Next Article in Journal
Investigation of the Associations of Novel Inflammatory Biomarkers—Systemic Inflammatory Index (SII) and Systemic Inflammatory Response Index (SIRI)—With the Severity of Coronary Artery Disease and Acute Coronary Syndrome Occurrence
Next Article in Special Issue
Patient-Derived Multiple Myeloma 3D Models for Personalized Medicine—Are We There Yet?
Previous Article in Journal
HER2 c-Terminal Fragments Are Expressed via Internal Translation of the HER2 mRNA
Previous Article in Special Issue
Alpha-Ketoglutarate or 5-HMF: Single Compounds Effectively Eliminate Leukemia Cells via Caspase-3 Apoptosis and Antioxidative Pathways
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of RNA-Binding Proteins in Hematological Malignancies

by
Pedro Aguilar-Garrido
1,2,†,
Álvaro Otero-Sobrino
1,2,†,
Miguel Ángel Navarro-Aguadero
1,2,
María Velasco-Estévez
1,2,*,‡ and
Miguel Gallardo
1,2,*,‡
1
H12O-CNIO Hematological Malignancies Clinical Research Unit, CNIO, 28029 Madrid, Spain
2
Department of Hematology, Hospital Universitario 12 de Octubre, Instituto de Investigación Sanitaria Hospital 12 de Octubre (imas12), 28041 Madrid, Spain
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2022, 23(17), 9552; https://doi.org/10.3390/ijms23179552
Submission received: 8 August 2022 / Revised: 18 August 2022 / Accepted: 21 August 2022 / Published: 23 August 2022

Abstract

:
Hematological malignancies comprise a plethora of different neoplasms, such as leukemia, lymphoma, and myeloma, plus a myriad of dysplasia, such as myelodysplastic syndromes or anemias. Despite all the advances in patient care and the development of new therapies, some of these malignancies remain incurable, mainly due to resistance and refractoriness to treatment. Therefore, there is an unmet clinical need to identify new biomarkers and potential therapeutic targets that play a role in treatment resistance and contribute to the poor outcomes of these tumors. RNA-binding proteins (RBPs) are a diverse class of proteins that interact with transcripts and noncoding RNAs and are involved in every step of the post-transcriptional processing of transcripts. Dysregulation of RBPs has been associated with the development of hematological malignancies, making them potential valuable biomarkers and potential therapeutic targets. Although a number of dysregulated RBPs have been identified in hematological malignancies, there is a critical need to understand the biology underlying their contribution to pathology, such as the spatiotemporal context and molecular mechanisms involved. In this review, we emphasize the importance of deciphering the regulatory mechanisms of RBPs to pinpoint novel therapeutic targets that could drive or contribute to hematological malignancy biology.

1. Introduction

Hematological malignancies comprise a heterogeneous group of cancers that includes diverse and biologically different subgroups of neoplasms. The proposed WHO classification of hematological malignancies stratifies these neoplasms according to their original cell lineage: myeloid neoplasms, lymphoid neoplasms, mast cell disorders, and histiocytic neoplasms [1]. Each type varies in incidence, etiology, prognosis, and survival, and they are all essential contributors to the global cancer burden. In 2020, hematological neoplasms accounted for more than 7% of all diagnosed cancers worldwide and this number rose to 10% in the United States of America (USA) [2]. The incidence of hematological neoplasms varies depending on subtype, age, sex, and socioeconomic state of the patient [3]. In the USA alone, circa 200,000 people get diagnosed annually with hematological malignancies, the most common one being non-Hodgkin lymphoma followed by leukemia and multiple myeloma [2,4], and they account for 9.7% of all cancer deaths [2].
Relative survival for many hematological neoplasms has significantly increased in the past decade. Most hematological neoplasms have had the same robust frontline treatment for the past decades, such as Rituximab to treat lymphoma (approved in 1997), Bortezomib and Lenalidomide to treat myeloma (approved in 2003 and 2005, respectively), or Imatinib for chronic myeloid leukemia (CML) (approved in 2005). Indeed, there is a 10-year relative survival greater than 50% for most hematological malignancies, except for acute leukemia and multiple myeloma [5]. However, in the past 5 years, the emergence of novel therapies targeting poor prognosis neoplasms and/or relapse subset of patients is starting to change the clinical situation. Currently, these malignancies can be treated with multiple treatment regimens and therapy combinations, including precision medicine such as target therapy molecules (e.g., small molecule inhibitors), monoclonal antibodies, antibody-drug conjugates (ADC), bispecific T-cell engagers (BiTEs), immunotoxins, immune checkpoint inhibitors and Chimeric Antigen Receptor cell technology, efficiently improving the survival of these patients [6,7,8,9,10,11].
However, although there has been an improvement in the development of new therapies and medical care of patients, the heterogeneity of hematological malignancies constitutes a therapeutic obstacle, and the current small molecule inhibitors available in the clinic only target a subset of patients. Tumor cells have various strategies to evade/escape therapy, such as gain of genetic mutations, epigenetic alterations, abnormal drug metabolism, signaling pathway dysfunction, tumor microenvironment rewiring and stem cell persistence [12,13,14]. This way, malignant cells circumvent current therapies and survive, leading to therapy resistance, relapse, and failure of treatments.
Therefore, there is an urgent unmet clinical need to identify new biomarkers and potential therapeutic targets that would contribute to overcoming the current therapy resistance and poor outcomes of hematological neoplasms. Here, we review the role and potential clinical significance of RNA-binding proteins (RBPs) in hematological malignancies, both as biomarkers and therapeutic targets for different neoplasms.

2. RNA-Binding Proteins

Cancer involves a myriad of cellular and molecular mechanisms; however, transcription and translation have gained more interest in cancer research as they are critical in gene expression and protein biogenesis and seem to highly contribute to the development and progression of a number of different tumors. Nevertheless, the underlying mechanisms causing these alterations remain unknown.
Gene transcription and mRNA translation are tightly regulated processes where RBPs are closely involved. These RBPs represent a group of regulatory proteins with structural domains with high affinity for specific RNA sequences, which are key components of RNA biogenesis [15]. RBPs are increasingly becoming more popular as they are essential for the proper formation of blood cellular components (termed hematopoiesis) [16]. Indeed, studies of gene expression profiling have demonstrated that different blood cell differentiation stages are not only regulated by transcription factors but also influenced by post-transcriptional regulatory mechanisms, such as those controlled by RBPs [17,18]. For this reason, RBPs can be key players in the onset and progression of hematological malignancies.
RBPs are involved in many aspects of the post-transcriptional modulation of mRNAs and are deeply regulated by post-translational modifications (PTMs), such as phosphorylation, acetylation, or methylation. Depending on the PTMs acquired and the balance between them, the properties of RBPs dramatically change, affecting their interaction with other molecules, their biological function, and their subcellular localization [19]. Because RBPs are implicated in multiple cellular processes, perturbations in RBP–RNA complex activity have been associated with cancer development and progression. In recent years, an increasing number of studies have pointed out the key role of RBPs in hematological malignancies [20,21,22].
To date, more than 500 RBPs have been characterized in humans, all of which have different affinities for RNA [23]. Although they can also interact with proteins and DNA, most of the identified RBPs specifically bind to RNAs (including mRNA) through the recognition of specific sequences mostly located in the untranslated regions (UTRs) [24,25]. Within the structure of most known RBPs, there is a small group of “classic” specific domains known as RNA-binding domains (RBDs) which mediate the interaction between RBPs and RNAs in a sequence- and structure-dependent manner. These include the zinc-finger and DEAD/DEAH box helicase, RNA recognition motif (RRM), or K homology (KH), among others (Table 1). These domains usually contain between 60–100 amino acids that normally adopt an αβ topology [26].
Commonly, RBPs contain multiple repeats of the same RBD, which contributes to an enhanced specificity and affinity for their target mRNAs [27,54]. Nonetheless, other RBPs are also able to interact with mRNA through “nonclassic” RBDs, thus allowing them to bind to a wider spectrum of molecules [24].
RBPs and transcripts assemble in what are termed ribonucleocomplexes to control gene expression. These dynamic interactions determine the fate of the transcripts in every step of its cycle, from synthesis to decay, depending on the combination and balance of each type of RBP forming the ribonucleocomplex at each time point [59,60]. Because the same recognition sequences could be shared by many different RNAs, the expression of multiple mRNAs could be controlled by a single RBP. Therefore, each RBP can regulate a plethora of molecular pathways and different physiological processes depending on which mRNAs are expressed [61].
Despite the complexity of the genomic and mRNA network, this system has great versatility due to the continuous modification of the transcriptome as a result of the high activity rate of the mRNA machinery. Depending on the cellular needs at a given time, the gene regulatory networks persistently adjust mRNA and protein levels, influencing from transcription to protein degradation, which includes mRNA transport and maturation and translation [62,63,64].

3. mRNA Regulation by RBPs

The information coded in the DNA must be transcribed by RNA polymerases into RNA molecules to generate proteins. These primary transcripts are usually modified and processed to mature them into functional molecules. Therefore, the extensive processing of protein-encoding pre-mRNA transcripts is essential in eukaryotic gene expression [65]. As mentioned before, RBPs are involved in many steps of RNA metabolism and have the ability to control post-transcriptional gene regulation by many mechanisms.

3.1. mRNA 5′ Capping

The production of mRNA requires the synthesis of a pre-mRNA followed by a 5′ capping, deletion of introns by splicing, and a final 3′ cleavage and polyadenylation, in order to produce a mature mRNA. The addition of a 7-methylguanosine to the first nucleotide of the 5′ of a pre-mRNA through a triphosphate bound is known as capping [66]. Just before finishing transcription, pre-mRNAs are capped. This process is essential for the growth of eukaryotic cells, as it participates in many functions such as the regulation of nuclear export, the prevention of mRNA degradation, the fostering of translation and the promotion of 5′ proximal intron excision [66,67]. This 5′ capping is a tightly regulated process that allows the generation of a stable and mature mRNA molecule able to undergo translation by the ribosome.
In mammals, all 5′ caps of pre-mRNAs bind to a heterodimer known as the nuclear cap binding complex (CBC), which is formed by two different proteins: CBP20 (cap binding protein 20) and CBP80 (cap binding protein 80). This heterodimer protein complex participates in monitoring mRNA quality through the nonsense-mediated mRNA decay pathway (NMD), in which there is an elimination of mRNA transcripts that contain premature stop codons, leading to a reduction in errors in gene expression [68].
Prevention of mRNA degradation is one of the principal functions of mRNA capping. It prevents the digestion of mRNAs by the exoribonuclease activity of Xrn1 and Xrn2, to continue its journey to the ribosome. Once in the cytoplasm, 5′ capping acts as a trigger for the recruitment of the translational machinery to the 5′ end of protein-encoding RNAs, inducing the recognition of the cap by translation initiation factors (e.g., eIF4E) and leading to the recruitment of the 40S ribosome subunit to the 5′ end [69,70].

3.2. Splicing of Pre-mRNAs

The process of removing introns (noncoding regions of RNA) and splicing exons (coding regions) back together from a pre-mRNA to generate a mature mRNA is known as RNA splicing, and it is crucial for bona fide protein expression.
In eukaryotes, splicing mainly occurs in a series of reactions catalyzed by the spliceosome, a huge macromolecular complex consisting of small nuclear ribonucleoproteins (snRNPs) together with more than 100 different other proteins. However, most transcripts can undergo alternative splicing as well, in which a selection of alternative exons could be executed, resulting in the production of different isoforms of the same mRNA [71].
Due to this alternative splicing, different mRNA isoforms are produced in different tissues and cell types within them. This process alters gene coding and has an impact on mRNA stability. It has been recognized as a mechanism to increase the functional diversity of the proteome, leading to an additional layer of complexity for gene expression regulation. Splicing processes can occur both during transcription of mRNAs (cotranscriptional) or after this process (post-transcriptional) [72,73].
Many proteins of the spliceosome are well-studied RBPs, such as DExD/H type RNA-dependent ATPase. These are RBPs with helicase activity and they are implicated in rearrangements, single-strand RNA translocation, strand annealing and protein displacement. Due to the RBPs’ wide range of molecular processes implication, a plethora of diseases are linked to mutations in many RBPs involved in splicing and splicing regulation [26,74].

3.3. Cleavage and 3′ End Formation

The processes of cleavage and 3′ end formation are fundamental steps in the maturation of pre-mRNAs into functional mRNAs that can be exported from the nucleus to the cytosol.
The final stage of transcription is marked by endonucleolytic cleavage, prior to 3′ end formation, where the 3′ end of a newly pre-mRNA is first cleaved-off by a set of proteins, followed by the addition of a poly(A) tail at the 3′ end of an mRNA (polyadenylation) of approximately 200 adenosine residues [75].
The poly(A) tail is important for nuclear export to the cytosol, translation initiation and stability of mRNA. Similar to splicing, in some genes, transcripts can be polyadenylated at several possible sites, leading to the generation of more than one transcript from a single gene. This process is known as alternative polyadenylation and alters the stability, localization and transport of these pre-mRNAs. Within the human genome, it is estimated that half of the genes, at least, are subject to this alternative splicing and polyadenylation, generating isoforms that differ in the 3′ UTR length or, even, encoding different proteins [76].
The polyadenylation machinery consists of six multimeric proteins that act together to, first, mediate the cleavage of newly nascent pre-mRNA and, second, allow the polyadenylation of the transcript. It is as assembled by a cleavage stimulatory factor (CstF), cleavage and polyadenylation specificity factor (CPSF), cleavage factors I and II (CFI and CFII), poly(A) polymerase (PAP) and poly(A) binding protein II (PABP II) [75,76].

3.4. mRNA Export

The largest cell organelle is the nucleus, surrounded by the nuclear envelope, which is pierced by nuclear pore complexes (NPCs). These are the main gateways through which mRNAs are transported between the nucleus and the cytosol, although this translocation requires specific transport receptors to break this barrier.
The mRNA export machinery includes numerous RBPs (e.g., hnRNP C tetramer), adaptor proteins (EJC and SR) and proteins that associate with NPCs (TREX complex). All of these proteins interact with the mRNA at some point, facilitating the transport from the nucleus to the cytosol [77,78]. After mRNA is exported from the nucleus, all the interacting RBPs can either remain in the nucleus or follow the transcript into the cytosol. Once in this compartment, mRNAs could remain interacting with the same RBPs or they can be recruited by others, which determines mRNA cytosol compartment localization [79].

3.5. mRNA Stability

The abundance of an mRNA depends not only on its synthesis, processing and nuclear export rates, but also on its degradation rate in the cytosol. Thus, mRNA stability maintenance plays a key role in the translation of essential proteins and in the degradation of transcripts mediated by the proteasome.
Cytosolic stability of mRNAs relies upon mRNA interactions with RBPs and other types of proteins, which are crucial for the modulation of this stability. This process can be quickly modulated to alter specific gene expression, giving a huge flexibility in affecting changes in patterns of protein synthesis [80].
RBPs contribute to controlling the degradation of transcripts, which is an important step and is highly variable. One of the main processes observed in the alteration of mRNA stability is the alteration of the length of the poly(A) tail, which is related to the earlier decapping [81].

3.6. Translation

The translation process is one of the most well-studied and best-detailed processes in which RBPs play a key role. Briefly, the regulation of the translation process is controlled by three main mechanisms: poly(A) tail modification, interaction and association with RBPs, and assembly of the translation machinery [82].
The increase in the poly(A) tail length of an mRNA can significantly improve translation rates, which depends on the activity of PABP, an RBP responsible for recruiting many factors essential for the initiation of translation [83].
Transcript-specific cytosolic compartmentalization of RBPs constitutes a targeted location for processing, sorting, storage or degradation. Some RBPs can be associated with exonucleases to be degraded, while others can be protected and remain silent until they are needed in the ribosome for protein production [84].

3.7. RNA Editing

The process of insertion, deletion or subtraction of nucleotides in RNA sequences is known as RNA editing. The most common RNA editing process is the exchange of adenosine to inosine (A-I), catalyzed by the RBP family ADARs, such ADAR1, but with the regulation and collaboration of other RBPs such as DROSHA or TARDBP [85]. A-I editing promotes deep modification of transcriptomic landscape by protein modification, RNA splicing, export or microRNAs activity.

4. RNA-Binding Proteins in Hematological Malignancies

Diseases such as neurodegeneration, dysplasia, cancer, ribosomopathies and hematological malignancies can all be caused by defects in the normal functionality or expression of RBPs. Currently, RBPs are understood to be crucial for healthy hematopoiesis and to play a significant role in hematological neoplasms by functioning as both oncogenes and tumor suppressors [86,87,88,89].
RBP modifications result in biological changes such as altered hematopoietic lineage development or abnormalities that, in many circumstances, can result in bone marrow failure [90,91]. Likewise, alterations in several RBPs also cause variations in splicing in myeloid and lymphoid cell lineages [16]. Because RBPs act as nodes for multiple signaling pathways and take part in hematological malignancies, they offer a therapeutic opportunity in the clinic, as potential new biomarkers as well as new therapeutic targets for hematological neoplasms Figure 1.
Some of the most significant RBPs, whose dysregulation has been associated with the development of hematological malignancies, can be found summarized in Table 2.

4.1. RBM39–mRNA Splicing

RNA-binding motif protein 39 (RBM39), also known as HCC1, CAPER or CAPERɑ, is a serine/arginine-rich RNA-binding protein implicated in the regulation of gene expression at several levels, such as transcription, alternative splicing, and translation [118]. RBM39 contains two central RRMs, an N-terminal arginine- and serine-rich domain (RS domain) which can be phosphorylated, and a C-terminal U2AF homology motif (UHM). Both the N- and C-terminal domains enable RBM39 to interact with other proteins. Thus, RBM39 plays a critical role in the activation of transcription factors, such as activating protein-1 (AP-1)/Jun, steroid receptors and transcriptional coactivators, such as ASC-2 [119,120]. Furthermore, c-Abl phosphorylates RBM39, enhancing its transcriptional activity [121]. RBM39 has been described as a potential splicing factor that acts in pre-mRNA alternative splicing by binding to RNA or recruiting specific splicing factors to regulate this process [113]. RBM39 is implicated in the recognition of 3′ splice sites and can interact with several molecules in early spliceosome assembly that contain the U2AF ligand motif (ULM), such as U2AF65, SF3B155, and RSRC1.
RBM39 is involved in several solid cancers, such as breast and lung cancer [114,122,123], but it is also implicated in hematological malignancies. Wang et al. [124] used a comprehensive screen to pinpoint RBM39 as a crucial RBP necessary for AML survival, and demonstrated therapeutic effects by pharmacologically reducing RBM39 protein in spliceosome mutant AML models. They identified RBM39 as a key AML-specific RBP, involved in AML malignant cell growth and maintenance [124]. Moreover, its depletion leads to a decrease in leukemia progression and an increase in overall survival. Likewise, it has been shown that RBM39 plays an important role in multiple myeloma, by promoting multiple myeloma cell proliferation and inhibiting apoptosis under hypoxic conditions in vitro [125]. Tong and collaborators also described that a long noncoding RNA, DARS-AS1, inhibits RBM39 degradation under hypoxic conditions. They demonstrated that DARS-AS1 is regulated by HIF-1α and that DARS-AS1 regulates mTOR signaling via RBM39, thus concluding that the HIF1α/DARS-AS1/RBM39 axis could be a useful target in multiple myeloma.

4.2. Musashi Proteins (MSI1 and MSI2)-mRNA Translation

The Musashi protein family is a group of RNA-binding proteins that play necessary roles in controlling stem and progenitor cell functions [109,110,126,127,128]. This family contains Musashi-1 (MSI1) and Musashi-2 (MSI2), two highly conserved homologous proteins [109]. MSI2 is transcribed ubiquitously, in contrast with MSI1, which is mainly found in neural stem cells or embryonic progenitor cells and is involved in neural stem cell proliferation and cell cycle regulation [128]. On the other hand, MSI2 is implicated in the self-renewal and differentiation of embryonic [129], neuronal [110] and hematopoietic stem cells [111]. This suggests that each member of this family is regulated in a context-dependent manner.
The Musashi RNA-binding proteins constitute a subgroup within the heterogeneous nuclear ribonucleoprotein (hnRNP) A/B type of proteins that contain two N-terminal copies of RNA recognition motifs (RRMs) in addition to a C-terminal PABP-interacting domain [109]. RRM1 mainly binds to RNA whereas RRM2 has a complementary role, stabilizing the protein–RNA complex and increasing RNA binding affinity in combination with RRM1 [41]. MSI1 and MSI2 recognize different mRNA motifs, preferentially within the 3′-UTR (3′ untranslated region) of target transcripts, although it is thought that they regulate similar mRNA targets as there is a partial overlap between MSI1 and MSI2 binding targets [130]. The C-terminal region of MSI proteins allows protein–protein interactions, containing several protein-binding domains and contributing to the stimulation or repression of protein translation. In contrast with MSI1, which only has one isoform, the MSI2 gene produces four different isoforms (MSI2a, MSI2b, MSI2c and MSI2d), all of them containing the same RRM motifs and PABP domain, but differing on the C-ter and N-ter amino acids, which makes them susceptible to different PTMs [131]. The canonical MSI2a can be phosphorylated at its C-terminal, converting it from a translational repressor to an antivator, while MSI2b lacks this regulatory site and thus cannot promote translation of target mRNAs [132]. Therefore, MSI2 can lead to the translation or repression of different target mRNAs, and despite the exact underlying molecular mechanisms being not yet fully clear, it seems that the regulation by MSI2 depends on the isoform, phosphorylation status, and cellular context [131,132].
Elevated expression of Musashi proteins has been identified in several solid tumors [132] and this overexpression is extensively correlated with cancer cell proliferation. In contrast with MSI1, MSI2 expression has been predominantly reported in hematological neoplasms, fundamentally in leukemia, probably due to the fact that MSI2 is mainly expressed in the hematopoietic system and is an important regulator of hematopoietic stem cells (HSCs) [111]. High MSI2 expression levels are found in nearly all hematological malignancies and are associated with poor prognosis [130,133].
Functional studies have shown an essential role of MSI2 in leukemic progression and stem cell renewal. Kharas et al. described that MSI2 dysregulation is associated with aggressive AML cases and that its inhibition in human AML cell lines resulted in reduced proliferation and increased apoptosis [111]. HoxA9, c-Myc and Ikzf2 were shown to be targets of MSI2, maintaining the oncogenic leukemia stem cell (LSC) self-renewal program in AML [134]. The most prominent cancer pathway regulated by Musashi proteins is the Notch pathway. The main studied target of Musashi proteins is Numb, an inhibitor of the Notch signaling pathway [135]. The Musashi2-Numb axis is crucial to regulate chronic myeloid leukemia (CML) progression to blast crisis, and its depletion results in reduced leukemia development in CML models [136]. In chronic lymphocytic leukemia (CLL), MSI2 is implicated in CLL cell growth and survival, and correlates with worse clinical outcome [137]. Palacios and colleagues described critical functions of MSI2 in CLL progression, probing its driving role in proliferative CLL cells. They also demonstrated that Ro 08-2750, an MSI2 small molecule inhibitor whose efficacy was previously reported in a murine AML model [138], preferentially and more effectively targeted leukemic B cells and myeloid cells while not affecting HSCs [105]. MSI2 is also involved in myelodysplastic syndrome (MDS) progression, as conditional deletion of MSI2 in mouse models led to a reversal of MDS phenotypes, and its overexpression resulted in MDS progression [105].

4.3. IGF2BP3–mRNA Localization, mRNA Stability and mRNA Degradation

Insulin-like growth factor 2 mRNA binding protein 3 (IGF2BP3), also known as IMP3, KOC, CT98, KOC1, and VICKZ3, belongs to the IGF2BP family, which is composed of three highly structurally and functionally related proteins. These RBPs carry two amino-terminal RRMs and four carboxy-terminal KH domains [139]. Although several studies report that these proteins interact with RNA via KH domains [140,141], it has been subsequently described that RRMs also contribute to RNA binding, enabling a very specific interaction with RNA targets [142]. IGF2BP3 expression follows an oncofetal pattern, since it is highly expressed during embryonic development [139], while its levels decrease at the late stages of embryogenesis and it is upregulated in several adult cancers [143,144].
IGF2BP3 controls the localization, stability and degradation of its mRNA targets [143]. This protein is primarily cytoplasmic, but it can also act in the nucleus where it binds target transcripts and facilitates their nuclear export [145] due to the presence of nuclear export signals within the KH2 and KH4 domains [139]. IGF2BP3 has the bimodal capacity to regulate mRNA fate, avoiding or promoting binding to the RNA-induced silencing complex (RISC). Functional studies reveal that IGF2BP3 directly interacts with ribonuclease XRN2 [146] and deubiquitinase USP10 [147], leading to EIF4EBP mRNA degradation or tumor suppressor p53 reduction, respectively. It has been observed that IGF2BP3 is highly expressed in the majority of solid cancers, promoting a proliferative phenotype, metastatic activity and poor outcomes of these tumors [106].
This RBP has been identified as a key player in many hematological malignancies as well. IGF2BP3 is necessary for B-cell acute lymphoblastic leukemia (B-ALL) cell survival. Its overexpression in the bone marrow of mice causes abnormal growth of hematopoietic stem cells [107]. In 2021, Tran et al. demonstrated the critical role of IGF2BP3 in mixed-lineage leukemia (MLL)-rearranged leukemia, a subtype of acute leukemia with poor outcomes, high relapse and therapy refractoriness. They proved that IGF2BP3 itself is a direct target of MLL-AF4 and regulates a post-transcriptional operon that increases the expression of leukemogenic genes [148]. A recent study by Mäkinen et al. also showed that IGF2BP3 is associated with active cellular proliferation in B-cell blasts [108]. Surprisingly, they found that high IGF2BP3 mRNA levels were related to improved survival in a high-risk pediatric B-ALL cohort.
In multiple myeloma, IGF2BP3 expression is inhibited by miR-9-5p, which is upregulated after AR-42 treatment, a histone deacetylase inhibitor [149]. IGF2BP3 controls CD44, a glycoprotein that has been associated with current therapy resistance in multiple myeloma.
Related to myelodysplastic syndromes, apoptotic bodies from cytosine arabinose-resistant cells notably contain IGF2BP3, which promotes the survival of recipient cells by activating the PI3K-Akt and p42-44 MAPK pathways via c-Myc [150].

4.4. hnRNP K–mRNA Localization, Transcription, Translation and Splicing

Heterogeneous nuclear ribonucleoprotein K (hnRNP K) is a multifunctional RBP implicated in a myriad of biological processes, such as transcription, translation, splicing, RNA stability and chromatin remodeling, among others [29,95]. It contains three KH domains (KH1, KH2 and KH3) responsible for nucleic acid binding [151], one K protein interactive domain (KI) responsible for protein binding, and a nuclear-cytoplasmic shuttling domain (KNS) that confers the ability to translocate bidirectionally through the nuclear pore complex [152].
hnRNP K is an intracellular protein that localizes both in the nucleus and the cytoplasm and is ubiquitously expressed in all tissues. This protein could be modified by many other proteins, as it can be methylated, acetylated, sumoylated or phosphorylated. Depending on this PTM balance, hnRNP K can regulate its interactions with different molecules and influence its functions, having a huge impact on many biological pathways [95]. Because of its multiple interactions with many proteins, alterations of its protein levels can be oncogenic by deregulating the transcription and/or translation of multiple oncogenes or tumor suppressors, supporting the hypothesis that RBPs are involved in tumorigenesis [96].
Many studies have demonstrated the link between hnRNP K deregulation and cancer. hnRNP K is a key player in DNA damage response as it is required for p53-mediated transcription of cell cycle checkpoint genes [97,153], and thus could promote cell proliferation via p21, a negative regulator of cyclin-dependent kinases (CDKs), and its subsequent reduction in G0/G1 stage cells [154]. Numerous studies have shown that long noncoding RNAs (lncRNAs) influence the growth of tumors by either promoting or inhibiting hnRNP K expression [98,154,155,156,157].
It has been reported that approximately 2% of AML patients have a 9q21.32 deletion that encompasses the HNRNPK gene [99,100,101,158]. For the first time, Gallardo et al. described that this haploinsufficiency is linked to the onset of AML phenotypes in animal models, pointing out hnRNP K’s tumor suppressor role [21]. Moreover, years later, Naarman-de Vries IS et al. found that the 9q21.32 deletion produces a reduction in hnRNP K-targeted gene expression, such as CDKN1A and CEBPA, which is related to AML del(9q) pathogenesis. Furthermore, it seems that hnRNP K influences the differentiation of AML-derived cells through its interaction with the transcription factor PU.1 [102].
Related to leukemia, several studies have demonstrated that hnRNP K also has an impact in chronic myeloid leukemia (CML) [103,104]. The activity of BCR/ABL kinase induces hnRNP K expression by enhancing hnRNP K gene transcription and mRNA stability. As mentioned before, c-Myc mRNA can be overexpressed due to hnRNP K overexpression, a situation that is essential for the phenotype of CML cell progenitors.
Recently, Gallardo et al. found that diffuse large B-cell lymphoma (DLBCL) patients with high hnRNP K expression suffered poor clinical outcomes compared with patients with low hnRNP K expression [22]. In contrast with previous studies [21], they showed that hnRNP K could act as an oncogene forcing c-Myc expression and contributing to DLBCL pathogenesis.
Multiple myeloma (MM) pathogenesis is also correlated with hnRNP K deregulation. The study of Evans JR et al. showed that at least 42% of MM patients carry a mutated version of the c-Myc IRES that binds more tightly to hnRNP K. Thus, the upregulation of hnRNP K leads to an increase in c-Myc, contributing to MM development [159].
Moving out of neoplasms, the presence of hnRNP K autoantibodies has been found in the serum of patients with immune-mediated aplastic aplasia (AA), a type of bone marrow failure syndrome. Qi Z et al. suggested that specific immune responses to hnRNP K may induce the polarization of Th1 CD4+ cells and may contribute to the development of AA. It seems that the destruction of immature hematopoietic cells overexpressing hnRNP K may induce a specific immune response to this RBP in patients with immune-mediated AA [90].
Lastly, Aguilar-Garrido et al. described the role of hnRNP K in non-neoplastic hematological malignancies. In vivo overexpression of hnRNP K promoted the exhaustion of hematopoietic stem cells by ribosome dysfunction, inducing a bone marrow failure phenotype [91].

5. Discussion

The fate of mRNAs mostly depends on their relationship and network with RBPs, but the principles underlying these interactions remain poorly understood. How RBPs achieve binding specificity through their different RBDs and how they interact and compete with other RBPs is something not yet fully characterized [160]. Furthermore, the precise mRNA sequences that RBPs bind are still being fully identified, and it is unknown whether they are cell context specific.
RBPs are crucial for controlling gene expression at every stage, and their dysregulation can result in a wide variety of hematopoietic malignancies [88]. Excitement and rising interest in this new topic have been generated by our growing understanding of the discovery of these unregulated RBPs and their function in hematological malignancies.
However, the list of unconventional RBPs with newly discovered biological roles continues to grow, as is the case for RBM39, MSI1, MSI2, IGFBP3 and hnRNP K.
Strikingly, several RBPs have both oncogenic and tumor suppressor functions, behaving or regulating both oncogenes and tumor suppressor genes [161]. Which of these two paths is chosen depends on the spatial–temporal context: the type of cell/tissue, the stage the cell is at and the microenvironment surrounding the tumor [162].
Now, we are becoming aware of the complex control enforced after the point of transcription. RBPs play a significant role in this regulation, which allows coordinated remodeling of the transcriptome and proteome in response to microenvironmental influences.
It has long been assumed that a protein’s activity is mostly regulated and influenced by interactions with other proteins: however, it is also possible that this situation could also be influenced by its interaction with RNA, in the case of RBPs.
Additionally, developments in sequencing technology and novel techniques to decipher the RBP landscape are improving our understanding of the key players involved in gene expression, translation and post-transcriptional regulation mechanisms. Genetic, post-transcriptional or post-translational modifications to a single RBP may have an impact on a variety of RNA targets but also at multiple points of their RNA metabolism. Thus, analyzing these protein–RNA structural interactions will be enlightening.
New approaches for therapeutic targeting are now emerging from a better understanding of how RBPs affect this regulatory network in healthy and malignant hematopoiesis. These strategies may focus on the RBP itself, its RNA interaction, the up/downstream alterations to the proteome brought on by changes in RBP function, or any combination of these possibilities. For instance, new therapeutic strategies target the RBPs themselves in hematological malignancies, such as the use of XPO1 inhibitors for both myeloma (e.g., STORM clinical trial, ID: NCT02336815) and lymphoma (ID: NCT05422066) [163], MSI2 inhibitors for AML (ID: NCT01546038) [138], or SRSF2 inhibitors for MDS [164]; but also the inhibition of their interaction and/or downregulation of their targets comprise a new therapeutic horizon, such as the targeting of HNRNPK and c-MYC [22].
Although there is an increasing development of new drugs targeting RBPs and new therapeutic strategies are being discovered, there are still many questions that need to be answered. There are multiple dilemmas regarding the use of RBPs inhibitors. For instance, many RBPs act as master regulators in cancer, behaving as both oncogenes and tumor suppressors; most RBPs have multiple locations in the cell related with their function and a promiscuous nature with multiple targets and functions. All of these can lead to unspecificity, side effects and toxicity. Thus, we need to understand the physiological implications of altered RBPs, which type of complexes they form and their dynamics, what is the PTM’s role on altered RBPs and what kind of structural information is needed to target RBPs (isolated RBP versus RBP-RNA complex). Nevertheless, these are similar disadvantages to the traditional chemotherapy, which is still very effective and kept as the current frontline treatment for many cancers.

6. Conclusions

In conclusion, we highlight how crucial it is to comprehend the regulation of RBPs, which are emerging biomarkers of hematological malignancies that behave as master switches in cancer, aging and hematopoiesis, in order to identify new targets and pathways that drive or contribute to hematological diseases, neoplasms and dysplasia. In this context, future research will be essential in identifying relevant RBPs, their targets and their working network.

Author Contributions

Conceptualization, M.V.-E. and M.G.; writing—original draft preparation, P.A.-G. and Á.O.-S.; writing—review and editing, M.Á.N.-A., M.V.-E. and M.G.; supervision, M.V.-E. and M.G.; project administration, M.V.-E. and M.G.; funding acquisition, M.V.-E. and M.G. All authors have read and agreed to the published version of the manuscript.

Funding

M.G.: This study has been funded by Instituto de Salud Carlos III (ISCIII) through the project “PI21/00191” and co-funded by the European Union. This study has been funded by Instituto de Salud Carlos III through the funding “CP19/00140“and project “PI18/00295“(Co-funded by European Regional Development Fund/European Social Fund “A way to make Europe/Investing in your future”. This study has been funded by CRIS contra el Cancer Foundation. This study has been funded by AECC Accelerator project from “Asociación Española contra el Cancer (AECC). P.A.: This study has been funded by FEHH grant 2021. M.V.E.: “This project has received funding from the European Union’s Horizon 2020 research and innovation programme under the Marie Sklodowska-Curie grant agreement No 101027864”.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Swerdlow, S.C.E.; Harris, N.; Jaffe, E.; Pileri, S.; Stein, H.; Thiele, J.; Vardiman, J. WHO Classification of Tumours of Haematopoietic and Lymphoid Tissues, 4th ed.; World Health Organization: Geneva, Switzerland, 2017.
  2. Global Cancer Observatory. Available online: https://gco.iarc.fr/ (accessed on 6 August 2022).
  3. Keykhaei, M.; Masinaei, M.; Mohammadi, E.; Azadnajafabad, S.; Rezaei, N.; Saeedi Moghaddam, S.; Rezaei, N.; Nasserinejad, M.; Abbasi-Kangevari, M.; Malekpour, M.R.; et al. A global, regional, and national survey on burden and Quality of Care Index (QCI) of hematologic malignancies; global burden of disease systematic analysis 1990–2017. Exp. Hematol. Oncol. 2021, 10, 11. [Google Scholar] [CrossRef] [PubMed]
  4. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer statistics, 2022. CA Cancer J. Clin. 2022, 72, 7–33. [Google Scholar] [CrossRef]
  5. Pulte, D.; Jansen, L.; Brenner, H. Changes in long term survival after diagnosis with common hematologic malignancies in the early 21st century. Blood Cancer J. 2020, 10, 56. [Google Scholar] [CrossRef] [PubMed]
  6. Yu, B.; Liu, D. Antibody-drug conjugates in clinical trials for lymphoid malignancies and multiple myeloma. J. Hematol. Oncol. 2019, 12, 94. [Google Scholar] [CrossRef] [PubMed]
  7. Tian, Z.; Liu, M.; Zhang, Y.; Wang, X. Bispecific T cell engagers: An emerging therapy for management of hematologic malignancies. J. Hematol. Oncol. 2021, 14, 75. [Google Scholar] [CrossRef]
  8. Allahyari, H.; Heidari, S.; Ghamgosha, M.; Saffarian, P.; Amani, J. Immunotoxin: A new tool for cancer therapy. Tumour Biol. J. Int. Soc. Oncodev. Biol. Med. 2017, 39, 1010428317692226. [Google Scholar] [CrossRef] [Green Version]
  9. Salik, B.; Smyth, M.J.; Nakamura, K. Targeting immune checkpoints in hematological malignancies. J. Hematol. Oncol. 2020, 13, 111. [Google Scholar] [CrossRef]
  10. Gill, S.; Brudno, J.N. CAR T-Cell Therapy in Hematologic Malignancies: Clinical Role, Toxicity, and Unanswered Questions. Am. Soc. Clin. Oncol. Educ. Book. Am. Soc. Clin. Oncol. Annu. Meet. 2021, 41, 1–20. [Google Scholar] [CrossRef]
  11. Lamb, M.G.; Rangarajan, H.G.; Tullius, B.P.; Lee, D.A. Natural killer cell therapy for hematologic malignancies: Successes, challenges, and the future. Stem Cell Res. Ther. 2021, 12, 211. [Google Scholar] [CrossRef]
  12. Pinto, V.; Bergantim, R.; Caires, H.R.; Seca, H.; Guimaraes, J.E.; Vasconcelos, M.H. Multiple Myeloma: Available Therapies and Causes of Drug Resistance. Cancers 2020, 12, 407. [Google Scholar] [CrossRef] [Green Version]
  13. Zhang, J.; Gu, Y.; Chen, B. Mechanisms of drug resistance in acute myeloid leukemia. OncoTargets Ther. 2019, 12, 1937–1945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Camicia, R.; Winkler, H.C.; Hassa, P.O. Novel drug targets for personalized precision medicine in relapsed/refractory diffuse large B-cell lymphoma: A comprehensive review. Mol. Cancer 2015, 14, 207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Glisovic, T.; Bachorik, J.L.; Yong, J.; Dreyfuss, G. RNA-binding proteins and post-transcriptional gene regulation. FEBS Lett. 2008, 582, 1977–1986. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Hodson, D.J.; Screen, M.; Turner, M. RNA-binding proteins in hematopoiesis and hematological malignancy. Blood 2019, 133, 2365–2373. [Google Scholar] [CrossRef] [Green Version]
  17. Zandhuis, N.D.; Nicolet, B.P.; Wolkers, M.C. RNA-Binding Protein Expression Alters Upon Differentiation of Human B Cells and T Cells. Front. Immunol. 2021, 12, 717324. [Google Scholar] [CrossRef]
  18. Feng, M.; Xie, X.; Han, G.; Zhang, T.; Li, Y.; Li, Y.; Yin, R.; Wang, Q.; Zhang, T.; Wang, P.; et al. YBX1 is required for maintaining myeloid leukemia cell survival by regulating BCL2 stability in an m6A-dependent manner. Blood 2021, 138, 71–85. [Google Scholar] [CrossRef]
  19. Lovci, M.T.; Bengtson, M.H.; Massirer, K.B. Post-Translational Modifications and RNA-Binding Proteins. Adv. Exp. Med. Biol. 2016, 907, 297–317. [Google Scholar] [CrossRef]
  20. Carotenuto, P.; Pecoraro, A.; Palma, G.; Russo, G.; Russo, A. Therapeutic Approaches Targeting Nucleolus in Cancer. Cells 2019, 8, 1090. [Google Scholar] [CrossRef] [Green Version]
  21. Gallardo, M.; Lee, H.J.; Zhang, X.; Bueso-Ramos, C.; Pageon, L.R.; McArthur, M.; Multani, A.; Nazha, A.; Manshouri, T.; Parker-Thornburg, J.; et al. hnRNP K Is a Haploinsufficient Tumor Suppressor that Regulates Proliferation and Differentiation Programs in Hematologic Malignancies. Cancer Cell 2015, 28, 486–499. [Google Scholar] [CrossRef] [Green Version]
  22. Gallardo, M.; Malaney, P.; Aitken, M.J.L.; Zhang, X.; Link, T.M.; Shah, V.; Alybayev, S.; Wu, M.H.; Pageon, L.R.; Ma, H.; et al. Uncovering the Role of RNA-Binding Protein hnRNP K in B-Cell Lymphomas. J. Natl. Cancer Inst. 2020, 112, 95–106. [Google Scholar] [CrossRef]
  23. 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] [PubMed]
  24. 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]
  25. 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. A large-scale binding and functional map of human RNA-binding proteins. Nature 2020, 583, 711–719. [Google Scholar] [CrossRef] [PubMed]
  26. Lukong, K.E.; Chang, K.W.; Khandjian, E.W.; Richard, S. RNA-binding proteins in human genetic disease. Trends Genet. TIG 2008, 24, 416–425. [Google Scholar] [CrossRef]
  27. Valverde, R.; Edwards, L.; Regan, L. Structure and function of KH domains. FEBS J. 2008, 275, 2712–2726. [Google Scholar] [CrossRef]
  28. Du, Z.; Lee, J.K.; Fenn, S.; Tjhen, R.; Stroud, R.M.; James, T.L. X-ray crystallographic and NMR studies of protein-protein and protein-nucleic acid interactions involving the KH domains from human poly(C)-binding protein-2. RNA 2007, 13, 1043–1051. [Google Scholar] [CrossRef] [Green Version]
  29. Bomsztyk, K.; Denisenko, O.; Ostrowski, J. hnRNP K: One protein multiple processes. BioEssays News Rev. Mol. Cell. Dev. Biol. 2004, 26, 629–638. [Google Scholar] [CrossRef]
  30. Shibata, R.; Bessho, Y.; Shinkai, A.; Nishimoto, M.; Fusatomi, E.; Terada, T.; Shirouzu, M.; Yokoyama, S. Crystal structure and RNA-binding analysis of the archaeal transcription factor NusA. Biochem. Biophys. Res. Commun. 2007, 355, 122–128. [Google Scholar] [CrossRef]
  31. Daubner, G.M.; Brummer, A.; Tocchini, C.; Gerhardy, S.; Ciosk, R.; Zavolan, M.; Allain, F.H. Structural and functional implications of the QUA2 domain on RNA recognition by GLD-1. Nucleic Acids Res. 2014, 42, 8092–8105. [Google Scholar] [CrossRef] [Green Version]
  32. Teplova, M.; Malinina, L.; Darnell, J.C.; Song, J.; Lu, M.; Abagyan, R.; Musunuru, K.; Teplov, A.; Burley, S.K.; Darnell, R.B.; et al. Protein-RNA and protein-protein recognition by dual KH1/2 domains of the neuronal splicing factor Nova-1. Structure 2011, 19, 930–944. [Google Scholar] [CrossRef] [Green Version]
  33. Lewis, H.A.; Musunuru, K.; Jensen, K.B.; Edo, C.; Chen, H.; Darnell, R.B.; Burley, S.K. Sequence-specific RNA binding by a Nova KH domain: Implications for paraneoplastic disease and the fragile X syndrome. Cell 2000, 100, 323–332. [Google Scholar] [CrossRef] [Green Version]
  34. Baber, J.L.; Libutti, D.; Levens, D.; Tjandra, N. High precision solution structure of the C-terminal KH domain of heterogeneous nuclear ribonucleoprotein K, a c-myc transcription factor. J. Mol. Biol. 1999, 289, 949–962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. 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]
  36. Agrawal, A.A.; Salsi, E.; Chatrikhi, R.; Henderson, S.; Jenkins, J.L.; Green, M.R.; Ermolenko, D.N.; Kielkopf, C.L. An extended U2AF(65)-RNA-binding domain recognizes the 3’ splice site signal. Nat. Commun. 2016, 7, 10950. [Google Scholar] [CrossRef] [Green Version]
  37. Okuwaki, M.; Saotome-Nakamura, A.; Yoshimura, M.; Saito, S.; Hirawake-Mogi, H.; Sekiya, T.; Nagata, K. RNA-recognition motifs and glycine and arginine-rich region cooperatively regulate the nucleolar localization of nucleolin. J. Biochem. 2021, 169, 87–100. [Google Scholar] [CrossRef] [PubMed]
  38. Hargous, Y.; Hautbergue, G.M.; Tintaru, A.M.; Skrisovska, L.; Golovanov, A.P.; Stevenin, J.; Lian, L.Y.; Wilson, S.A.; Allain, F.H. Molecular basis of RNA recognition and TAP binding by the SR proteins SRp20 and 9G8. EMBO J. 2006, 25, 5126–5137. [Google Scholar] [CrossRef] [Green Version]
  39. Dominguez, C.; Allain, F.H. NMR structure of the three quasi RNA recognition motifs (qRRMs) of human hnRNP F and interaction studies with Bcl-x G-tract RNA: A novel mode of RNA recognition. Nucleic Acids Res. 2006, 34, 3634–3645. [Google Scholar] [CrossRef]
  40. Krepl, M.; Blatter, M.; Clery, A.; Damberger, F.F.; Allain, F.H.T.; Sponer, J. Structural study of the Fox-1 RRM protein hydration reveals a role for key water molecules in RRM-RNA recognition. Nucleic Acids Res. 2017, 45, 8046–8063. [Google Scholar] [CrossRef] [Green Version]
  41. Zearfoss, N.R.; Deveau, L.M.; Clingman, C.C.; Schmidt, E.; Johnson, E.S.; Massi, F.; Ryder, S.P. A conserved three-nucleotide core motif defines Musashi RNA binding specificity. J. Biol. Chem. 2014, 289, 35530–35541. [Google Scholar] [CrossRef] [Green Version]
  42. Bussiere, D.E.; Xie, L.; Srinivas, H.; Shu, W.; Burke, A.; Be, C.; Zhao, J.; Godbole, A.; King, D.; Karki, R.G.; et al. Structural basis of indisulam-mediated RBM39 recruitment to DCAF15 E3 ligase complex. Nat. Chem. Biol. 2020, 16, 15–23. [Google Scholar] [CrossRef]
  43. Berg, J.M.; Godwin, H.A. Lessons from zinc-binding peptides. Annu. Rev. Biophys. Biomol. Struct. 1997, 26, 357–371. [Google Scholar] [CrossRef] [PubMed]
  44. Hale, M.A.; Richardson, J.I.; Day, R.C.; McConnell, O.L.; Arboleda, J.; Wang, E.T.; Berglund, J.A. An engineered RNA binding protein with improved splicing regulation. Nucleic Acids Res. 2018, 46, 3152–3168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Lu, D.; Searles, M.A.; Klug, A. Crystal structure of a zinc-finger-RNA complex reveals two modes of molecular recognition. Nature 2003, 426, 96–100. [Google Scholar] [CrossRef] [PubMed]
  46. Loughlin, F.E.; Mansfield, R.E.; Vaz, P.M.; McGrath, A.P.; Setiyaputra, S.; Gamsjaeger, R.; Chen, E.S.; Morris, B.J.; Guss, J.M.; Mackay, J.P. The zinc fingers of the SR-like protein ZRANB2 are single-stranded RNA-binding domains that recognize 5’ splice site-like sequences. Proc. Natl. Acad. Sci. USA 2009, 106, 5581–5586. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Banerjee, S.; Barraud, P. Functions of double-stranded RNA-binding domains in nucleocytoplasmic transport. RNA Biol. 2014, 11, 1226–1232. [Google Scholar] [CrossRef] [Green Version]
  48. Brooks, R.; Eckmann, C.R.; Jantsch, M.F. The double-stranded RNA-binding domains of Xenopus laevis ADAR1 exhibit different RNA-binding behaviors. FEBS Lett. 1998, 434, 121–126. [Google Scholar] [CrossRef] [Green Version]
  49. Doyle, M.; Badertscher, L.; Jaskiewicz, L.; Guttinger, S.; Jurado, S.; Hugenschmidt, T.; Kutay, U.; Filipowicz, W. The double-stranded RNA binding domain of human Dicer functions as a nuclear localization signal. RNA 2013, 19, 1238–1252. [Google Scholar] [CrossRef] [Green Version]
  50. Wang, S.; Hu, Y.; Overgaard, M.T.; Karginov, F.V.; Uhlenbeck, O.C.; McKay, D.B. The domain of the Bacillus subtilis DEAD-box helicase YxiN that is responsible for specific binding of 23S rRNA has an RNA recognition motif fold. RNA 2006, 12, 959–967. [Google Scholar] [CrossRef] [Green Version]
  51. Lorsch, J.R.; Herschlag, D. The DEAD box protein eIF4A. 2. A cycle of nucleotide and RNA-dependent conformational changes. Biochemistry 1998, 37, 2194–2206. [Google Scholar] [CrossRef]
  52. Bates, G.J.; Nicol, S.M.; Wilson, B.J.; Jacobs, A.M.; Bourdon, J.C.; Wardrop, J.; Gregory, D.J.; Lane, D.P.; Perkins, N.D.; Fuller-Pace, F.V. The DEAD box protein p68: A novel transcriptional coactivator of the p53 tumour suppressor. EMBO J. 2005, 24, 543–553. [Google Scholar] [CrossRef] [Green Version]
  53. Gustafson, E.A.; Wessel, G.M. DEAD-box helicases: Posttranslational regulation and function. Biochem. Biophys. Res. Commun. 2010, 395, 1–6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Wang, X.; Zamore, P.D.; Hall, T.M. Crystal structure of a Pumilio homology domain. Mol. Cell 2001, 7, 855–865. [Google Scholar] [CrossRef]
  55. Wang, M.; Oge, L.; Perez-Garcia, M.D.; Hamama, L.; Sakr, S. The PUF Protein Family: Overview on PUF RNA Targets, Biological Functions, and Post Transcriptional Regulation. Int. J. Mol. Sci. 2018, 19, 410. [Google Scholar] [CrossRef] [Green Version]
  56. Oberstrass, F.C.; Lee, A.; Stefl, R.; Janis, M.; Chanfreau, G.; Allain, F.H. Shape-specific recognition in the structure of the Vts1p SAM domain with RNA. Nat. Struct. Mol. Biol. 2006, 13, 160–167. [Google Scholar] [CrossRef] [PubMed]
  57. Ou, H.D.; Lohr, F.; Vogel, V.; Mantele, W.; Dotsch, V. Structural evolution of C-terminal domains in the p53 family. EMBO J. 2007, 26, 3463–3473. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Lee, H.J.; Hota, P.K.; Chugha, P.; Guo, H.; Miao, H.; Zhang, L.; Kim, S.J.; Stetzik, L.; Wang, B.C.; Buck, M. NMR structure of a heterodimeric SAM:SAM complex: Characterization and manipulation of EphA2 binding reveal new cellular functions of SHIP2. Structure 2012, 20, 41–55. [Google Scholar] [CrossRef] [Green Version]
  59. Fu, X.D.; Ares, M., Jr. Context-dependent control of alternative splicing by RNA-binding proteins. Nat. Rev. Genet. 2014, 15, 689–701. [Google Scholar] [CrossRef]
  60. Muller-McNicoll, M.; Neugebauer, K.M. How cells get the message: Dynamic assembly and function of mRNA-protein complexes. Nat. Rev. Genet. 2013, 14, 275–287. [Google Scholar] [CrossRef]
  61. Keene, J.D. Minireview: Global regulation and dynamics of ribonucleic Acid. Endocrinology 2010, 151, 1391–1397. [Google Scholar] [CrossRef] [Green Version]
  62. Raghavan, A.; Ogilvie, R.L.; Reilly, C.; Abelson, M.L.; Raghavan, S.; Vasdewani, J.; Krathwohl, M.; Bohjanen, P.R. Genome-wide analysis of mRNA decay in resting and activated primary human T lymphocytes. Nucleic Acids Res. 2002, 30, 5529–5538. [Google Scholar] [CrossRef]
  63. Gebauer, F.; Hentze, M.W. Molecular mechanisms of translational control. Nat. Rev. Mol. Cell Biol. 2004, 5, 827–835. [Google Scholar] [CrossRef] [PubMed]
  64. Muckenthaler, M.; Gray, N.K.; Hentze, M.W. IRP-1 binding to ferritin mRNA prevents the recruitment of the small ribosomal subunit by the cap-binding complex eIF4F. Mol. Cell 1998, 2, 383–388. [Google Scholar] [CrossRef]
  65. Neve, J.; Patel, R.; Wang, Z.; Louey, A.; Furger, A.M. Cleavage and polyadenylation: Ending the message expands gene regulation. RNA Biol. 2017, 14, 865–890. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Flaherty, S.M.; Fortes, P.; Izaurralde, E.; Mattaj, I.W.; Gilmartin, G.M. Participation of the nuclear cap binding complex in pre-mRNA 3’ processing. Proc. Natl. Acad. Sci. USA 1997, 94, 11893–11898. [Google Scholar] [CrossRef] [Green Version]
  67. Bauren, G.; Belikov, S.; Wieslander, L. Transcriptional termination in the Balbiani ring 1 gene is closely coupled to 3’-end formation and excision of the 3’-terminal intron. Genes Dev. 1998, 12, 2759–2769. [Google Scholar] [CrossRef] [Green Version]
  68. Schoenberg, D.R.; Maquat, L.E. Re-capping the message. Trends Biochem. Sci. 2009, 34, 435–442. [Google Scholar] [CrossRef]
  69. Schwer, B.; Mao, X.; Shuman, S. Accelerated mRNA decay in conditional mutants of yeast mRNA capping enzyme. Nucleic Acids Res. 1998, 26, 2050–2057. [Google Scholar] [CrossRef] [Green Version]
  70. Gingras, A.C.; Raught, B.; Sonenberg, N. eIF4 initiation factors: Effectors of mRNA recruitment to ribosomes and regulators of translation. Annu. Rev. Biochem. 1999, 68, 913–963. [Google Scholar] [CrossRef]
  71. Wahl, M.C.; Will, C.L.; Luhrmann, R. The spliceosome: Design principles of a dynamic RNP machine. Cell 2009, 136, 701–718. [Google Scholar] [CrossRef] [Green Version]
  72. Han, J.; Xiong, J.; Wang, D.; Fu, X.D. Pre-mRNA splicing: Where and when in the nucleus. Trends Cell Biol. 2011, 21, 336–343. [Google Scholar] [CrossRef]
  73. Perales, R.; Bentley, D. “Cotranscriptionality”: The transcription elongation complex as a nexus for nuclear transactions. Mol. Cell 2009, 36, 178–191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Cooper, T.A.; Wan, L.; Dreyfuss, G. RNA and disease. Cell 2009, 136, 777–793. [Google Scholar] [CrossRef] [Green Version]
  75. Proudfoot, N.J.; Furger, A.; Dye, M.J. Integrating mRNA processing with transcription. Cell 2002, 108, 501–512. [Google Scholar] [CrossRef] [Green Version]
  76. Proudfoot, N. New perspectives on connecting messenger RNA 3’ end formation to transcription. Curr. Opin. Cell Biol. 2004, 16, 272–278. [Google Scholar] [CrossRef] [PubMed]
  77. Ohno, M. Size matters in RNA export. RNA Biol. 2012, 9, 1413–1417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. McCloskey, A.; Taniguchi, I.; Shinmyozu, K.; Ohno, M. hnRNP C tetramer measures RNA length to classify RNA polymerase II transcripts for export. Science 2012, 335, 1643–1646. [Google Scholar] [CrossRef]
  79. Stewart, M. Ratcheting mRNA out of the nucleus. Mol. Cell 2007, 25, 327–330. [Google Scholar] [CrossRef]
  80. Wu, X.; Brewer, G. The regulation of mRNA stability in mammalian cells: 2.0. Gene 2012, 500, 10–21. [Google Scholar] [CrossRef] [Green Version]
  81. Balagopal, V.; Fluch, L.; Nissan, T. Ways and means of eukaryotic mRNA decay. Biochim. Biophys. Acta 2012, 1819, 593–603. [Google Scholar] [CrossRef] [Green Version]
  82. Vasudevan, S.; Tong, Y.; Steitz, J.A. Switching from repression to activation: microRNAs can up-regulate translation. Science 2007, 318, 1931–1934. [Google Scholar] [CrossRef] [Green Version]
  83. Sheets, M.D.; Fox, C.A.; Hunt, T.; Vande Woude, G.; Wickens, M. The 3’-untranslated regions of c-mos and cyclin mRNAs stimulate translation by regulating cytoplasmic polyadenylation. Genes Dev. 1994, 8, 926–938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Iguchi, N.; Tobias, J.W.; Hecht, N.B. Expression profiling reveals meiotic male germ cell mRNAs that are translationally up- and down-regulated. Proc. Natl. Acad. Sci. USA 2006, 103, 7712–7717. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Quinones-Valdez, G.; Tran, S.S.; Jun, H.I.; Bahn, J.H.; Yang, E.W.; Zhan, L.; Brümmer, A.; Wei, X.; Van Nostrand, E.L.; Prat, G.A.; et al. Regulation of RNA editing by RNA-binding proteins in human cells. Commun. Biol. 2019, 2, 19. [Google Scholar] [CrossRef] [Green Version]
  86. Brooke, G.N.; Culley, R.L.; Dart, D.A.; Mann, D.J.; Gaughan, L.; McCracken, S.R.; Robson, C.N.; Spencer-Dene, B.; Gamble, S.C.; Powell, S.M.; et al. FUS/TLS is a novel mediator of androgen-dependent cell-cycle progression and prostate cancer growth. Cancer Res. 2011, 71, 914–924. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Ortiz-Zapater, E.; Pineda, D.; Martinez-Bosch, N.; Fernandez-Miranda, G.; Iglesias, M.; Alameda, F.; Moreno, M.; Eliscovich, C.; Eyras, E.; Real, F.X.; et al. Key contribution of CPEB4-mediated translational control to cancer progression. Nat. Med. 2011, 18, 83–90. [Google Scholar] [CrossRef]
  88. 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]
  89. Wurth, L.; Papasaikas, P.; Olmeda, D.; Bley, N.; Calvo, G.T.; Guerrero, S.; Cerezo-Wallis, D.; Martinez-Useros, J.; Garcia-Fernandez, M.; Huttelmaier, S.; et al. UNR/CSDE1 Drives a Post-transcriptional Program to Promote Melanoma Invasion and Metastasis. Cancer Cell 2019, 36, 337. [Google Scholar] [CrossRef] [Green Version]
  90. Qi, Z.; Takamatsu, H.; Espinoza, J.L.; Lu, X.; Sugimori, N.; Yamazaki, H.; Okawa, K.; Nakao, S. Autoantibodies specific to hnRNP K: A new diagnostic marker for immune pathophysiology in aplastic anemia. Ann. Hematol. 2010, 89, 1255–1263. [Google Scholar] [CrossRef]
  91. Aguilar-Garrido, P.V.M.; Hernández Sánchez, M.; Navarro Aguadero, M.Á.; Malaney, P.; Aitken, M.J.L.; Zhang, X.; Young, K.H.; Duan, R.; Hu, P.; Kornblau, S.; et al. S104: RBPs dysregulation cause hyper-nucleoli and ribosome gain-of-function driving bone marrow failure. HemaSphere 2022, 6, 5–6. [Google Scholar] [CrossRef]
  92. Jumper, J.; Evans, R.; Pritzel, A.; Green, T.; Figurnov, M.; Ronneberger, O.; Tunyasuvunakool, K.; Bates, R.; Žídek, A.; Potapenko, A.; et al. Highly accurate protein structure prediction with AlphaFold. Nature 2021, 596, 583–589. [Google Scholar] [CrossRef]
  93. Varadi, M.; Anyango, S.; Deshpande, M.; Nair, S.; Natassia, C.; Yordanova, G.; Yuan, D.; Stroe, O.; Wood, G.; Laydon, A.; et al. AlphaFold Protein Structure Database: Massively expanding the structural coverage of protein-sequence space with high-accuracy models. Nucleic Acids Res. 2022, 50, D439–D444. [Google Scholar] [CrossRef] [PubMed]
  94. Grech, G.; von Lindern, M. The Role of Translation Initiation Regulation in Haematopoiesis. Comp. Funct. Genom. 2012, 2012, 576540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Barboro, P.; Ferrari, N.; Balbi, C. Emerging roles of heterogeneous nuclear ribonucleoprotein K (hnRNP K) in cancer progression. Cancer Lett. 2014, 352, 152–159. [Google Scholar] [CrossRef] [PubMed]
  96. Malaney, P.; Velasco-Estevez, M.; Aguilar-Garrido, P.; Aitken, M.J.L.; Chan, L.E.; Zhang, X.; Post, S.M.; Gallardo, M. The Emicro-hnRNP K Murine Model of Lymphoma: Novel Insights into the Role of hnRNP K in B-Cell Malignancies. Front. Immunol. 2021, 12, 634584. [Google Scholar] [CrossRef]
  97. Moumen, A.; Masterson, P.; O’Connor, M.J.; Jackson, S.P. hnRNP K: An HDM2 target and transcriptional coactivator of p53 in response to DNA damage. Cell 2005, 123, 1065–1078. [Google Scholar] [CrossRef] [Green Version]
  98. Yang, X.; Wen, Y.; Liu, S.; Duan, L.; Liu, T.; Tong, Z.; Wang, Z.; Gu, Y.; Xi, Y.; Wang, X.; et al. LCDR regulates the integrity of lysosomal membrane by hnRNP K-stabilized LAPTM5 transcript and promotes cell survival. Proc. Natl. Acad. Sci. USA 2022, 119, e2110428119. [Google Scholar] [CrossRef]
  99. Grimwade, D.; Walker, H.; Oliver, F.; Wheatley, K.; Harrison, C.; Harrison, G.; Rees, J.; Hann, I.; Stevens, R.; Burnett, A.; et al. The importance of diagnostic cytogenetics on outcome in AML: Analysis of 1,612 patients entered into the MRC AML 10 trial. The Medical Research Council Adult and Children’s Leukaemia Working Parties. Blood 1998, 92, 2322–2333. [Google Scholar] [CrossRef] [Green Version]
  100. Dayyani, F.; Wang, J.; Yeh, J.R.; Ahn, E.Y.; Tobey, E.; Zhang, D.E.; Bernstein, I.D.; Peterson, R.T.; Sweetser, D.A. Loss of TLE1 and TLE4 from the del(9q) commonly deleted region in AML cooperates with AML1-ETO to affect myeloid cell proliferation and survival. Blood 2008, 111, 4338–4347. [Google Scholar] [CrossRef] [Green Version]
  101. Kronke, J.; Bullinger, L.; Teleanu, V.; Tschurtz, F.; Gaidzik, V.I.; Kuhn, M.W.; Rucker, F.G.; Holzmann, K.; Paschka, P.; Kapp-Schworer, S.; et al. Clonal evolution in relapsed NPM1-mutated acute myeloid leukemia. Blood 2013, 122, 100–108. [Google Scholar] [CrossRef] [Green Version]
  102. Nika, E.; Brugnoli, F.; Piazzi, M.; Lambertini, E.; Grassilli, S.; Bavelloni, A.; Piva, R.; Capitani, S.; Bertagnolo, V. hnRNP K in PU.1-containing complexes recruited at the CD11b promoter: A distinct role in modulating granulocytic and monocytic differentiation of AML-derived cells. Biochem. J. 2014, 463, 115–122. [Google Scholar] [CrossRef]
  103. Eiring, A.M.; Neviani, P.; Santhanam, R.; Oaks, J.J.; Chang, J.S.; Notari, M.; Willis, W.; Gambacorti-Passerini, C.; Volinia, S.; Marcucci, G.; et al. Identification of novel posttranscriptional targets of the BCR/ABL oncoprotein by ribonomics: Requirement of E2F3 for BCR/ABL leukemogenesis. Blood 2008, 111, 816–828. [Google Scholar] [CrossRef] [PubMed]
  104. Notari, M.; Neviani, P.; Santhanam, R.; Blaser, B.W.; Chang, J.S.; Galietta, A.; Willis, A.E.; Roy, D.C.; Caligiuri, M.A.; Marcucci, G.; et al. A MAPK/HNRPK pathway controls BCR/ABL oncogenic potential by regulating MYC mRNA translation. Blood 2006, 107, 2507–2516. [Google Scholar] [CrossRef] [PubMed]
  105. Taggart, J.; Ho, T.C.; Amin, E.; Xu, H.; Barlowe, T.S.; Perez, A.R.; Durham, B.H.; Tivnan, P.; Okabe, R.; Chow, A.; et al. MSI2 is required for maintaining activated myelodysplastic syndrome stem cells. Nat. Commun. 2016, 7, 10739. [Google Scholar] [CrossRef] [Green Version]
  106. Mancarella, C.; Scotlandi, K. IGF2BP3 From Physiology to Cancer: Novel Discoveries, Unsolved Issues, and Future Perspectives. Front. Cell Dev. Biol. 2019, 7, 363. [Google Scholar] [CrossRef] [PubMed]
  107. Palanichamy, J.K.; Tran, T.M.; Howard, J.M.; Contreras, J.R.; Fernando, T.R.; Sterne-Weiler, T.; Katzman, S.; Toloue, M.; Yan, W.; Basso, G.; et al. RNA-binding protein IGF2BP3 targeting of oncogenic transcripts promotes hematopoietic progenitor proliferation. J. Clin. Investig. 2016, 126, 1495–1511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Makinen, A.; Nikkila, A.; Haapaniemi, T.; Oksa, L.; Mehtonen, J.; Vanska, M.; Heinaniemi, M.; Paavonen, T.; Lohi, O. IGF2BP3 Associates with Proliferative Phenotype and Prognostic Features in B-Cell Acute Lymphoblastic Leukemia. Cancers 2021, 13, 1505. [Google Scholar] [CrossRef] [PubMed]
  109. Sakakibara, S.; Nakamura, Y.; Satoh, H.; Okano, H. Rna-binding protein Musashi2: Developmentally regulated expression in neural precursor cells and subpopulations of neurons in mammalian CNS. J. Neurosci. Off. J. Soc. Neurosci. 2001, 21, 8091–8107. [Google Scholar] [CrossRef] [Green Version]
  110. Nakamura, M.; Okano, H.; Blendy, J.A.; Montell, C. Musashi, a neural RNA-binding protein required for Drosophila adult external sensory organ development. Neuron 1994, 13, 67–81. [Google Scholar] [CrossRef]
  111. Kharas, M.G.; Lengner, C.J.; Al-Shahrour, F.; Bullinger, L.; Ball, B.; Zaidi, S.; Morgan, K.; Tam, W.; Paktinat, M.; Okabe, R.; et al. Musashi-2 regulates normal hematopoiesis and promotes aggressive myeloid leukemia. Nat. Med. 2010, 16, 903–908. [Google Scholar] [CrossRef] [Green Version]
  112. Xu, C.; Chen, X.; Zhang, X.; Zhao, D.; Dou, Z.; Xie, X.; Li, H.; Yang, H.; Li, Q.; Zhang, H.; et al. RNA-binding protein 39: A promising therapeutic target for cancer. Cell Death Discov. 2021, 7, 214. [Google Scholar] [CrossRef]
  113. Xu, Y.; Nijhuis, A.; Keun, H.C. RNA-binding motif protein 39 (RBM39): An emerging cancer target. Br. J. Pharmacol. 2022, 179, 2795–2812. [Google Scholar] [CrossRef] [PubMed]
  114. Mercier, I.; Gonzales, D.M.; Quann, K.; Pestell, T.G.; Molchansky, A.; Sotgia, F.; Hulit, J.; Gandara, R.; Wang, C.; Pestell, R.G.; et al. CAPER, a novel regulator of human breast cancer progression. Cell Cycle 2014, 13, 1256–1264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Komeno, Y.; Huang, Y.J.; Qiu, J.; Lin, L.; Xu, Y.; Zhou, Y.; Chen, L.; Monterroza, D.D.; Li, H.; DeKelver, R.C.; et al. SRSF2 is Essential for Hematopoiesis, and its Myeloidysplastic Syndrome-Related Mutations Dysregulate Alternatie pre-mRNA Splicing. Mol. Cell Biol. 2015, 35, 3071–3082. [Google Scholar] [CrossRef] [Green Version]
  116. Dolashad, H.; Pellagati, A.; Fernandez-Mercado, M.; Yip, B.H.; Malcovati, L.; Attwood, M.; Przychodzen, B.; Sahgal, N.; Kanapin, A.A.; Lockstone, H.; et al. Disruption of SF3B1 results in deregulated expression and splicing of key genes and pathways in myelodysplastic syndrome hematopoietic stem and progenitor cells. Leukemia 2015, 29, 1092–1103. [Google Scholar] [CrossRef]
  117. Baou, M.; Norton, J.D.; Murphy, J.J. AU-rich RNA binding proteins in hematopoiesis and leukemogenesis. Blood 2011, 118, 5732–5740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Dowhan, D.H.; Hong, E.P.; Auboeuf, D.; Dennis, A.P.; Wilson, M.M.; Berget, S.M.; O’Malley, B.W. Steroid hormone receptor coactivation and alternative RNA splicing by U2AF65-related proteins CAPERalpha and CAPERbeta. Mol. Cell 2005, 17, 429–439. [Google Scholar] [CrossRef]
  119. Jung, D.J.; Na, S.Y.; Na, D.S.; Lee, J.W. Molecular cloning and characterization of CAPER, a novel coactivator of activating protein-1 and estrogen receptors. J. Biol. Chem. 2002, 277, 1229–1234. [Google Scholar] [CrossRef] [Green Version]
  120. Song, Y.; Guo, Y.; Li, X.; Sun, R.; Zhu, M.; Shi, J.; Tan, Z.; Zhang, L.; Huang, J. RBM39 Alters Phosphorylation of c-Jun and Binds to Viral RNA to Promote PRRSV Proliferation. Front. Immunol. 2021, 12, 664417. [Google Scholar] [CrossRef]
  121. Mai, S.; Qu, X.; Li, P.; Ma, Q.; Liu, X.; Cao, C. Functional interaction between nonreceptor tyrosine kinase c-Abl and SR-Rich protein RBM39. Biochem. Biophys. Res. Commun. 2016, 473, 355–360. [Google Scholar] [CrossRef]
  122. Campbell, M.C.; Pontiggia, L.; Russell, A.Y.; Schwarting, R.; Camacho, J.; Jasmin, J.F.; Mercier, I. CAPER as a therapeutic target for triple negative breast cancer. Oncotarget 2018, 9, 30340–30354. [Google Scholar] [CrossRef]
  123. Chai, Y.; Liu, X.; Dai, L.; Li, Y.; Liu, M.; Zhang, J.Y. Overexpression of HCC1/CAPERalpha may play a role in lung cancer carcinogenesis. Tumour Biol. J. Int. Soc. Oncodev. Biol. Med. 2014, 35, 6311–6317. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Wang, E.; Lu, S.X.; Pastore, A.; Chen, X.; Imig, J.; Chun-Wei Lee, S.; Hockemeyer, K.; Ghebrechristos, Y.E.; Yoshimi, A.; Inoue, D.; et al. Targeting an RNA-Binding Protein Network in Acute Myeloid Leukemia. Cancer Cell 2019, 35, 369–384.e367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Tong, J.; Xu, X.; Zhang, Z.; Ma, C.; Xiang, R.; Liu, J.; Xu, W.; Wu, C.; Li, J.; Zhan, F.; et al. Hypoxia-induced long non-coding RNA DARS-AS1 regulates RBM39 stability to promote myeloma malignancy. Haematologica 2020, 105, 1630–1640. [Google Scholar] [CrossRef]
  126. Okano, H.; Imai, T.; Okabe, M. Musashi: A translational regulator of cell fate. J. Cell Sci. 2002, 115, 1355–1359. [Google Scholar] [CrossRef]
  127. Okano, H.; Kawahara, H.; Toriya, M.; Nakao, K.; Shibata, S.; Imai, T. Function of RNA-binding protein Musashi-1 in stem cells. Exp. Cell Res. 2005, 306, 349–356. [Google Scholar] [CrossRef]
  128. Forouzanfar, M.; Lachinani, L.; Dormiani, K.; Nasr-Esfahani, M.H.; Gure, A.O.; Ghaedi, K. Intracellular functions of RNA-binding protein, Musashi1, in stem and cancer cells. Stem Cell Res. Ther. 2020, 11, 193. [Google Scholar] [CrossRef]
  129. Wuebben, E.L.; Mallanna, S.K.; Cox, J.L.; Rizzino, A. Musashi2 is required for the self-renewal and pluripotency of embryonic stem cells. PLoS ONE 2012, 7, e34827. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Kudinov, A.E.; Karanicolas, J.; Golemis, E.A.; Boumber, Y. Musashi RNA-Binding Proteins as Cancer Drivers and Novel Therapeutic Targets. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2017, 23, 2143–2153. [Google Scholar] [CrossRef] [Green Version]
  131. Li, M.; Li, A.Q.; Zhou, S.L.; Lv, H.; Wei, P.; Yang, W.T. RNA-binding protein MSI2 isoforms expression and regulation in progression of triple-negative breast cancer. J. Exp. Clin. Cancer Res. 2020, 39, 92. [Google Scholar] [CrossRef]
  132. MacNicol, M.C.; Cragle, C.E.; McDaniel, F.K.; Hardy, L.L.; Wang, Y.; Arumugam, K.; Rahmatallah, Y.; Glazko, G.V.; Wilczynska, A.; Childs, G.V.; et al. Evasion of regulatory phosphorylation by an alternative spliced isoform of Musashi2. Sci. Rep. 2017, 7, 11503. [Google Scholar] [CrossRef] [Green Version]
  133. Kharas, M.G.; Lengner, C.J. Stem Cells, Cancer, and MUSASHI in Blood and Guts. Trends Cancer 2017, 3, 347–356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Park, S.M.; Gonen, M.; Vu, L.; Minuesa, G.; Tivnan, P.; Barlowe, T.S.; Taggart, J.; Lu, Y.; Deering, R.P.; Hacohen, N.; et al. Musashi2 sustains the mixed-lineage leukemia-driven stem cell regulatory program. J. Clin. Investig. 2015, 125, 1286–1298. [Google Scholar] [CrossRef] [Green Version]
  135. Imai, T.; Tokunaga, A.; Yoshida, T.; Hashimoto, M.; Mikoshiba, K.; Weinmaster, G.; Nakafuku, M.; Okano, H. The neural RNA-binding protein Musashi1 translationally regulates mammalian numb gene expression by interacting with its mRNA. Mol. Cell. Biol. 2001, 21, 3888–3900. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Ito, T.; Kwon, H.Y.; Zimdahl, B.; Congdon, K.L.; Blum, J.; Lento, W.E.; Zhao, C.; Lagoo, A.; Gerrard, G.; Foroni, L.; et al. Regulation of myeloid leukaemia by the cell-fate determinant Musashi. Nature 2010, 466, 765–768. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Palacios, F.; Yan, X.J.; Ferrer, G.; Chen, S.S.; Vergani, S.; Yang, X.; Gardner, J.; Barrientos, J.C.; Rock, P.; Burack, R.; et al. Musashi 2 influences chronic lymphocytic leukemia cell survival and growth making it a potential therapeutic target. Leukemia 2021, 35, 1037–1052. [Google Scholar] [CrossRef]
  138. Minuesa, G.; Albanese, S.K.; Xie, W.; Kazansky, Y.; Worroll, D.; Chow, A.; Schurer, A.; Park, S.M.; Rotsides, C.Z.; Taggart, J.; et al. Small-molecule targeting of MUSASHI RNA-binding activity in acute myeloid leukemia. Nat. Commun. 2019, 10, 2691. [Google Scholar] [CrossRef] [Green Version]
  139. Nielsen, J.; Christiansen, J.; Lykke-Andersen, J.; Johnsen, A.H.; Wewer, U.M.; Nielsen, F.C. A family of insulin-like growth factor II mRNA-binding proteins represses translation in late development. Mol. Cell. Biol. 1999, 19, 1262–1270. [Google Scholar] [CrossRef] [Green Version]
  140. Nielsen, J.; Kristensen, M.A.; Willemoes, M.; Nielsen, F.C.; Christiansen, J. Sequential dimerization of human zipcode-binding protein IMP1 on RNA: A cooperative mechanism providing RNP stability. Nucleic Acids Res. 2004, 32, 4368–4376. [Google Scholar] [CrossRef] [Green Version]
  141. 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]
  142. Schneider, T.; Hung, L.H.; Aziz, M.; Wilmen, A.; Thaum, S.; Wagner, J.; Janowski, R.; Muller, S.; Schreiner, S.; Friedhoff, P.; et al. Combinatorial recognition of clustered RNA elements by the multidomain RNA-binding protein IMP3. Nat. Commun. 2019, 10, 2266. [Google Scholar] [CrossRef] [Green Version]
  143. Bell, J.L.; Wachter, K.; Muhleck, B.; Pazaitis, N.; Kohn, M.; Lederer, M.; Huttelmaier, S. Insulin-like growth factor 2 mRNA-binding proteins (IGF2BPs): Post-transcriptional drivers of cancer progression? Cell. Mol. Life Sci. CMLS 2013, 70, 2657–2675. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Lederer, M.; Bley, N.; Schleifer, C.; Huttelmaier, S. The role of the oncofetal IGF2 mRNA-binding protein 3 (IGF2BP3) in cancer. Semin. Cancer Biol. 2014, 29, 3–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Rivera Vargas, T.; Boudoukha, S.; Simon, A.; Souidi, M.; Cuvellier, S.; Pinna, G.; Polesskaya, A. Post-transcriptional regulation of cyclins D1, D3 and G1 and proliferation of human cancer cells depend on IMP-3 nuclear localization. Oncogene 2014, 33, 2866–2875. [Google Scholar] [CrossRef] [PubMed]
  146. 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]
  147. Zhao, W.; Lu, D.; Liu, L.; Cai, J.; Zhou, Y.; Yang, Y.; Zhang, Y.; Zhang, J. Insulin-like growth factor 2 mRNA binding protein 3 (IGF2BP3) promotes lung tumorigenesis via attenuating p53 stability. Oncotarget 2017, 8, 93672–93687. [Google Scholar] [CrossRef] [Green Version]
  148. Tran, T.M.; Philipp, J.; Bassi, J.S.; Nibber, N.; Draper, J.M.; Lin, T.L.; Palanichamy, J.K.; Jaiswal, A.K.; Silva, O.; Paing, M.; et al. The RNA-binding protein IGF2BP3 is critical for MLL-AF4-mediated leukemogenesis. Leukemia 2022, 36, 68–79. [Google Scholar] [CrossRef]
  149. Canella, A.; Cordero Nieves, H.; Sborov, D.W.; Cascione, L.; Radomska, H.S.; Smith, E.; Stiff, A.; Consiglio, J.; Caserta, E.; Rizzotto, L.; et al. HDAC inhibitor AR-42 decreases CD44 expression and sensitizes myeloma cells to lenalidomide. Oncotarget 2015, 6, 31134–31150. [Google Scholar] [CrossRef] [Green Version]
  150. Gou, J.; Li, H.; Bi, J.; Pang, X.; Li, X.; Wang, Y. Transfer of IGF2BP3 Through Ara-C-Induced Apoptotic Bodies Promotes Survival of Recipient Cells. Front. Oncol. 2022, 12, 801226. [Google Scholar] [CrossRef]
  151. Makeyev, A.V.; Liebhaber, S.A. The poly(C)-binding proteins: A multiplicity of functions and a search for mechanisms. Rna 2002, 8, 265–278. [Google Scholar] [CrossRef] [Green Version]
  152. Michael, W.M.; Eder, P.S.; Dreyfuss, G. The K nuclear shuttling domain: A novel signal for nuclear import and nuclear export in the hnRNP K protein. EMBO J. 1997, 16, 3587–3598. [Google Scholar] [CrossRef] [Green Version]
  153. Moumen, A.; Magill, C.; Dry, K.L.; Jackson, S.P. ATM-dependent phosphorylation of heterogeneous nuclear ribonucleoprotein K promotes p53 transcriptional activation in response to DNA damage. Cell Cycle 2013, 12, 698–704. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Tang, F.; Li, W.; Chen, Y.; Wang, D.; Han, J.; Liu, D. Downregulation of hnRNP K by RNAi inhibits growth of human lung carcinoma cells. Oncol. Lett. 2014, 7, 1073–1077. [Google Scholar] [CrossRef] [PubMed]
  155. Li, D.; Wang, X.; Mei, H.; Fang, E.; Ye, L.; Song, H.; Yang, F.; Li, H.; Huang, K.; Zheng, L.; et al. Long Noncoding RNA pancEts-1 Promotes Neuroblastoma Progression through hnRNPK-Mediated beta-Catenin Stabilization. Cancer Res. 2018, 78, 1169–1183. [Google Scholar] [CrossRef] [Green Version]
  156. Qin, G.; Tu, X.; Li, H.; Cao, P.; Chen, X.; Song, J.; Han, H.; Li, Y.; Guo, B.; Yang, L.; et al. Long Noncoding RNA p53-Stabilizing and Activating RNA Promotes p53 Signaling by Inhibiting Heterogeneous Nuclear Ribonucleoprotein K deSUMOylation and Suppresses Hepatocellular Carcinoma. Hepatology 2020, 71, 112–129. [Google Scholar] [CrossRef]
  157. Zhang, Z.; Zhou, C.; Chang, Y.; Zhang, Z.; Hu, Y.; Zhang, F.; Lu, Y.; Zheng, L.; Zhang, W.; Li, X.; et al. Long non-coding RNA CASC11 interacts with hnRNP-K and activates the WNT/beta-catenin pathway to promote growth and metastasis in colorectal cancer. Cancer Lett. 2016, 376, 62–73. [Google Scholar] [CrossRef] [PubMed]
  158. Sweetser, D.A.; Peniket, A.J.; Haaland, C.; Blomberg, A.A.; Zhang, Y.; Zaidi, S.T.; Dayyani, F.; Zhao, Z.; Heerema, N.A.; Boultwood, J.; et al. Delineation of the minimal commonly deleted segment and identification of candidate tumor-suppressor genes in del(9q) acute myeloid leukemia. Genes Chromosomes Cancer 2005, 44, 279–291. [Google Scholar] [CrossRef]
  159. Evans, J.R.; Mitchell, S.A.; Spriggs, K.A.; Ostrowski, J.; Bomsztyk, K.; Ostarek, D.; Willis, A.E. Members of the poly (rC) binding protein family stimulate the activity of the c-myc internal ribosome entry segment in vitro and in vivo. Oncogene 2003, 22, 8012–8020. [Google Scholar] [CrossRef] [Green Version]
  160. Prieto, C.; Kharas, M.G. RNA Regulators in Leukemia and Lymphoma. Cold Spring Harb. Perspect. Med. 2020, 10, a034967. [Google Scholar] [CrossRef]
  161. Shen, L.; Shi, Q.; Wang, W. Double agents: Genes with both oncogenic and tumor-suppressor functions. Oncogenesis 2018, 7, 25. [Google Scholar] [CrossRef]
  162. Mohibi, S.; Chen, X.; Zhang, J. Cancer the’RBP’eutics-RNA-binding proteins as therapeutic targets for cancer. Pharmacol. Ther. 2019, 203, 107390. [Google Scholar] [CrossRef]
  163. Azmi, A.S.; Uddin, M.H.; Mohammad, R.M. The nuclear export protein XPO1-from biology to targeted therapy. Nat. Rev. Clin. Oncol. 2021, 18, 152–169. [Google Scholar] [CrossRef] [PubMed]
  164. Flach, J.; Jann, J.C.; Knaflic, A.; Riabov, V.; Streuer, A.; Altrock, E.; Xu, Q.; Schmitt, N.; Obländer, J.; Nowak, V.; et al. Replication stress signaling is a therapeutic target in myelodysplastic syndromes with splicing factor mutations. Haematologica 2021, 106, 2906. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structure of (A) MSI2, (B) hnRNP K, and (C) EIF4A1 proteins. Structure schema obtained from AlphaFold (AF-Q96DH6, AF-P61979, and AF-P60842, respectively) (https://alphafold.ebi.ac.uk/assets/License-Disclaimer.pdf, accessed on 7 August 2022, “not changes were made”) [92,93].
Figure 1. Structure of (A) MSI2, (B) hnRNP K, and (C) EIF4A1 proteins. Structure schema obtained from AlphaFold (AF-Q96DH6, AF-P61979, and AF-P60842, respectively) (https://alphafold.ebi.ac.uk/assets/License-Disclaimer.pdf, accessed on 7 August 2022, “not changes were made”) [92,93].
Ijms 23 09552 g001
Table 1. RNA-binding domains classification.
Table 1. RNA-binding domains classification.
Type of Binding DomainStructure InterfaceInteraction between RBP and RNANucleic Acid
Affinity
RBPs Containing RBDs
KH type I (hnRNP K homology type I)Structure approximately 70 amino acids long. It typically adopts βααββα, forming by a β-sheet composed of 3 antiparallel β-strands and 3 α-helices [27,28]Four single-stranded nucleotides are recognized by the invariant Gly–X–X–Gly motif, the near helices, and the β-strand that follows α2 (type I)C-rich ssDNA and ssRNA [27,28]hnRNP K [29], Nus A [30], SF1 [31], Nova-2 [32]
KH type II (hnRNP K homology type II)Structure approximately 70 amino acids long. It is displayed as a αββααβ [27]Same as KH I, where four single-stranded nucleotides are recognized by α3 (instead of α2 as in KH I)C-rich ssDNA and ssRNA [27,28]hnRNP K [29], Nus A [30], SF1 [31], Nova-2 [32]
KH type III (hnRNP K homology type III)Structure approximately 75-80 amino acids long. It typically adopts a spatial configuration of βααββα, forming a 3-stranded β-sheet held against a 3-helix cluster [27]Same as KH type IC-rich ssDNA and ssRNA [27,28]Nova-2 [33], hnRNP K [34]
RRMs (RNA recognition motifs)Structure approximately 80-90 amino acids long. It typically adopts topology of βαββαβ, forming a 4-stranded β-sheet and 2 α-helices [35]Main interaction between the binding domain and RNA is mediated through the β-sheetPolypyrimidine (mainly C- and U-rich sequences) ssRNA [35,36]U2AF65 [36], nucleolin [37], SRp20 [38], hnRNP F [39], FOX1 [40], Musashi 1, Musashi 2 [41], RBM39 [42]
ZnF (Zinc Fingers)Structure approximately 30 amino acids long. It displays a ββα topology, forming a β-hairpin and an α-helix together with a Zn2+ ion [43]Binding to nucleic acids through the α-helixdsDNA, ssRNA, dsRNA [44]MBNL1 [44], TFIIIA [45], ZRANB2 [46]
dsRBDs (double-stranded RNA binding domains)Structure approximately 65-70 amino acids long. It typically adopts a αβββα topology, where 2 α-helices are packed along a 3-stranded anti-parallel β-sheet [47]Binding to dsRNA backbone through α2 and the β1- β2 loop. Additional interactions occur through the α1dsRNA [47]ADAR1 [48], Dicer [49]
DEAD-boxStructure approximately 300-400 amino acids long. It adopts a βαββαβ topology, forming a 4-stranded β-sheet and 2 α-helices, similar to RRM binding domains [50]Helicase core binds to the backbone of the RNA, without contact with the nucleotidesPolypyrimidine ssRNA [50]eIF4A1/DDX2 [51], p68 [52], p72 [53]
PUF (Pumilio-fem-3 binding factor)Structure approximately 6-8 tandem copies of a 35 amino acids long sequence. It adopts a topology of 3 α- helices, forming a triangle [54]Binding to RNA is through the α 2 in each tandem repeatssRNA [55]PUM1, PUM2 [55]
SAM (Sterile alpha motif)Structure approximately 150-160 amino acids long. It displays a topology of 6 α-helices, packed by a hydrophobic core [56]Traditionally known to bind protein, but has recently been shown to bind RNAHairpin RNA [56]p63, p73 [57], p73, EPHA2 [58]
Table 2. Summary of RBPs implicated in hematological malignancies.
Table 2. Summary of RBPs implicated in hematological malignancies.
RBPRNA Binding MotifRole in Normal HematopoiesisRelated Hematological Malignancies
ADAR1dsRBD [48]Regulation of HSCs differentiation via base editing activityEnhanced editing activity in CML, increasing self-renewal capacity
DDX3X, DDX5DEAD Box [51]Essential for the innate immune response and normal hematopoiesisFrequently mutated in hematological malignances, and upregulated upon Imatinib treatment
DDX21DEAD Box [51]HSCs self-renewalIncreases leukemia stem cell proliferation in AML
EIF4EDEAD BoxTranscription factor, cell differentiation [94]Blockage of myeloid differentiation, leading to leukemogenesis
hnRNP KKH1, KH2, KH3 [95]DNA damage response and cell cycle arrest [96,97]Deletion in AML [98,99,100,101], overexpression in CML [102,103], oncogen in DLBCL [22] and MM [104]
IGF2BP3RRM, KH [105]Self-renewal of HSCs [106]B-ALL cell survival [106], MLL [107], and therapeutic resistance in MM [108]
MSI2RRM [41]Self-renewal of HSCs [109]Upregulated in most hematological malignancies and associated to poor prognosis [110,111]
RBM39RRM [42]Part of the spliceosome [112]AML malignant cell growth and maintenance [113], as well as myeloma progression [114]
SRSF2RRM [38]Essential for myeloid hematopoiesis [115]Mutations associated to poor survival in MDS [115]
SF3B1RRM [36]HSCs homeostasis [116]Mutations associated with MDS [116]
ZFP36ZnF [43]Hematopoiesis and cell differentiation [117]Loss of function leads to leukemogenesis [117]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Aguilar-Garrido, P.; Otero-Sobrino, Á.; Navarro-Aguadero, M.Á.; Velasco-Estévez, M.; Gallardo, M. The Role of RNA-Binding Proteins in Hematological Malignancies. Int. J. Mol. Sci. 2022, 23, 9552. https://doi.org/10.3390/ijms23179552

AMA Style

Aguilar-Garrido P, Otero-Sobrino Á, Navarro-Aguadero MÁ, Velasco-Estévez M, Gallardo M. The Role of RNA-Binding Proteins in Hematological Malignancies. International Journal of Molecular Sciences. 2022; 23(17):9552. https://doi.org/10.3390/ijms23179552

Chicago/Turabian Style

Aguilar-Garrido, Pedro, Álvaro Otero-Sobrino, Miguel Ángel Navarro-Aguadero, María Velasco-Estévez, and Miguel Gallardo. 2022. "The Role of RNA-Binding Proteins in Hematological Malignancies" International Journal of Molecular Sciences 23, no. 17: 9552. https://doi.org/10.3390/ijms23179552

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