellular localization [102]. RBPs associated with cancer often bind to RNAs to coordinate their localization and translation [95,103]. RBPs typically associate transcripts with cytoskeletal molecular motors by binding to sequences in the 3′-UTR, which in turn deliver RNPs to specific subcellular compartments, thereby localizing mRNAs within the cell. This mechanism is essential for establishing and maintaining cell polarity but is frequently altered during carcinogenesis [104].
Among the CPEB family members, CPEB1 regulates the localization of ZO-1 mRNA, which encodes a key tight junction component. When CPEB1 is depleted, ZO-1 mRNA is randomly distributed, and the central lumen does not form properly, resulting in a loss of epithelial cell polarity [105].
RBP Tia1 can interact with a variety of cancer-related mRNAs and is involved in multiple aspects of cancer development and progression, such as cancer cell proliferation, apoptosis, invasion, metastasis, angiogenesis, and immune escape [106,107]. Tia1 functions to regulate the translational silencing and localization of cellular stress-related transcripts (p53 mRNA), a function that is achieved by binding to an mRNA, and this function is achieved by binding to a subset of P53 mRNA [108]. When the DNA of Tia1 is disrupted, Tia1 and mRNA dissociate, and P53 mRNA is released from the stress granule and bound to the polyplex, which causes mRNA migration and translation.
IGF2BP1 is extremely important for the development of cancer. After transcriptional processes have occurred, they influence mRNA expression in oncogenes by regulating the subcellular localization of mRNAs encoding β-actin, E-calmodulin, α-actin, and Arp-16 (components of the Arp 2/3 complex), thereby promoting tumor cell proliferation and growth, invasion, and chemical metabolism [109,110]. IGF2BP1 promotes the transfer of polar cell ACTB transcripts to actin-rich protrusions in polarized cells [98]. IGF2BP1 binds to β-actin transcripts in the nucleus and induces translational silencing of mRNAs in the cytoplasm without affecting their stability. Elevated expression of IGF2BP1 can be found in primary tumor tissues such as breast, colon, and non-small cell lung cancers [111,112,113]. In contrast, IGF2BP1 expression is reduced in metastatic cells. Reduced expression of IGF2BP1 inhibits the transport and local expression of adhesion- and migration-related target mRNAs [110,114]. When IGF2BP1 is silenced, intercellular junctions are disrupted, and cell adhesion is reduced, resulting in enhanced cell migration and invasion.

4.5. Translation

The translation process involves the coordination of three major components: initiation, elongation, and termination. For the regulation of mRNA translation, most interventions are made at the initiation step [115,116]. EIF4E (Eukaryotic translation initiation factor 4 E) is a 5′ cap-dependent translation initiation factor, and due to the broad function and expression pattern of translation factors and associated RBPs, the RBP of elF4E plays a crucial role in the induction of translation by facilitating the ribosome assembly and loading process [117,118,119,120]. Almost all major oncogenic signaling pathways that are altered in cancer, such as PI3K/AKT/mTOR, RAS/MAPK, and Wnt/b-linked proteins, result in translational dysregulation [57,119,120].
The translation of the oncogenic program is guided by structure- and sequence-specific regulatory elements. One of the most widely studied 5′-UTR structural elements is the internal ribosome entry site (IRES), which promotes translation in a cap-dependent manner by directly recruiting ribosomes in association with IRES trans-acting factors (ITAFs). IFAFs are composed of a set of RBPs that act on IRES RNA elements located in the 5′-UTR in a cap-independent manner, recruiting ribosomal subunits, which in turn undergo translation initiation [115,116]. The development of cancer is associated with the dysregulation of cap structure-dependent translation.
La ribonucleoprotein structural domain family member 3 (LARP 3) can promote cancer cell survival and invasion by inducing IRES-mediated translation of anti-apoptotic XIAP and EMT-associated LAMB 1 mRNAs [121,122].
EIF4E is a component of the eIF4F translation initiation complex. EIF4E interacts with eIF4A (RNA decapping enzyme) and eIF4G (scaffold molecule) and binds the 5′-terminal m7G cap of mRNA. EIF4F, the 5′-cap binding complex, induces mRNA cyclization and translation activation via ribosome loading. Among the components of the eIF4F complex, eIF4E is commonly overexpressed in different types of tumors, which is associated with a poor prognosis [123]. A study showed that eIF4E haploinsufficient mice were physiologically normal but significantly resistant to tumor formation [124]. This suggests that mammalian cells express eIF4E above the threshold required for normal translational control but that the factor may be upregulated by cancer cells to induce the translation of specific subgroups of oncogenic mRNAs.
HuR can enhance the translation of transcripts encoding ProTα, P53, and MSI 1 by binding to the 3β-UTR in a miRNA-independent manner, and increased cytoplasmic ProTα mRNA levels have been associated with anti-apoptotic effects in cancer cells [125,126]. Meanwhile, HuR can regulate MSI1 mRNA and its translation in glioblastoma. HuR and PTB (also known as hnRNP I) bind to the HIF-1α5 β-UTR and 3 β-UTR, respectively, thereby synergistically increasing HIF-1α translation in the presence of hypoxia [127]. In addition, HuR can negatively regulate mRNA translation in a miRNA-dependent manner [128].
Here, we summary the molecular mechanisms of RBP in the cancer (Table 1).

5. RBP-Targeted Cancer Therapy

5.1. Small Molecules

Cancer therapeutics targeting RBPs, represented by small-molecule inhibitors, are in high demand [159]. Small-molecule probes that target and interfere with RNA binding are new approaches that have recently been developed to block RBP function and have been characterized by researchers [160]. It has been shown that some small-molecule drugs can regulate RBP function by targeting it in various human diseases, including cancer, thus promising to improve therapeutic resistance to RBP transmission (Figure 7b).
Abnormal expression of HuR is often seen in various types of cancer, and this is closely associated with treatment resistance in cancer. A range of small-molecule inhibitors targeting HuR have been developed to alleviate treatment resistance in cancer. Among them are, for example, MS-444, DHTS, and AZA-9, nanomolecular inhibitors of HuR that block its RNA-binding activity by targeting RRM1 and RRM2 of HuR. In the presence of these inhibitors, the ARE binding activity of HuR to the target RNA is significantly reduced [161,162]. Since HuR-ARE interactions are important for the stability of many mRNAs associated with therapeutic resistance, the potential exists for the use of these inhibitors to reverse therapeutic resistance [163]. In colorectal cancer, after treatment with small molecule inhibitors, HuR is unable to bind to ARE-containing mRNA targets (e.g., IL-2, IL-1β, TNF-α, COX-2, and C-fos), its function is disrupted, and it exhibits selective anticancer effects [161,162,164,165]. In the treatment of non-small cell lung and thyroid cancer progression, CMLD-2 has been shown to be an effective HuR inhibitor that binds competitively to HuR, thereby inhibiting HuR’s cell growth and proliferation-promoting effects and the expression of related genes, increasing apoptosis in tumor cells, and thereby slowing the progression of cancer [165,166]. Wu et al. identified, by establishing a high-throughput screening system, several compounds that inhibited the activity of HuR by disrupting HuR-mRNA interactions. Therefore, these compounds could be used as cancer therapeutic agents targeting HuR [167].
EIF4E has also been associated with therapeutic resistance in a variety of cancers. Overexpression of eIF4E has been reported in a variety of cancers, which correlates with the aggressive phenotype of tumors. Ribavirin, an antiviral guanosine analog, blocks the translocation and translation of oncogenes regulated by eIF4E [168]. As another eIF4E inhibitor, N-7 benzyl guanosine monophosphate tryptamine phosphonamidite prodrug (4Ei-1) upregulates gemcitabine chemosensitivity in lung and breast cancers by inhibiting the mRNA cap-binding ability of eIF4E and degrading the eIF4E protease [169].
Musashi RNA-binding protein 1 (MSI-1) has a role in enhancing therapeutic resistance in cancer. MSI-1 enhances therapeutic resistance by increasing the expression of DNA repair-associated proteins DNA-PKcs and EGFR, and therefore cancer treatment, in a way that identifies potential inhibitors of MSI proteins using high-throughput analysis [170]. In one experiment, a natural compound called (−) -gossypol, a small molecule known to have anticancer effects on various cancers, was screened by fluorescence polarization. (−) -Gossypol can occupy the consensus RNA binding site of MSI-1, thereby disrupting the interaction between MSI-1 and its target mRNA to achieve inhibition of MSI-1 [171]. Another study found that oleic acid could interact with the RRM1 motif in MSI and induce a change in MSI conformation, thereby inhibiting MSI binding to target mRNAs. In addition, oleic acid can also influence cancer progression by upregulating MSI-1 expression levels to affect cell division [172].

5.2. Therapeutic Peptides

Therapeutic peptides contain 55 amino acids or less and have advantages such as high specificity, selectivity, small size, ease of modification, and biocompatibility [173]. Currently, various therapeutic peptides based on the reversal of therapeutic resistance are being designed for cancer treatment (Figure 7c). The 4EBP-based therapeutic peptides hinder the progression of ovarian cancer by binding to eIF4E and disrupting tumor growth. At the same time, the anti-cancer effects of the fusion peptide are even more pronounced when this therapeutic peptide is combined with an analog of gonadotropin-releasing hormone (GnRH). The GnRH-4EBP fusion peptide can inhibit ovarian tumor growth in an epithelial ovarian cancer xenograft model without any cytotoxicity [174]. P53 as an oncogene can be inhibited by RBM38 by suppressing eIF53E-mediated translation of P4 mRNA. Another 4EBP-based therapeutic peptide, Pep8, which mimics the eIF8E bonding domain, can antagonize RBM38, thereby promoting P53 expression, reducing cancer development, and slowing down cancer progression [175].

5.3. ASO and siRNA-Based Strategies

ASOs (antisense oligonucleotides) and siRNAs are commonly used to regulate gene expression, and both oligonucleotides can bind their target RNAs via Watson-Crick base pairing to regulate splicing, target miRNAs, and inhibit translation. SiRNAs are used to target RBPs to reverse their abnormal expression. SiRNAs can bind to target genes and subsequently cause gene silencing. SiRNA-based therapies have shown efficacy and clinical safety in clinical practice. SiRNA efficacy has been investigated in several tumor types (Figure 7c). SiRNAs targeting eIF4E inhibit tumor growth and stimulate the cytotoxic effects of cisplatin in human breast cancer in vitro and in vivo, suggesting that cisplatin treatment in combination with eIF4E-siRNA therapy would be more successful [176]. HuR can also be targeted with siRNA, a combined chemical-biological drug delivery system that can be used to actively target desired cells or tissues, leading to increased reactive oxidative stress, which increases radiosensitivity and sensitizes TNBC cells to radiation [177]. Furthermore, loading siRNA targeting HuR into folic acid (FA)-coupled polyamide-amine dendrimer (Den)-based nanoparticles revealed that the formulation was effective in reducing HuR expression and cell proliferation in lung cancer cells. In combination with cis-diaminoplatinum (CDDP), this nanoparticle exhibited synergistically improved anticancer effects and reduced cytotoxicity [178]. Similarly, when siRNA was targeted to HuR in a mouse lung cancer model, HuR expression was disrupted and tumor growth was inhibited in a liposomal nanoparticle delivery system [179]. Specific ASOs targeting eIF4E can inhibit tumor growth by suppressing the translation of target mRNAs such as VEGF, Survivin, C-Myc, Cyclin D1, and BCL-2 [137,180,181]. A number of FDA-approved or clinically tested drugs enable gene-specific silencing, and although not yet approved for cancer therapy, many researchers are conducting relevant clinical and non-clinical studies [182]. For example, in the context of targeting RBP, ISIS 183750, an ASO drug targeting eIF4E, disrupts the proliferation of colorectal cancer cells [183]. In addition, ASOs and siRNAs targeting MSI can inhibit tumor growth in pancreatic and ovarian cancers in vitro and in vivo [184,185] (Figure 7d).

5.4. Binding Partners and Downstream Effectors

The strategy of targeting RBPs for cancer treatment is not a bad option, but the success rate has been a stumbling block to this strategy due to the difficulty of either direct targeting or specific selection of RBPs. However, it is also reassuring to note that RBPs have great potential to become markers for different types of cancer, both diagnostically and prognostically, due to their aberrant expression profiles in different environments and the variation in their regulatory functions on target mRNAs. Gene regulation of RBP in rodent tumor models has shown exciting inhibitory effects on tumor development in different types of cancer studies, including leukemia. RBP’s therapeutic potential has been extensively studied, as some RBPs have been found to be promising biomarkers of prognosis for patients with different types of cancer [186,187]. For example, the aggressive phenotype or poor prognosis of several cancers has been associated with the upregulation of IGF2BP3 [163,188].
In order to improve the success rate of cancer treatment, many researchers have focused on binding partners of RBPs and downstream effectors of RBPs. With advances in bioinformatics methods and experimental techniques, the veil of RBP-binding chaperones involved in cancer progression is gradually being lifted. Direct targets of various RBPs in different cancer types have been identified through microarray and sequencing techniques [189].
Downstream effectors of RBPs are also good choices for therapeutic targets. Among them, MYC is, on the one hand, a common effector of many RBPs. On the other hand, it can also inversely regulate the stability of several RBPs, including HuR, hnRNPA1, and hnRNPH, to participate in cancer progression. Because of its ability to inversely regulate the function of RBPs, it is possible to intervene in RBP-influenced cancer progression by targeting and regulating the transcriptional levels and activity of MYC. For example, small molecule inhibitors such as I-BET, JQ1, and MMS417 have been identified as transcriptional repressors of MYC, and these small molecule inhibitors can significantly inhibit cancer progression. ASO is another MYC transcriptional repressor that can down-regulate MYC transcriptional levels by reducing MYC translation and splicing events. In addition to suppressing MYC activity, it is also possible to inhibit MYC and its binding partners by mutating MYC and using small molecule inhibitors on the interaction between MYC and its binding partners (Figure 7d).
Here, we summary information on small molecule inhibitors, therapeutic peptides, antisense oligonucleotides (ASOs), siRNAs, binding partners and downstream effectors (Table 2).

6. Conclusions and Future Perspectives

6.1. Conclusions

Dysregulation of RBPs is observed across multiple cancer types, leading to modifications in the expression of genes associated with tumorigenesis and influencing the progression of cancer. Simultaneously, RBPs possess the capability to either induce or suppress the expression of genes implicated in cancer, thereby contributing to the heterogeneity observed in cancer progression. On the one hand, RBPs govern the delivery of genetic material from DNA to RNA; on the other hand, in the context of cancer, RBPs participate in the energy metabolism of cancer cells and facilitate their evasion of immune surveillance by modulating mRNA stability and the translation process. Therefore, within the field of cancer biology, RBP has great research and therapeutic potential.
Although much has been explored and many researchers have achieved much in identifying RBPs and RBP-RNA interactions, there are still many questions and unknowns that deserve continued in-depth investigation. For example, the use of RBP- RNA-specific binding for targeted drug therapy still requires work. In order to introduce RBP-targeted drugs into cancer therapy, we need better identification of RBPs and a deeper understanding of the target RNAs that interact with them, as well as the signaling pathways and regulatory mechanisms involved in the interactions, so that we can target RBPs accordingly for different cancer patients. Methods such as primer extension (SHAPE), chromatin immunoprecipitation (CHIP), RNA immunoprecipitation (RIP), and Ribosome analysis (Ribo-Seq) can be used to investigate the feasibility and effectiveness of targeted RBP therapy. Therefore, the use of targeted RBP therapy in cancer requires us to explore more deeply in the future to study the interaction between RBPs and RNA.
On the therapeutic side, several researchers have experimentally validated the use of small molecule inhibitors of RBPs, therapeutic peptides, and ASO-specific antagonism of RBPs in vitro, and the experimental results have shown encouraging effectiveness. Nevertheless, questions remain as to whether RBPs can be used as an effective marker for cancer diagnosis, prognosis, and treatment.

6.2. Future Perspectives

In summary, we make some suggestions for future RBP research. First, although the mechanical alterations triggered by dysregulated RBP expression have been shown, this is only the tip of the iceberg, and there are still a large number of unidentified mechanistic alterations mediated by RBPs that require more and more in-depth scientific studies. The deeper our understanding of RBP’s dysregulation-mediated mechanical alterations, the better our understanding of RBP’s function will be. Secondly, RBPs act primarily through binding to target RNAs to form RNP complexes and a complex regulatory network that regulates cellular processes through various post-transcriptional mechanisms. A change in one node of the regulatory network can have a butterfly effect, so further deciphering the unknown post-transcriptional mechanisms is urgent. Thirdly, RBP is currently seen mainly in tumor tissues but can be detected by extracting RBPs from blood, urine, and other secretion samples for combined diagnostic testing. Finally, in the process of targeted cancer therapy, how to reduce or avoid the delivery of RBPs-targeted drugs to tumor tissue by the patient’s autoimmune response and how to better ensure the effectiveness and effective amount of drugs are also directions we need to work on. In the future, as our comprehension of RBP-mediated cancer regulatory networks advances, the integration of numerous data resources will encounter substantial challenges. Consequently, the development of comprehensive analysis software pertaining to this domain assumes utmost importance.

Author Contributions

Conception, writing—original draft preparation, Y.F.; writing—review and editing, S.Z., T.L., G.Z. and B.S.; supervision, J.L., B.L., C.J., Q.F., P.W. and D.W. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Natural Science Foundation of China (82074399).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

RBPsRNA-binding proteins
RBDsRNA-binding domains
ncRNAsnon-coding RNAs
lncRNAlong non-coding RNA
miRNAsmicroRNAs
hnRNPKheterogeneous nuclear ribonucleoprotein K
lncRNAslong-stranded non-coding RNAs
dsRBDdouble-stranded RNA binding structures
dsRNAdouble-stranded RNA
siRNAssmall interfering RNAs
snRNPssmall nuclear proteins
snoRNAsmall nucleolar RNA
snRNAsmall nucleolar RNA
TTPTritetraproline
AUF1AU-rich elemental RNA-binding protein 1
HuRHuman antigen R
IGF2BPInsulin-like growth factor 2 mRNA-binding protein
UTRsUntranslated regions
CRCColorectal cancer
ESRP1Epithelial splicing regulatory protein 1
RBM38RNA-binding motif protein 38
m7G7-methylguanosine
AUBPARE-binding protein
MSI1Musashi RNA-binding protein 1
eIF4Eeukaryotic translation initiation factor 4E
RNPsRibonucleoprotein complexes
RRMRNA Recognition Motifs
KHK homology structures domain
ZnFsZinc finger structural domains
DsRBDsDouble-stranded RNA-binding structural domains
ASAlternative splicing
APAAlternative polyadenylation
CPACleavage and polyadenylation
CPECytoplasmic polyadenylation elements
APCAdenomatous polyposis protein
EMTEpithelial-mesenchymal transition
3′UTR3′untranslated regions
5′UTR5′untranslated regions
ASOsAntisense oligonucleotides

References

  1. Lin, S.; Gregory, R.I. MicroRNA biogenesis pathways in cancer. Nat. Rev. Cancer 2015, 15, 321–333. [Google Scholar] [CrossRef]
  2. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [Green Version]
  3. Sahu, A.; Singhal, U.; Chinnaiyan, A.M. Long noncoding RNAs in cancer: From function to translation. Trends Cancer 2015, 1, 93–109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Ma, X.M.; Blenis, J. Molecular mechanisms of mTOR-mediated translational control. Nat. Rev. Mol. Cell Biol. 2009, 10, 307–318. [Google Scholar] [CrossRef] [PubMed]
  5. Van Nostrand, E.L.; Freese, P.; Pratt, G.A.; Wang, X.; Wei, X.; Xiao, R.; Blue, S.M.; Chen, J.Y.; Cody, N.A.L.; Dominguez, D.; et al. Author Correction: A large-scale binding and functional map of human RNA-binding proteins. Nature 2021, 589, E5. [Google Scholar] [CrossRef] [PubMed]
  6. Yao, R.W.; Wang, Y.; Chen, L.L. Cellular functions of long noncoding RNAs. Nat. Cell Biol. 2019, 21, 542–551. [Google Scholar] [CrossRef]
  7. Mitchell, S.F.; Parker, R. Principles and properties of eukaryotic mRNPs. Mol. Cell 2014, 54, 547–558. [Google Scholar] [CrossRef] [Green Version]
  8. Anantharaman, V.; Koonin, E.V.; Aravind, L. Comparative genomics and evolution of proteins involved in RNA metabolism. Nucleic Acids Res. 2002, 30, 1427–1464. [Google Scholar] [CrossRef]
  9. Matia-González, A.M.; Laing, E.E.; Gerber, A.P. Conserved mRNA-binding proteomes in eukaryotic organisms. Nat. Struct. Mol. Biol. 2015, 22, 1027–1033. [Google Scholar] [CrossRef] [Green Version]
  10. Xiao, R.; Chen, J.-Y.; Liang, Z.; Luo, D.; Chen, G.; Lu, Z.J.; Chen, Y.; Zhou, B.; Li, H.; Du, X.; et al. Pervasive Chromatin-RNA Binding Protein Interactions Enable RNA-Based Regulation of Transcription. Cell 2019, 178, 107–121. [Google Scholar] [CrossRef]
  11. David, C.J.; Manley, J.L. Alternative pre-mRNA splicing regulation in cancer: Pathways and programs unhinged. Genes Dev. 2010, 24, 2343–2364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Masamha, C.P.; Xia, Z.; Yang, J.; Albrecht, T.R.; Li, M.; Shyu, A.-B.; Li, W.; Wagner, E.J. CFIm25 links alternative polyadenylation to glioblastoma tumour suppression. Nature 2014, 510, 412–416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Guo, X.; Hartley, R.S. HuR contributes to cyclin E1 deregulation in MCF-7 breast cancer cells. Cancer Res. 2006, 66, 7948–7956. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Hüttelmaier, S.; Zenklusen, D.; Lederer, M.; Dictenberg, J.; Lorenz, M.; Meng, X.; Bassell, G.J.; Condeelis, J.; Singer, R.H. Spatial regulation of beta-actin translation by Src-dependent phosphorylation of ZBP1. Nature 2005, 438, 512–515. [Google Scholar] [CrossRef]
  15. Gerstberger, S.; Hafner, M.; Tuschl, T. A census of human RNA-binding proteins. Nat. Rev. Genet. 2014, 15, 829–845. [Google Scholar] [CrossRef]
  16. Neelamraju, Y.; Hashemikhabir, S.; Janga, S.C. The human RBPome: From genes and proteins to human disease. J. Proteom. 2015, 127, 61–70. [Google Scholar] [CrossRef] [Green Version]
  17. Kang, D.; Lee, Y.; Lee, J.S. RNA-Binding Proteins in Cancer: Functional and Therapeutic Perspectives. Cancers 2020, 12, 2699. [Google Scholar] [CrossRef]
  18. Corley, M.; Burns, M.C.; Yeo, G.W. How RNA-Binding Proteins Interact with RNA: Molecules and Mechanisms. Mol. Cell 2020, 78, 9–29. [Google Scholar] [CrossRef]
  19. Li, W.; Deng, X.; Chen, J. RNA-binding proteins in regulating mRNA stability and translation: Roles and mechanisms in cancer. Semin. Cancer Biol. 2022, 86, 664–677. [Google Scholar] [CrossRef]
  20. Liao, J.Y.; Yang, B.; Zhang, Y.C.; Wang, X.J.; Ye, Y.; Peng, J.W.; Yang, Z.Z.; He, J.H.; Zhang, Y.; Hu, K.; et al. EuRBPDB: A comprehensive resource for annotation, functional and oncological investigation of eukaryotic RNA binding proteins (RBPs). Nucleic Acids Res. 2020, 48, D307–D313. [Google Scholar] [CrossRef] [Green Version]
  21. Hentze, M.W.; Castello, A.; Schwarzl, T.; Preiss, T. A brave new world of RNA-binding proteins. Nat. Rev. Mol. Cell Biol. 2018, 19, 327–341. [Google Scholar] [CrossRef] [PubMed]
  22. Qin, H.; Ni, H.; Liu, Y.; Yuan, Y.; Xi, T.; Li, X.; Zheng, L. RNA-binding proteins in tumor progression. J. Hematol. Oncol. 2020, 13, 90. [Google Scholar] [CrossRef]
  23. Lunde, B.M.; Moore, C.; Varani, G. RNA-binding proteins: Modular design for efficient function. Nat. Rev. Mol. Cell Biol. 2007, 8, 479–490. [Google Scholar] [CrossRef] [Green Version]
  24. Auweter, S.D.; Oberstrass, F.C.; Allain, F.H.T. Sequence-specific binding of single-stranded RNA: Is there a code for recognition? Nucleic Acids Res. 2006, 34, 4943–4959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Hall, T.M.T. Multiple modes of RNA recognition by zinc finger proteins. Curr. Opin. Struct. Biol. 2005, 15, 367–373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Maris, C.; Dominguez, C.; Allain, F.H.T. The RNA recognition motif, a plastic RNA-binding platform to regulate post-transcriptional gene expression. FEBS J. 2005, 272, 2118–2131. [Google Scholar] [CrossRef]
  27. Mazza, C.; Segref, A.; Mattaj, I.W.; Cusack, S. Large-scale induced fit recognition of an m(7)GpppG cap analogue by the human nuclear cap-binding complex. EMBO J. 2002, 21, 5548–5557. [Google Scholar] [CrossRef] [Green Version]
  28. Oberstrass, F.C.; Auweter, S.D.; Erat, M.; Hargous, Y.; Henning, A.; Wenter, P.; Reymond, L.; Amir-Ahmady, B.; Pitsch, S.; Black, D.L.; et al. Structure of PTB bound to RNA: Specific binding and implications for splicing regulation. Science 2005, 309, 2054–2057. [Google Scholar] [CrossRef]
  29. Bono, F.; Ebert, J.; Lorentzen, E.; Conti, E. The crystal structure of the exon junction complex reveals how it maintains a stable grip on mRNA. Cell 2006, 126, 713–725. [Google Scholar] [CrossRef] [Green Version]
  30. Bono, F.; Ebert, J.; Unterholzner, L.; Güttler, T.; Izaurralde, E.; Conti, E. Molecular insights into the interaction of PYM with the Mago-Y14 core of the exon junction complex. EMBO Rep. 2004, 5, 304–310. [Google Scholar] [CrossRef] [Green Version]
  31. Fribourg, S.; Gatfield, D.; Izaurralde, E.; Conti, E. A novel mode of RBD-protein recognition in the Y14-Mago complex. Nat. Struct. Biol. 2003, 10, 433–439. [Google Scholar] [CrossRef] [PubMed]
  32. Kadlec, J.; Izaurralde, E.; Cusack, S. The structural basis for the interaction between nonsense-mediated mRNA decay factors UPF2 and UPF3. Nat. Struct. Mol. Biol. 2004, 11, 330–337. [Google Scholar] [CrossRef] [PubMed]
  33. Kielkopf, C.L.; Rodionova, N.A.; Green, M.R.; Burley, S.K. A novel peptide recognition mode revealed by the X-ray structure of a core U2AF35/U2AF65 heterodimer. Cell 2001, 106, 595–605. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Selenko, P.; Gregorovic, G.; Sprangers, R.; Stier, G.; Rhani, Z.; Krämer, A.; Sattler, M. Structural basis for the molecular recognition between human splicing factors U2AF65 and SF1/mBBP. Mol. Cell 2003, 11, 965–976. [Google Scholar] [CrossRef] [PubMed]
  35. Nicastro, G.; Taylor, I.A.; Ramos, A. KH-RNA interactions: Back in the groove. Curr. Opin. Struct. Biol. 2015, 30, 63–70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Grishin, N.V. KH domain: One motif, two folds. Nucleic Acids Res. 2001, 29, 638–643. [Google Scholar] [CrossRef]
  37. Valverde, R.; Edwards, L.; Regan, L. Structure and function of KH domains. FEBS J. 2008, 275, 2712–2726. [Google Scholar] [CrossRef]
  38. Masliah, G.; Barraud, P.; Allain, F.H.T. RNA recognition by double-stranded RNA binding domains: A matter of shape and sequence. Cell. Mol. Life Sci. 2013, 70, 1875–1895. [Google Scholar] [CrossRef] [Green Version]
  39. Ramos, A.; Grunert, S.; Adams, J.; Micklem, D.R.; Proctor, M.R.; Freund, S.; Bycroft, M.; St Johnston, D.; Varani, G. RNA recognition by a Staufen double-stranded RNA-binding domain. EMBO J. 2000, 19, 997–1009. [Google Scholar] [CrossRef] [Green Version]
  40. Stefl, R.; Xu, M.; Skrisovska, L.; Emeson, R.B.; Allain, F.H. Structure and specific RNA binding of ADAR2 double-stranded RNA binding motifs. Structure 2006, 14, 345–355. [Google Scholar] [CrossRef] [Green Version]
  41. Stephens, O.M.; Haudenschild, B.L.; Beal, P.A. The binding selectivity of ADAR2′s dsRBMs contributes to RNA-editing selectivity. Chem. Biol. 2004, 11, 1239–1250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Xu, M.; Wells, K.S.; Emeson, R.B. Substrate-dependent contribution of double-stranded RNA-binding motifs to ADAR2 function. Mol. Biol. Cell. 2006, 17, 3211–3220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Stefl, R.; Oberstrass, F.C.; Hood, J.L.; Jourdan, M.; Zimmermann, M.; Skrisovska, L.; Maris, C.; Peng, L.; Hofr, C.; Emeson, R.B.; et al. The solution structure of the ADAR2 dsRBM-RNA complex reveals a sequence-specific readout of the minor groove. Cell 2010, 143, 225–237. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, Z.H.; Hartman, E.; Roy, K.; Chanfreau, G.; Feigon, J. Structure of a Yeast RNase III dsRBD Complex with a Noncanonical RNA Substrate Provides New Insights into Binding Specificity of dsRBDs. Structure 2011, 19, 999–1010. [Google Scholar] [CrossRef] [Green Version]
  45. Picard, B.; Wegnez, M. Isolation of a 7S particle from Xenopus laevis oocytes: A 5S RNA-protein complex. Proc. Natl. Acad. Sci. USA 1979, 76, 241–245. [Google Scholar] [CrossRef]
  46. Ngwa, C.J.; Farrukh, A.; Pradel, G. Zinc finger proteins of Plasmodium falciparum. Cell Microbiol. 2021, 23, e13387. [Google Scholar] [CrossRef]
  47. Lai, W.S.; Carballo, E.; Thorn, J.M.; Kennington, E.A.; Blackshear, P.J. Interactions of CCCH zinc finger proteins with mRNA. Binding of tristetraprolin-related zinc finger proteins to Au-rich elements and destabilization of mRNA. J. Biol. Chem. 2000, 275, 17827–17837. [Google Scholar] [CrossRef] [Green Version]
  48. Wolfe, S.A.; Nekludova, L.; Pabo, C.O. DNA recognition by Cys2His2 zinc finger proteins. Annu. Rev. Biophys. Biomol. Struct. 2000, 29, 183–212. [Google Scholar] [CrossRef]
  49. Biamonti, G.; Catillo, M.; Pignataro, D.; Montecucco, A.; Ghigna, C. The alternative splicing side of cancer. Semin. Cell Dev. Biol. 2014, 32, 30–36. [Google Scholar] [CrossRef]
  50. Long, J.C.; Caceres, J.F. The SR protein family of splicing factors: Master regulators of gene expression. Biochem. J. 2009, 417, 15–27. [Google Scholar] [CrossRef] [Green Version]
  51. Zhu, J.; Mayeda, A.; Krainer, A.R. Exon identity established through differential antagonism between exonic splicing silencer-bound hnRNP A1 and enhancer-bound SR proteins. Mol. Cell 2001, 8, 1351–1361. [Google Scholar] [CrossRef] [PubMed]
  52. Erson-Bensan, A.E.; Can, T. Alternative Polyadenylation: Another Foe in Cancer. Mol. Cancer Res. 2016, 14, 507–517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Blake, D.; Lynch, K.W. The three as: Alternative splicing, alternative polyadenylation and their impact on apoptosis in immune function. Immunol. Rev. 2021, 304, 30–50. [Google Scholar] [CrossRef]
  54. Garneau, N.L.; Wilusz, J.; Wilusz, C.J. The highways and byways of mRNA decay. Nat. Rev. Mol. Cell Biol. 2007, 8, 113–126. [Google Scholar] [CrossRef] [PubMed]
  55. Sidali, A.; Teotia, V.; Solaiman, N.S.; Bashir, N.; Kanagaraj, R.; Murphy, J.J.; Surendranath, K. AU-Rich Element RNA Binding Proteins: At the Crossroads of Post-Transcriptional Regulation and Genome Integrity. Int. J. Mol. Sci. 2021, 23, 96. [Google Scholar] [CrossRef]
  56. Perron, G.; Jandaghi, P.; Solanki, S.; Safisamghabadi, M.; Storoz, C.; Karimzadeh, M.; Papadakis, A.I.; Arseneault, M.; Scelo, G.; Banks, R.E.; et al. A General Framework for Interrogation of mRNA Stability Programs Identifies RNA-Binding Proteins that Govern Cancer Transcriptomes. Cell Rep. 2018, 23, 1639–1650. [Google Scholar] [CrossRef] [Green Version]
  57. Truitt, M.L.; Ruggero, D. New frontiers in translational control of the cancer genome. Nat. Rev. Cancer. 2016, 16, 288–304. [Google Scholar] [CrossRef] [Green Version]
  58. Durie, D.; Lewis, S.M.; Liwak, U.; Kisilewicz, M.; Gorospe, M.; Holcik, M. RNA-binding protein HuR mediates cytoprotection through stimulation of XIAP translation. Oncogene 2011, 30, 1460–1469. [Google Scholar] [CrossRef] [Green Version]
  59. Hussey, G.S.; Chaudhury, A.; Dawson, A.E.; Lindner, D.J.; Knudsen, C.R.; Wilce, M.C.J.; Merrick, W.C.; Howe, P.H. Identification of an mRNP complex regulating tumorigenesis at the translational elongation step. Mol. Cell 2011, 41, 419–431. [Google Scholar] [CrossRef] [Green Version]
  60. Biyanee, A.; Ohnheiser, J.; Singh, P.; Klempnauer, K.H. A novel mechanism for the control of translation of specific mRNAs by tumor suppressor protein Pdcd4: Inhibition of translation elongation. Oncogene 2015, 34, 1384–1392. [Google Scholar] [CrossRef]
  61. Chaudhury, A.; Hussey, G.S.; Ray, P.S.; Jin, G.; Fox, P.L.; Howe, P.H. TGF-beta-mediated phosphorylation of hnRNP E1 induces EMT via transcript-selective translational induction of Dab2 and ILEI. Nat. Cell Biol. 2010, 12, 286–293. [Google Scholar] [CrossRef] [Green Version]
  62. Zhang, M.; Weng, W.; Zhang, Q.; Wu, Y.; Ni, S.; Tan, C.; Xu, M.; Sun, H.; Liu, C.; Wei, P.; et al. The lncRNA NEAT1 activates Wnt/β-catenin signaling and promotes colorectal cancer progression via interacting with DDX5. J. Hematol. Oncol. 2018, 11, 113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Ha, M.; Kim, V.N. Regulation of microRNA biogenesis. Nat. Rev. Mol. Cell Biol. 2014, 15, 509–524. [Google Scholar] [CrossRef] [PubMed]
  64. Melo, S.A.; Ropero, S.; Moutinho, C.; Aaltonen, L.A.; Yamamoto, H.; Calin, G.A.; Rossi, S.; Fernandez, A.F.; Carneiro, F.; Oliveira, C.; et al. A TARBP2 mutation in human cancer impairs microRNA processing and DICER1 function. Nat. Genet. 2009, 41, 365–370. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Quattrone, A.; Dassi, E. The Architecture of the Human RNA-Binding Protein Regulatory Network. iScience 2019, 21, 706–719. [Google Scholar] [CrossRef] [Green Version]
  66. Keene, J.D. RNA regulons: Coordination of post-transcriptional events. Nat. Rev. Genet. 2007, 8, 533–543. [Google Scholar] [CrossRef]
  67. Lin, Y.; Li, Z.; Ozsolak, F.; Kim, S.W.; Arango-Argoty, G.; Liu, T.T.; Tenenbaum, S.A.; Bailey, T.; Monaghan, A.P.; Milos, P.M.; et al. An in-depth map of polyadenylation sites in cancer. Nucleic Acids Res. 2012, 40, 8460–8471. [Google Scholar] [CrossRef] [Green Version]
  68. Bonnal, S.C.; Lopez-Oreja, I.; Valcarcel, J. Roles and mechanisms of alternative splicing in cancer—Implications for care. Nat. Rev. Clin. Oncol. 2020, 17, 457–474. [Google Scholar] [CrossRef]
  69. Gallego-Paez, L.M.; Bordone, M.C.; Leote, A.C.; Saraiva-Agostinho, N.; Ascensao-Ferreira, M.; Barbosa-Morais, N.L. Alternative splicing: The pledge, the turn, and the prestige: The key role of alternative splicing in human biological systems. Hum. Genet. 2017, 136, 1015–1042. [Google Scholar] [CrossRef] [Green Version]
  70. Baralle, F.E.; Giudice, J. Alternative splicing as a regulator of development and tissue identity. Nat. Rev. Mol. Cell Bio. 2017, 18, 437–451. [Google Scholar] [CrossRef]
  71. Climente-Gonzalez, H.; Porta-Pardo, E.; Godzik, A.; Eyras, E. The Functional Impact of Alternative Splicing in Cancer. Cell Rep. 2017, 20, 2215–2226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Danan-Gotthold, M.; Golan-Gerstl, R.; Eisenberg, E.; Meir, K.; Karni, R.; Levanon, E.Y. Identification of recurrent regulated alternative splicing events across human solid tumors. Nucleic Acids Res. 2015, 43, 5130–5144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Kahles, A.; Lehmann, K.-V.; Toussaint, N.C.; Hüser, M.; Stark, S.G.; Sachsenberg, T.; Stegle, O.; Kohlbacher, O.; Sander, C.; Rätsch, G. Comprehensive Analysis of Alternative Splicing Across Tumors from 8705 Patients. Cancer Cell 2018, 34. [Google Scholar] [CrossRef] [Green Version]
  74. Nilsen, T.W.; Graveley, B.R. Expansion of the eukaryotic proteome by alternative splicing. Nature 2010, 463, 457–463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Eléouët, M.; Lu, C.; Zhou, Y.; Yang, P.; Ma, J.; Xu, G. Insights on the biological functions and diverse regulation of RNA-binding protein 39 and their implication in human diseases. Biochim. Biophys. Acta Gene. Regul. Mech. 2023, 1866, 194902. [Google Scholar] [CrossRef]
  76. Dvinge, H.; Kim, E.; Abdel-Wahab, O.; Bradley, R.K. RNA splicing factors as oncoproteins and tumour suppressors. Nat. Rev. Cancer 2016, 16, 413–430. [Google Scholar] [CrossRef] [PubMed]
  77. Jimenez, M.; Arechederra, M.; Avila, M.A.; Berasain, C. Splicing alterations contributing to cancer hallmarks in the liver: Central role of dedifferentiation and genome instability. Transl. Gastroenterol. Hepatol. 2018, 3, 84. [Google Scholar] [CrossRef]
  78. Li, X.; Qian, X.; Peng, L.X.; Jiang, Y.; Hawke, D.H.; Zheng, Y.; Xia, Y.; Lee, J.H.; Cote, G.; Wang, H.; et al. Corrigendum: A splicing switch from ketohexokinase-C to ketohexokinase-A drives hepatocellular carcinoma formation. Nat. Cell Biol. 2016, 18, 709. [Google Scholar] [CrossRef] [Green Version]
  79. Elkon, R.; Ugalde, A.P.; Agami, R. Alternative cleavage and polyadenylation: Extent, regulation and function. Nat. Rev. Genet. 2013, 14, 496–506. [Google Scholar] [CrossRef]
  80. Xiang, Y.; Ye, Y.; Lou, Y.; Yang, Y.; Cai, C.; Zhang, Z.; Mills, T.; Chen, N.Y.; Kim, Y.; Muge Ozguc, F.; et al. Comprehensive Characterization of Alternative Polyadenylation in Human Cancer. J. Natl. Cancer. Inst. 2018, 110, 379–389. [Google Scholar] [CrossRef]
  81. Pereira, B.; Billaud, M.; Almeida, R. RNA-Binding Proteins in Cancer: Old Players and New Actors. Trends Cancer 2017, 3, 506–528. [Google Scholar] [CrossRef] [PubMed]
  82. Tian, B.; Manley, J.L. Alternative polyadenylation of mRNA precursors. Nat. Rev. Mol. Cell Biol. 2017, 18, 18–30. [Google Scholar] [CrossRef] [PubMed]
  83. Fernández-Miranda, G.; Méndez, R. The CPEB-family of proteins, translational control in senescence and cancer. Ageing Res. Rev. 2012, 11, 460–472. [Google Scholar] [CrossRef] [PubMed]
  84. Bava, F.-A.; Eliscovich, C.; Ferreira, P.G.; Miñana, B.; Ben-Dov, C.; Guigó, R.; Valcárcel, J.; Méndez, R. CPEB1 coordinates alternative 3′-UTR formation with translational regulation. Nature 2013, 495, 121–125. [Google Scholar] [CrossRef]
  85. Nagaoka, K.; Fujii, K.; Zhang, H.; Usuda, K.; Watanabe, G.; Ivshina, M.; Richter, J.D. CPEB1 mediates epithelial-to-mesenchyme transition and breast cancer metastasis. Oncogene 2016, 35, 2893–2901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Hasan, A.; Cotobal, C.; Duncan, C.D.S.; Mata, J. Systematic Analysis of the Role of RNA-Binding Proteins in the Regulation of RNA Stability. PLoS Genet. 2014, 10, e1004684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Stoecklin, G.; Anderson, P. Posttranscriptional mechanisms regulating the inflammatory response. Adv. Immunol. 2006, 89, 1–37. [Google Scholar] [CrossRef]
  88. Gruber, A.R.; Fallmann, J.; Kratochvill, F.; Kovarik, P.; Hofacker, I.L. AREsite: A database for the comprehensive investigation of AU-rich elements. Nucleic Acids Res. 2011, 39, D66–D69. [Google Scholar] [CrossRef] [Green Version]
  89. Fallmann, J.; Sedlyarov, V.; Tanzer, A.; Kovarik, P.; Hofacker, I.L. AREsite2: An enhanced database for the comprehensive investigation of AU/GU/U-rich elements. Nucleic Acids Res. 2016, 44, D90–D95. [Google Scholar] [CrossRef] [Green Version]
  90. Hitti, E.; Bakheet, T.; Al-Souhibani, N.; Moghrabi, W.; Al-Yahya, S.; Al-Ghamdi, M.; Al-Saif, M.; Shoukri, M.M.; Lanczky, A.; Grepin, R.; et al. Systematic Analysis of AU-Rich Element Expression in Cancer Reveals Common Functional Clusters Regulated by Key RNA-Binding Proteins. Cancer Res. 2016, 76, 4068–4080. [Google Scholar] [CrossRef] [Green Version]
  91. Khabar, K.S. Hallmarks of cancer and AU-rich elements. Wiley Interdiscip Rev. RNA 2017, 8, e1368. [Google Scholar] [CrossRef] [Green Version]
  92. Moore, A.E.; Chenette, D.M.; Larkin, L.C.; Schneider, R.J. Physiological networks and disease functions of RNA-binding protein AUF1. Wires RNA 2014, 5, 549–564. [Google Scholar] [CrossRef] [PubMed]
  93. Zhu, H.F.; Berkova, Z.; Mathur, R.; Sehgal, L.; Khashab, T.; Tao, R.H.; Ao, X.; Feng, L.; Sabichi, A.L.; Blechacz, B.; et al. HuR Suppresses Fas Expression and Correlates with Patient Outcome in Liver Cancer. Mol. Cancer Res. 2015, 13, 809–818. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Hu, Y.P.; Jin, Y.P.; Wu, X.S.; Yang, Y.; Li, Y.S.; Li, H.F.; Xiang, S.S.; Song, X.L.; Jiang, L.; Zhang, Y.J.; et al. LncRNA-HGBC stabilized by HuR promotes gallbladder cancer progression by regulating miR-502-3p/SET/AKT axis. Mol. Cancer 2021, 20, 167. [Google Scholar] [CrossRef]
  95. Jonas, K.; Calin, G.A.; Pichler, M. RNA-Binding Proteins as Important Regulators of Long Non-Coding RNAs in Cancer. Int. J. Mol. Sci. 2020, 21, 2969. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Filippova, N.; Yang, X.; Wang, Y.; Gillespie, G.Y.; Langford, C.; King, P.H.; Wheeler, C.; Nabors, L.B. The RNA-binding protein HuR promotes glioma growth and treatment resistance. Mol. Cancer Res. 2011, 9, 648–659. [Google Scholar] [CrossRef] [Green Version]
  97. Hou, P.; Li, L.; Chen, F.; Chen, Y.; Liu, H.; Li, J.; Bai, J.; Zheng, J. PTBP3-Mediated Regulation of ZEB1 mRNA Stability Promotes Epithelial-Mesenchymal Transition in Breast Cancer. Cancer Res. 2018, 78, 387–398. [Google Scholar] [CrossRef] [Green Version]
  98. Stöhr, N.; Köhn, M.; Lederer, M.; Glass, M.; Reinke, C.; Singer, R.H.; Hüttelmaier, S. IGF2BP1 promotes cell migration by regulating MK5 and PTEN signaling. Genes Dev. 2012, 26, 176–189. [Google Scholar] [CrossRef] [Green Version]
  99. Gutschner, T.; Hämmerle, M.; Pazaitis, N.; Bley, N.; Fiskin, E.; Uckelmann, H.; Heim, A.; Groβ, M.; Hofmann, N.; Geffers, R.; et al. Insulin-like growth factor 2 mRNA-binding protein 1 (IGF2BP1) is an important protumorigenic factor in hepatocellular carcinoma. Hepatology 2014, 59, 1900–1911. [Google Scholar] [CrossRef]
  100. Hämmerle, M.; Gutschner, T.; Uckelmann, H.; Ozgur, S.; Fiskin, E.; Gross, M.; Skawran, B.; Geffers, R.; Longerich, T.; Breuhahn, K.; et al. Posttranscriptional destabilization of the liver-specific long noncoding RNA HULC by the IGF2 mRNA-binding protein 1 (IGF2BP1). Hepatology 2013, 58, 1703–1712. [Google Scholar] [CrossRef]
  101. Mizutani, R.; Imamachi, N.; Suzuki, Y.; Yoshida, H.; Tochigi, N.; Oonishi, T.; Suzuki, Y.; Akimitsu, N. Oncofetal protein IGF2BP3 facilitates the activity of proto-oncogene protein eIF4E through the destabilization of EIF4E-BP2 mRNA. Oncogene 2016, 35, 3495–3502. [Google Scholar] [CrossRef] [PubMed]
  102. Carlevaro-Fita, J.; Johnson, R. Global Positioning System: Understanding Long Noncoding RNAs through Subcellular Localization. Mol. Cell 2019, 73, 869–883. [Google Scholar] [CrossRef] [Green Version]
  103. Degrauwe, N.; Suvà, M.-L.; Janiszewska, M.; Riggi, N.; Stamenkovic, I. IMPs: An RNA-binding protein family that provides a link between stem cell maintenance in normal development and cancer. Genes Dev. 2016, 30, 2459–2474. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Medioni, C.; Mowry, K.; Besse, F. Principles and roles of mRNA localization in animal development. Development 2012, 139, 3263–3276. [Google Scholar] [CrossRef] [Green Version]
  105. Nagaoka, K.; Udagawa, T.; Richter, J.D. CPEB-mediated ZO-1 mRNA localization is required for epithelial tight-junction assembly and cell polarity. Nat. Commun. 2012, 3, 675. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Hamdollah Zadeh, M.A.; Amin, E.M.; Hoareau-Aveilla, C.; Domingo, E.; Symonds, K.E.; Ye, X.; Heesom, K.J.; Salmon, A.; D’Silva, O.; Betteridge, K.B.; et al. Alternative splicing of TIA-1 in human colon cancer regulates VEGF isoform expression, angiogenesis, tumour growth and bevacizumab resistance. Mol. Oncol. 2015, 9, 167–178. [Google Scholar] [CrossRef]
  107. Liu, Y.; Liu, R.; Yang, F.; Cheng, R.; Chen, X.; Cui, S.; Gu, Y.; Sun, W.; You, C.; Liu, Z.; et al. miR-19a promotes colorectal cancer proliferation and migration by targeting TIA1. Mol. Cancer 2017, 16, 53. [Google Scholar] [CrossRef] [Green Version]
  108. Díaz-Muñoz, M.D.; Kiselev, V.Y.; Le Novère, N.; Curk, T.; Ule, J.; Turner, M. Tia1 dependent regulation of mRNA subcellular location and translation controls p53 expression in B cells. Nat. Commun. 2017, 8, 530. [Google Scholar] [CrossRef] [Green Version]
  109. Farina, K.L.; Huttelmaier, S.; Musunuru, K.; Darnell, R.; Singer, R.H. Two ZBP1 KH domains facilitate beta-actin mRNA localization, granule formation, and cytoskeletal attachment. J. Cell Biol. 2003, 160, 77–87. [Google Scholar] [CrossRef]
  110. Gu, W.; Katz, Z.; Wu, B.; Park, H.Y.; Li, D.; Lin, S.; Wells, A.L.; Singer, R.H. Regulation of local expression of cell adhesion and motility-related mRNAs in breast cancer cells by IMP1/ZBP1. J. Cell Sci. 2012, 125, 81–91. [Google Scholar] [CrossRef] [Green Version]
  111. Gu, W.; Pan, F.; Singer, R.H. Blocking beta-catenin binding to the ZBP1 promoter represses ZBP1 expression, leading to increased proliferation and migration of metastatic breast-cancer cells. J. Cell Sci. 2009, 122, 1895–1905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Hamilton, K.E.; Noubissi, F.K.; Katti, P.S.; Hahn, C.M.; Davey, S.R.; Lundsmith, E.T.; Klein-Szanto, A.J.; Rhim, A.D.; Spiegelman, V.S.; Rustgi, A.K. IMP1 promotes tumor growth, dissemination and a tumor-initiating cell phenotype in colorectal cancer cell xenografts. Carcinogenesis 2013, 34, 2647–2654. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Hamilton, K.E.; Chatterji, P.; Lundsmith, E.T.; Andres, S.F.; Giroux, V.; Hicks, P.D.; Noubissi, F.K.; Spiegelman, V.S.; Rustgi, A.K. Loss of Stromal IMP1 Promotes a Tumorigenic Microenvironment in the Colon. Mol. Cancer Res. 2015, 13, 1478–1486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Rebucci, M.; Sermeus, A.; Leonard, E.; Delaive, E.; Dieu, M.; Fransolet, M.; Arnould, T.; Michiels, C. miRNA-196b inhibits cell proliferation and induces apoptosis in HepG2 cells by targeting IGF2BP1. Mol. Cancer 2015, 14, 79. [Google Scholar] [CrossRef] [Green Version]
  115. Hinnebusch, A.G. The Scanning Mechanism of Eukaryotic Translation Initiation. Annu. Rev. Biochem. 2014, 83, 779–812. [Google Scholar] [CrossRef] [PubMed]
  116. Richter, J.D.; Coller, J. Pausing on Polyribosomes: Make Way for Elongation in Translational Control. Cell 2015, 163, 292–300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Truitt, M.L.; Ruggero, D. New frontiers in translational control of the cancer genome. Nat. Rev. Cancer 2017, 17, 332. [Google Scholar] [CrossRef] [Green Version]
  118. Janin, M.; Coll-SanMartin, L.; Esteller, M. Disruption of the RNA modifications that target the ribosome translation machinery in human cancer. Mol. Cancer 2020, 19, 7010. [Google Scholar] [CrossRef] [Green Version]
  119. Hinnebusch, A.G.; Ivanov, I.P.; Sonenberg, N. Translational control by 5‘-untranslated regions of eukaryotic mRNAs. Science 2016, 352, 1413–1416. [Google Scholar] [CrossRef]
  120. Bhat, M.; Robichaud, N.; Hulea, L.; Sonenberg, N.; Pelletier, J.; Topisirovic, I. Targeting the translation machinery in cancer. Nat. Rev. Drug. Discov. 2015, 14, 261–278. [Google Scholar] [CrossRef]
  121. Li, L.; Yang, Y.; Wu, M.; Yu, Z.; Wang, C.; Dou, G.; He, H.; Wang, H.; Yang, N.; Qi, H.; et al. beta-Asarone Induces Apoptosis and Cell Cycle Arrest of Human Glioma U251 Cells via Suppression of HnRNP A2/B1-Mediated Pathway In Vitro and In Vivo. Molecules 2018, 23, 1072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Golan-Gerstl, R.; Cohen, M.; Shilo, A.; Suh, S.-S.; Bakàcs, A.; Coppola, L.; Karni, R. Splicing factor hnRNP A2/B1 regulates tumor suppressor gene splicing and is an oncogenic driver in glioblastoma. Cancer Res. 2011, 71, 4464–4472. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Pisera, A.; Campo, A.; Campo, S. Structure and functions of the translation initiation factor eIF4E and its role in cancer development and treatment. J. Genet. Genom. 2018, 45, 13–24. [Google Scholar] [CrossRef] [PubMed]
  124. Truitt, M.L.; Conn, C.S.; Shi, Z.; Pang, X.; Tokuyasu, T.; Coady, A.M.; Seo, Y.; Barna, M.; Ruggero, D. Differential Requirements for eIF4E Dose in Normal Development and Cancer. Cell 2015, 162, 59–71. [Google Scholar] [CrossRef] [Green Version]
  125. Mazan-Mamczarz, K.; Galbán, S.; López de Silanes, I.; Martindale, J.L.; Atasoy, U.; Keene, J.D.; Gorospe, M. RNA-binding protein HuR enhances p53 translation in response to ultraviolet light irradiation. Proc. Natl. Acad. Sci. USA 2003, 100, 8354–8359. [Google Scholar] [CrossRef] [PubMed]
  126. Vo, D.T.; Abdelmohsen, K.; Martindale, J.L.; Qiao, M.; Tominaga, K.; Burton, T.L.; Gelfond, J.A.L.; Brenner, A.J.; Patel, V.; Trageser, D.; et al. The oncogenic RNA-binding protein Musashi1 is regulated by HuR via mRNA translation and stability in glioblastoma cells. Mol. Cancer Res. 2012, 10, 143–155. [Google Scholar] [CrossRef] [Green Version]
  127. Galban, S.; Kuwano, Y.; Pullmann, R., Jr.; Martindale, J.L.; Kim, H.H.; Lal, A.; Abdelmohsen, K.; Yang, X.; Dang, Y.; Liu, J.O.; et al. RNA-binding proteins HuR and PTB promote the translation of hypoxia-inducible factor 1alpha. Mol. Cell Biol. 2008, 28, 93–107. [Google Scholar] [CrossRef] [Green Version]
  128. Glorian, V.; Maillot, G.; Poles, S.; Iacovoni, J.S.; Favre, G.; Vagner, S. HuR-dependent loading of miRNA RISC to the mRNA encoding the Ras-related small GTPase RhoB controls its translation during UV-induced apoptosis. Cell Death Differ. 2011, 18, 1692–1701. [Google Scholar] [CrossRef] [PubMed]
  129. Al-Haidari, A.; Algaber, A.; Madhi, R.; Syk, I.; Thorlacius, H. MiR-155-5p controls colon cancer cell migration via post-transcriptional regulation of Human Antigen R (HuR). Cancer Lett. 2018, 421, 145–151. [Google Scholar] [CrossRef] [PubMed]
  130. Shu, C.; Yan, D.; Mo, Y.; Gu, J.; Shah, N.; He, J. Long noncoding RNA lncARSR promotes epithelial ovarian cancer cell proliferation and invasion by association with HuR and miR-200 family. Am. J. Cancer Res. 2018, 8, 981–992. [Google Scholar]
  131. Wu, J.I.; Lin, Y.P.; Tseng, C.W.; Chen, H.J.; Wang, L.H. Crabp2 Promotes Metastasis of Lung Cancer Cells via HuR and Integrin beta1/FAK/ERK Signaling. Sci. Rep. 2019, 9, 845. [Google Scholar] [CrossRef] [Green Version]
  132. Yang, F.; Hu, A.; Li, D.; Wang, J.; Guo, Y.; Liu, Y.; Li, H.; Chen, Y.; Wang, X.; Huang, K.; et al. Circ-HuR suppresses HuR expression and gastric cancer progression by inhibiting CNBP transactivation. Mol. Cancer 2019, 18, 158. [Google Scholar] [CrossRef] [Green Version]
  133. Huang, W.; Wei, X.-L.; Ni, W.; Cao, M.; Meng, L.; Yang, H. The expression of RNA-binding protein RBM38 decreased in renal cell carcinoma and represses renal cancer cell proliferation, migration, and invasion. Tumour. Biol. 2017, 39, 1010428317701635. [Google Scholar] [CrossRef] [Green Version]
  134. Wu, J.; Zhou, X.-J.; Sun, X.; Xia, T.-S.; Li, X.-X.; Shi, L.; Zhu, L.; Zhou, W.-B.; Wei, J.-F.; Ding, Q. RBM38 is involved in TGF-β-induced epithelial-to-mesenchymal transition by stabilising zonula occludens-1 mRNA in breast cancer. Br. J. Cancer 2017, 117, 675–684. [Google Scholar] [CrossRef]
  135. Zheng, L.; Zhang, Z.; Zhang, S.; Guo, Q.; Zhang, F.; Gao, L.; Ni, H.; Guo, X.; Xiang, C.; Xi, T. RNA Binding Protein RNPC1 Inhibits Breast Cancer Cell Metastasis via Activating STARD13-Correlated ceRNA Network. Mol. Pharm. 2018, 15, 2123–2132. [Google Scholar] [CrossRef]
  136. Zhou, X.-J.; Wu, J.; Shi, L.; Li, X.-X.; Zhu, L.; Sun, X.; Qian, J.-Y.; Wang, Y.; Wei, J.-F.; Ding, Q. PTEN expression is upregulated by a RNA-binding protein RBM38 via enhancing its mRNA stability in breast cancer. J. Exp. Clin. Cancer Res. 2017, 36, 149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Hsieh, A.C.; Ruggero, D. Targeting eukaryotic translation initiation factor 4E (eIF4E) in cancer. Clin. Cancer Res. 2010, 16, 4914–4920. [Google Scholar] [CrossRef] [Green Version]
  138. Wendel, H.G.; De Stanchina, E.; Fridman, J.S.; Malina, A.; Ray, S.; Kogan, S.; Cordon-Cardo, C.; Pelletier, J.; Lowe, S.W. Survival signalling by Akt and eIF4E in oncogenesis and cancer therapy. Nature 2004, 428, 332–337. [Google Scholar] [CrossRef]
  139. Xu, T.; Zong, Y.; Peng, L.; Kong, S.; Zhou, M.; Zou, J.; Liu, J.; Miao, R.; Sun, X.; Li, L. Overexpression of eIF4E in colorectal cancer patients is associated with liver metastasis. OncoTargets Ther. 2016, 9, 815–822. [Google Scholar] [CrossRef] [Green Version]
  140. Lal, A.; Mazan-Mamczarz, K.; Kawai, T.; Yang, X.; Martindale, J.L.; Gorospe, M. Concurrent versus individual binding of HuR and AUF1 to common labile target mRNAs. EMBO J. 2004, 23, 3092–3102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Pont, A.R.; Sadri, N.; Hsiao, S.J.; Smith, S.; Schneider, R.J. mRNA decay factor AUF1 maintains normal aging, telomere maintenance, and suppression of senescence by activation of telomerase transcription. Mol. Cell 2012, 47, 5–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Yoon, J.H.; De, S.; Srikantan, S.; Abdelmohsen, K.; Grammatikakis, I.; Kim, J.; Kim, K.M.; Noh, J.H.; White, E.J.; Martindale, J.L.; et al. PAR-CLIP analysis uncovers AUF1 impact on target RNA fate and genome integrity. Nat. Commun. 2014, 5, 5248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Zucconi, B.E.; Wilson, G.M. Modulation of neoplastic gene regulatory pathways by the RNA-binding factor AUF1. Front. Biosci. 2011, 16, 2307–2325. [Google Scholar] [CrossRef] [Green Version]
  144. Chaudhury, A.; Chander, P.; Howe, P.H. Heterogeneous nuclear ribonucleoproteins (hnRNPs) in cellular processes: Focus on hnRNP E1′s multifunctional regulatory roles. RNA 2010, 16, 1449–1462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Clower, C.V.; Chatterjee, D.; Wang, Z.; Cantley, L.C.; Vander Heiden, M.G.; Krainer, A.R. The alternative splicing repressors hnRNP A1/A2 and PTB influence pyruvate kinase isoform expression and cell metabolism. Proc. Natl. Acad. Sci. USA 2010, 107, 1894–1899. [Google Scholar] [CrossRef] [PubMed]
  146. David, C.J.; Chen, M.; Assanah, M.; Canoll, P.; Manley, J.L. HnRNP proteins controlled by c-Myc deregulate pyruvate kinase mRNA splicing in cancer. Nature 2010, 463, 364–368. [Google Scholar] [CrossRef] [Green Version]
  147. Tauler, J.; Zudaire, E.; Liu, H.; Shih, J.; Mulshine, J.L. hnRNP A2/B1 modulates epithelial-mesenchymal transition in lung cancer cell lines. Cancer Res. 2010, 70, 7137–7147. [Google Scholar] [CrossRef] [Green Version]
  148. Moran-Jones, K.; Grindlay, J.; Jones, M.; Smith, R.; Norman, J.C. hnRNP A2 regulates alternative mRNA splicing of TP53INP2 to control invasive cell migration. Cancer Res. 2009, 69, 9219–9227. [Google Scholar] [CrossRef] [Green Version]
  149. Ding, J.; Wang, D.Z. Mystery of Trbp, tale of a RBP in the miRNA pathway. Cell Cycle 2015, 14, 3007–3008. [Google Scholar] [CrossRef] [Green Version]
  150. Melo, S.A.; Ropero, S.; Moutinho, C.; Aaltonen, L.A.; Yamamoto, H.; Calin, G.A.; Rossi, S.; Fernandez, A.F.; Carneiro, F.; Oliveira, C.; et al. Retraction: A TARBP2 mutation in human cancer impairs microRNA processing and DICER1 function. Nat. Genet. 2016, 48, 221. [Google Scholar] [CrossRef] [Green Version]
  151. Noubissi, F.K.; Elcheva, I.; Bhatia, N.; Shakoori, A.; Ougolkov, A.; Liu, J.; Minamoto, T.; Ross, J.; Fuchs, S.Y.; Spiegelman, V.S. CRD-BP mediates stabilization of betaTrCP1 and c-myc mRNA in response to beta-catenin signalling. Nature 2006, 441, 898–901. [Google Scholar] [CrossRef] [PubMed]
  152. Weidensdorfer, D.; Stöhr, N.; Baude, A.; Lederer, M.; Köhn, M.; Schierhorn, A.; Buchmeier, S.; Wahle, E.; Hüttelmaier, S. Control of c-myc mRNA stability by IGF2BP1-associated cytoplasmic RNPs. RNA 2009, 15, 104–115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Vikesaa, J.; Hansen, T.V.O.; Jønson, L.; Borup, R.; Wewer, U.M.; Christiansen, J.; Nielsen, F.C. RNA-binding IMPs promote cell adhesion and invadopodia formation. EMBO J. 2006, 25, 1456–1468. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Elcheva, I.; Goswami, S.; Noubissi, F.K.; Spiegelman, V.S. CRD-BP protects the coding region of betaTrCP1 mRNA from miR-183-mediated degradation. Mol. Cell 2009, 35, 240–246. [Google Scholar] [CrossRef] [Green Version]
  155. Huang, H.; Weng, H.; Sun, W.; Qin, X.; Shi, H.; Wu, H.; Zhao, B.S.; Mesquita, A.; Liu, C.; Yuan, C.L.; et al. Recognition of RNA N6-methyladenosine by IGF2BP proteins enhances mRNA stability and translation. Nat. Cell Biol. 2018, 20, 285–295. [Google Scholar] [CrossRef] [PubMed]
  156. Janiszewska, M.; Suvà, M.L.; Riggi, N.; Houtkooper, R.H.; Auwerx, J.; Clément-Schatlo, V.; Radovanovic, I.; Rheinbay, E.; Provero, P.; Stamenkovic, I. Imp2 controls oxidative phosphorylation and is crucial for preserving glioblastoma cancer stem cells. Genes Dev. 2012, 26, 1926–1944. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Ye, S.; Song, W.; Xu, X.; Zhao, X.; Yang, L. IGF2BP2 promotes colorectal cancer cell proliferation and survival through interfering with RAF-1 degradation by miR-195. FEBS Lett. 2016, 590, 1641–1650. [Google Scholar] [CrossRef] [Green Version]
  158. Kim, H.-Y.; Ha Thi, H.T.; Hong, S. IMP2 and IMP3 cooperate to promote the metastasis of triple-negative breast cancer through destabilization of progesterone receptor. Cancer Lett. 2018, 415, 30–39. [Google Scholar] [CrossRef]
  159. Mohibi, S.; Chen, X.B.; Zhang, J. Cancer the’RBP’eutics-RNA-binding proteins as therapeutic targets for cancer. Pharmacol. Ther. 2019, 203, 107390. [Google Scholar] [CrossRef]
  160. Julio, A.R.; Backus, K.M. New approaches to target RNA binding proteins. Curr. Opin. Chem. Biol. 2021, 62, 13–23. [Google Scholar] [CrossRef]
  161. Lang, M.; Berry, D.; Passecker, K.; Mesteri, I.; Bhuju, S.; Ebner, F.; Sedlyarov, V.; Evstatiev, R.; Dammann, K.; Loy, A.; et al. HuR Small-Molecule Inhibitor Elicits Differential Effects in Adenomatosis Polyposis and Colorectal Carcinogenesis. Cancer Res. 2017, 77, 2424–2438. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Lal, P.; Cerofolini, L.; D’Agostino, V.G.; Zucal, C.; Fuccio, C.; Bonomo, I.; Dassi, E.; Giuntini, S.; Di Maio, D.; Vishwakarma, V.; et al. Regulation of HuR structure and function by dihydrotanshinone-I. Nucleic Acids Res. 2017, 45, 9514–9527. [Google Scholar] [CrossRef] [PubMed]
  163. Morimatsu, K.; Aishima, S.; Yamamoto, H.; Hayashi, A.; Nakata, K.; Oda, Y.; Shindo, K.; Fujino, M.; Tanaka, M.; Oda, Y. Insulin-like growth factor II messenger RNA-binding protein-3 is a valuable diagnostic and prognostic marker of intraductal papillary mucinous neoplasm. Hum. Pathol. 2013, 44, 1714–1721. [Google Scholar] [CrossRef]
  164. Filippova, N.; Yang, X.H.; Ananthan, S.; Sorochinsky, A.; Hackney, J.R.; Gentry, Z.; Bae, S.; King, P.; Nabors, L.B. Hu antigen R (HuR) multimerization contributes to glioma disease progression. J. Biol. Chem. 2017, 292, 16999–17010. [Google Scholar] [CrossRef] [Green Version]
  165. Kaur, K.; Wu, X.Q.; Fields, J.K.; Johnson, D.K.; Lan, L.; Pratt, M.; Somoza, A.D.; Wang, C.C.C.; Karanicolas, J.; Oakley, B.R.; et al. The fungal natural product azaphilone-9 binds to HuR and inhibits HuR-RNA interaction in vitro. PLoS ONE 2017, 12, e0175471. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Allegri, L.; Baldan, F.; Roy, S.; Aube, J.; Russo, D.; Filetti, S.; Damante, G. The HuR CMLD-2 inhibitor exhibits antitumor effects via MAD2 downregulation in thyroid cancer cells. Sci. Rep. 2019, 9, 7374. [Google Scholar] [CrossRef] [Green Version]
  167. Wu, X.; Lan, L.; Wilson, D.M.; Marquez, R.T.; Tsao, W.-C.; Gao, P.; Roy, A.; Turner, B.A.; McDonald, P.; Tunge, J.A.; et al. Identification and validation of novel small molecule disruptors of HuR-mRNA interaction. ACS Chem. Biol. 2015, 10, 1476–1484. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Kentsis, A.; Topisirovic, I.; Culjkovic, B.; Shao, L.; Borden, K.L.B. Ribavirin suppresses eIF4E-mediated oncogenic transformation by physical mimicry of the 7-methyl guanosine mRNA cap. Proc. Natl. Acad. Sci. USA 2004, 101, 18105–18110. [Google Scholar] [CrossRef] [PubMed]
  169. Li, S.; Jia, Y.; Jacobson, B.; McCauley, J.; Kratzke, R.; Bitterman, P.B.; Wagner, C.R. Treatment of breast and lung cancer cells with a N-7 benzyl guanosine monophosphate tryptamine phosphoramidate pronucleotide (4Ei-1) results in chemosensitization to gemcitabine and induced eIF4E proteasomal degradation. Mol. Pharm. 2013, 10, 523–531. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Troschel, F.M.; Minte, A.; Ismail, Y.M.; Kamal, A.; Abdullah, M.S.; Ahmed, S.H.; Deffner, M.; Kemper, B.; Kiesel, L.; Eich, H.T.; et al. Knockdown of Musashi RNA Binding Proteins Decreases Radioresistance but Enhances Cell Motility and Invasion in Triple-Negative Breast Cancer. Int. J. Mol. Sci. 2020, 21, 2169. [Google Scholar] [CrossRef] [Green Version]
  171. Lan, L.; Appelman, C.; Smith, A.R.; Yu, J.; Larsen, S.; Marquez, R.T.; Liu, H.; Wu, X.; Gao, P.; Roy, A.; et al. Natural product (−)-gossypol inhibits colon cancer cell growth by targeting RNA-binding protein Musashi-1. Mol. Oncol. 2015, 9, 1406–1420. [Google Scholar] [CrossRef] [PubMed]
  172. Clingman, C.C.; Deveau, L.M.; Hay, S.A.; Genga, R.M.; Shandilya, S.M.D.; Massi, F.; Ryder, S.P. Allosteric inhibition of a stem cell RNA-binding protein by an intermediary metabolite. Elife 2014, 3, e02848. [Google Scholar] [CrossRef] [PubMed]
  173. Marqus, S.; Pirogova, E.; Piva, T.J. Evaluation of the use of therapeutic peptides for cancer treatment. J. Biomed. Sci. 2017, 24, 21. [Google Scholar] [CrossRef] [Green Version]
  174. Konicek, B.W.; Stephens, J.R.; McNulty, A.M.; Robichaud, N.; Peery, R.B.; Dumstorf, C.A.; Dowless, M.S.; Iversen, P.W.; Parsons, S.; Ellis, K.E.; et al. Therapeutic inhibition of MAP kinase interacting kinase blocks eukaryotic initiation factor 4E phosphorylation and suppresses outgrowth of experimental lung metastases. Cancer Res. 2011, 71, 1849–1857. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Lucchesi, C.A.; Zhang, J.; Ma, B.; Chen, M.; Chen, X. Disruption of the Rbm38-eIF4E Complex with a Synthetic Peptide Pep8 Increases p53 Expression. Cancer Res. 2019, 79, 807–818. [Google Scholar] [CrossRef] [Green Version]
  176. Dong, K.; Wang, R.; Wang, X.; Lin, F.; Shen, J.-J.; Gao, P.; Zhang, H.-Z. Tumor-specific RNAi targeting eIF4E suppresses tumor growth, induces apoptosis and enhances cisplatin cytotoxicity in human breast carcinoma cells. Breast. Cancer Res. Treat. 2009, 113, 443–456. [Google Scholar] [CrossRef]
  177. Mehta, M.; Basalingappa, K.; Griffith, J.N.; Andrade, D.; Babu, A.; Amreddy, N.; Muralidharan, R.; Gorospe, M.; Herman, T.; Ding, W.-Q.; et al. HuR silencing elicits oxidative stress and DNA damage and sensitizes human triple-negative breast cancer cells to radiotherapy. Oncotarget 2016, 7, 64820–64835. [Google Scholar] [CrossRef] [Green Version]
  178. Amreddy, N.; Babu, A.; Panneerselvam, J.; Srivastava, A.; Muralidharan, R.; Chen, A.; Zhao, Y.D.; Munshi, A.; Ramesh, R. Chemo-biologic combinatorial drug delivery using folate receptor-targeted dendrimer nanoparticles for lung cancer treatment. Nanomedicine 2018, 14, 373–384. [Google Scholar] [CrossRef]
  179. Muralidharan, R.; Babu, A.; Amreddy, N.; Srivastava, A.; Chen, A.; Zhao, Y.D.; Kompella, U.B.; Munshi, A.; Ramesh, R. Tumor-targeted Nanoparticle Delivery of HuR siRNA Inhibits Lung Tumor Growth In Vitro and In Vivo By Disrupting the Oncogenic Activity of the RNA-binding Protein HuR. Mol. Cancer Ther. 2017, 16, 1470–1486. [Google Scholar] [CrossRef] [Green Version]
  180. Graff, J.R.; Konicek, B.W.; Vincent, T.M.; Lynch, R.L.; Monteith, D.; Weir, S.N.; Schwier, P.; Capen, A.; Goode, R.L.; Dowless, M.S.; et al. Therapeutic suppression of translation initiation factor eIF4E expression reduces tumor growth without toxicity. J. Clin. Investig. 2007, 117, 2638–2648. [Google Scholar] [CrossRef]
  181. Graff, J.R.; Konicek, B.W.; Lynch, R.L.; Dumstorf, C.A.; Dowless, M.S.; McNulty, A.M.; Parsons, S.H.; Brail, L.H.; Colligan, B.M.; Koop, J.W.; et al. eIF4E activation is commonly elevated in advanced human prostate cancers and significantly related to reduced patient survival. Cancer Res. 2009, 69, 3866–3873. [Google Scholar] [CrossRef] [Green Version]
  182. Chi, X.; Gatti, P.; Papoian, T. Safety of antisense oligonucleotide and siRNA-based therapeutics. Drug. Discov. Today. 2017, 22, 823–833. [Google Scholar] [CrossRef]
  183. Duffy, A.G.; Makarova-Rusher, O.V.; Ulahannan, S.V.; Rahma, O.E.; Fioravanti, S.; Walker, M.; Abdullah, S.; Raffeld, M.; Anderson, V.; Abi-Jaoudeh, N.; et al. Modulation of tumor eIF4E by antisense inhibition: A phase I/II translational clinical trial of ISIS 183750-an antisense oligonucleotide against eIF4E-in combination with irinotecan in solid tumors and irinotecan-refractory colorectal cancer. Int. J. Cancer 2016, 139, 1648–1657. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Chen, H.; Liu, J.; Wang, H.; Cheng, Q.; Zhou, C.; Chen, X.; Ye, F. Inhibition of RNA-Binding Protein Musashi-1 Suppresses Malignant Properties and Reverses Paclitaxel Resistance in Ovarian Carcinoma. J. Cancer 2019, 10, 1580–1592. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Fox, R.G.; Lytle, N.K.; Jaquish, D.V.; Park, F.D.; Ito, T.; Bajaj, J.; Koechlein, C.S.; Zimdahl, B.; Yano, M.; Kopp, J.; et al. Image-based detection and targeting of therapy resistance in pancreatic adenocarcinoma. Nature 2016, 534, 407–411. [Google Scholar] [CrossRef] [Green Version]
  186. King, C.E.; Cuatrecasas, M.; Castells, A.; Sepulveda, A.R.; Lee, J.S.; Rustgi, A.K. LIN28B promotes colon cancer progression and metastasis. Cancer Res. 2011, 71, 4260–4268. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Sánchez-Jiménez, C.; Ludeña, M.D.; Izquierdo, J.M. T-cell intracellular antigens function as tumor suppressor genes. Cell Death Dis. 2015, 6, e1669. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Burdelski, C.; Jakani-Karimi, N.; Jacobsen, F.; Möller-Koop, C.; Minner, S.; Simon, R.; Sauter, G.; Steurer, S.; Clauditz, T.S.; Wilczak, W. IMP3 overexpression occurs in various important cancer types and is linked to aggressive tumor features: A tissue microarray study on 8,877 human cancers and normal tissues. Oncol Rep 2018, 39. [Google Scholar] [CrossRef] [Green Version]
  189. Peritz, T.; Zeng, F.; Kannanayakal, T.J.; Kilk, K.; Eiríksdóttir, E.; Langel, U.; Eberwine, J. Immunoprecipitation of mRNA-protein complexes. Nat. Protoc. 2006, 1, 577–580. [Google Scholar] [CrossRef]
  190. Ko, S.Y.; Guo, H.; Barengo, N.; Naora, H. Inhibition of ovarian cancer growth by a tumor-targeting peptide that binds eukaryotic translation initiation factor 4E. Clin. Cancer Res. 2009, 15, 4336–4347. [Google Scholar] [CrossRef] [Green Version]
Figure 1. By binding to specific RNA structural domains, RBP interacts with RNA to fulfill its specific biological functions. RBP further disrupts post-transcriptional regulatory mechanisms, thereby impacting cancer progression.
Figure 1. By binding to specific RNA structural domains, RBP interacts with RNA to fulfill its specific biological functions. RBP further disrupts post-transcriptional regulatory mechanisms, thereby impacting cancer progression.
Pharmaceuticals 16 01114 g001
Figure 2. Several common RBDs are represented by different shapes and colors, as shown in the diagram. Examples include RRM, KH structural domains, DsRBD, and ZnFs domains. One or more RBDs are arranged and combined in different ways to form RBPs with different functions, which ultimately bind specifically to the target RNA.
Figure 2. Several common RBDs are represented by different shapes and colors, as shown in the diagram. Examples include RRM, KH structural domains, DsRBD, and ZnFs domains. One or more RBDs are arranged and combined in different ways to form RBPs with different functions, which ultimately bind specifically to the target RNA.
Pharmaceuticals 16 01114 g002
Figure 3. RBPs form complexes with snRNP and polypeptides that together regulate the process of selective splicing. U1-U6 represent spliceosomal components, or snRNPs, that bind to the 5′ and 3′ splice sites. As regulators, the RBP, SR protein family, and hnRNP proteins exert an enhancing (+) or inhibiting (−) effect on exon inclusion, respectively, as indicated by the arrows.
Figure 3. RBPs form complexes with snRNP and polypeptides that together regulate the process of selective splicing. U1-U6 represent spliceosomal components, or snRNPs, that bind to the 5′ and 3′ splice sites. As regulators, the RBP, SR protein family, and hnRNP proteins exert an enhancing (+) or inhibiting (−) effect on exon inclusion, respectively, as indicated by the arrows.
Pharmaceuticals 16 01114 g003
Figure 4. The important site where alternative polyadenylation occurs is at the mRNA 3′-UTR, where RBPs enhance (+) or inhibit (−) the use of PAS sites, resulting in 3′-UTRs of different lengths.
Figure 4. The important site where alternative polyadenylation occurs is at the mRNA 3′-UTR, where RBPs enhance (+) or inhibit (−) the use of PAS sites, resulting in 3′-UTRs of different lengths.
Pharmaceuticals 16 01114 g004
Figure 5. (a). RBPs such as AUF1, HuR, TTP, and IGF2BP family proteins regulate the stability of target RNAs by binding to the ARE sequence element of the 3′-UTR in mRNA. (b). The mRNA translation process is dominated by the 5′cap structure of the eIF4 cap-binding complex, which recognizes the relevant elements in the target RNA structure and thus efficiently synthesizes the protein.
Figure 5. (a). RBPs such as AUF1, HuR, TTP, and IGF2BP family proteins regulate the stability of target RNAs by binding to the ARE sequence element of the 3′-UTR in mRNA. (b). The mRNA translation process is dominated by the 5′cap structure of the eIF4 cap-binding complex, which recognizes the relevant elements in the target RNA structure and thus efficiently synthesizes the protein.
Pharmaceuticals 16 01114 g005
Figure 6. HnRNP assists in the translocation of pre-mRNA into the cytoplasm, which then enters the cytoplasm and becomes mRNA, a substrate for mRNA localization. RBP binds to pre-mRNA to form mRNA. RBPs assemble into multiunit complexes that bind RNA to the cytoskeleton and protein motors, ensuring that mRNA can be efficiently transported to the target site.
Figure 6. HnRNP assists in the translocation of pre-mRNA into the cytoplasm, which then enters the cytoplasm and becomes mRNA, a substrate for mRNA localization. RBP binds to pre-mRNA to form mRNA. RBPs assemble into multiunit complexes that bind RNA to the cytoskeleton and protein motors, ensuring that mRNA can be efficiently transported to the target site.
Pharmaceuticals 16 01114 g006
Figure 7. Current RBP-based targeted therapeutic strategies can be classified as either RBP-specific manipulation or RBP-RNA interactions. The cancer therapeutic approaches summarized in figure (a) are classified as small molecule inhibitors, therapeutic peptides, ASOs, siRNAs, RBP-binding partners, and downstream effectors. The main mechanisms are to inhibit RBP-RNA interactions by inducing degradation, inhibiting enzyme activity, blocking post-transcriptional modifications, or competing for selected RBPs through binding. Schematic representation of RBP-targeted therapeutic strategies is shown in figure. (b) Small molecule inhibitors, figure. (c) Therapeutic peptides and siRNAs, figure. (d) ASO, RBP-binding partners, and downstream effectors.
Figure 7. Current RBP-based targeted therapeutic strategies can be classified as either RBP-specific manipulation or RBP-RNA interactions. The cancer therapeutic approaches summarized in figure (a) are classified as small molecule inhibitors, therapeutic peptides, ASOs, siRNAs, RBP-binding partners, and downstream effectors. The main mechanisms are to inhibit RBP-RNA interactions by inducing degradation, inhibiting enzyme activity, blocking post-transcriptional modifications, or competing for selected RBPs through binding. Schematic representation of RBP-targeted therapeutic strategies is shown in figure. (b) Small molecule inhibitors, figure. (c) Therapeutic peptides and siRNAs, figure. (d) ASO, RBP-binding partners, and downstream effectors.
Pharmaceuticals 16 01114 g007
Table 1. Roles of RNA-binding proteins (RBPs) in cancer.
Table 1. Roles of RNA-binding proteins (RBPs) in cancer.
RBPThe Basic Mechanism of RBPs RegulationTumor TypeBiological FunctionsReferences
HuRSubcellular localization, mRNA stability,
mRNA translation
Gastric Cancer,
Breast Cancer,
Colon Cancer,
Lung Cancer,
Varian Cancer
Exerts
proliferation anti-apoptotic effects
[129,130,131,132]
RBM38mRNA stability,
mRNA translation,
post-transcriptional regulation,
mRNA splicing
Colorectal Cancer,
Acute Myeloid Leukemia,
Renal Cell Carcinoma,
Hepatocellular Carcinoma
Inhibit EMT, stemness,
invasiveness
[133,134,135,136]
eIF4ETranslationB-Cell Lymphoma,
Breast, Colon,
Lymphoma, Melanoma
Promotes apoptosis,
angiogenesis, EMT,
invasion,
metastasis
[137,138,139]
hnRNPD
(AUF1)
mRNA stabilityBreast Cancer,
Colon Cancer,
Stomach Cancer
Liver, Lung,
Pancreatic Cancer Sarcoma, Thyroid Cancer
Promotes proliferation,
Senescence
[92,140,141,142,143]
hnRNPA2/B1Alternative splicingBrain Tumor,
Lung Cancer
Promotes
proliferation, EMT,
metastasis
[122,144,145,146,147,148]
TRBPmRNA translation, mRNA stabilityBreast Carcinomas,
Colorectal Cancer,
Endometrial Cancer
Promotes or inhibits cell proliferation
and invasion
[149,150]
IGF2BP1
(IMP1/ZBP1)
Subcellular
localization,
mRNA stability
Breast Cancer,
Colon Cancer,
Lung Cancer,
Melanoma,
Ovarian Cancer,
Skin Cancer,
Liver Cancer
Promotes
proliferation, EMT,
invasion, metastasis
[99,100,110,111,151,152,153,154,155]
IGFBP2
(IMP2)
Subcellular
localization,
mRNA stability
Breast Cancer,
Leukemia,
Lung Cancer,
Colon Cancer
Promotes
EMT,
invasion,
metastasis
[155,156,157,158]
Table 2. RBP-based targeted therapeutic strategies in cancer.
Table 2. RBP-based targeted therapeutic strategies in cancer.
Therapeutic TypesRBPTherapeutic ApproachesFunctionsTumor TypesReferences
Small
molecule
inhibitors
HuRMS-444,
DHTS,
AZA-9
Targets RRM1 and RRM2 of HuR and inhibits RNA-binding activities of HuRPancreas, Colon, Melanoma, Brain, Breast[161,162,164,165]
eIF4ERibavirin
4Ei-1
Impedes eIF4E; antagonizes eIF4E cap binding; and initiates degradationHscc,
Aml, Breast, Lung, Mesothelioma
[168,169]
MSI-1(-)-gossypol;
Oleic acid
Blocks RBP binding site with RNA;
interfere with MSI-1 expression
Brain (CNS)
Colon
[171,172]
Therapeutic
peptides
eIF4EGnRH-4EBPBinds to eIF4E and disrupts eIF4E interacting with eIF4GOvary[175,190]
RBM38Pep8Antagonizes RBM38
and promotes p53 expression
Colon, Breast
ASOeIF4EISIS 183750Inhibits the proliferation of cancer cellsColon[183]
VEGF,
c-Myc,
Survivin,
BCL-2
Cyclin D1
Inhibit the translation of target mRNAs to inhibit tumor growth
siRNAeIF4E Stimulates the cytotoxic effects of cisplatinBreast[176]
HuR Silences HuR expression; interferes with the binding of HuR with mRNALung[178,179]
Binding partners
and
downstream effectors
HuRMYCTargeted regulation of MYC transcript levels and activity; intervenes in RBP-affected cancer progression [163,188]
hnRNPA1
hnRNPH
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Feng, Y.; Zhu, S.; Liu, T.; Zhi, G.; Shao, B.; Liu, J.; Li, B.; Jiang, C.; Feng, Q.; Wu, P.; et al. Surmounting Cancer Drug Resistance: New Perspective on RNA-Binding Proteins. Pharmaceuticals 2023, 16, 1114. https://doi.org/10.3390/ph16081114

AMA Style

Feng Y, Zhu S, Liu T, Zhi G, Shao B, Liu J, Li B, Jiang C, Feng Q, Wu P, et al. Surmounting Cancer Drug Resistance: New Perspective on RNA-Binding Proteins. Pharmaceuticals. 2023; 16(8):1114. https://doi.org/10.3390/ph16081114

Chicago/Turabian Style

Feng, Yiyuan, Sha Zhu, Tengwen Liu, Guoguo Zhi, Bingjie Shao, Jibin Liu, Baixue Li, Cen Jiang, Quansheng Feng, Peijie Wu, and et al. 2023. "Surmounting Cancer Drug Resistance: New Perspective on RNA-Binding Proteins" Pharmaceuticals 16, no. 8: 1114. https://doi.org/10.3390/ph16081114

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop