Skip to main content

Role of non-coding RNA networks in leukemia progression, metastasis and drug resistance

A Correction to this article was published on 29 December 2020

This article has been updated

Abstract

Early-stage detection of leukemia is a critical determinant for successful treatment of the disease and can increase the survival rate of leukemia patients. The factors limiting the current screening approaches to leukemia include low sensitivity and specificity, high costs, and a low participation rate. An approach based on novel and innovative biomarkers with high accuracy from peripheral blood offers a comfortable and appealing alternative to patients, potentially leading to a higher participation rate.

Recently, non-coding RNAs due to their involvement in vital oncogenic processes such as differentiation, proliferation, migration, angiogenesis and apoptosis have attracted much attention as potential diagnostic and prognostic biomarkers in leukemia. Emerging lines of evidence have shown that the mutational spectrum and dysregulated expression of non-coding RNA genes are closely associated with the development and progression of various cancers, including leukemia. In this review, we highlight the expression and functional roles of different types of non-coding RNAs in leukemia and discuss their potential clinical applications as diagnostic or prognostic biomarkers and therapeutic targets.

Introduction

Leukemia is a class of blood cancers characterized by an oligoclonal expansion of hematopoietic cells that infiltrate the bone marrow and can also invade the blood and other extramedullary tissues [1]. The proliferation of leukemic cells causes the expulsion of the normal hematopoietic cells and the loss of their functions, leading to severe symptoms, including thrombocytopenia, anemia, and immunodeficiency. Hematological cancers are ranked as the 11th common type of cancer and the 10th common cause of cancer-related death. More than 350,000 new leukemia cases and 265,000 leukemia deaths were estimated to have occurred in 2012 [2]. In the United States, leukemia accounts for approximately 4% of cancer-derived mortalities and 3.5% of all cancer cases. The incidence, mortality, and survival of leukemia depends on the diagnosis, prognosis, as well as natural history of neoplasms arising from the malignant transformation of hemopoietic stem cells or progenitor cells in the bone marrow [3].

Leukemia can be classified according to its progression pattern (acute or chronic) and affected lineage (lymphoid or myeloid). The four major subtypes are acute lymphoblastic leukemia (ALL), chronic lymphoblastic leukemia (CLL), acute myeloid leukemia (AML), and chronic myeloid leukemia (CML) [4, 5]. ALL is one of the most common types of malignancy in children worldwide [6], while the other subtypes are more common in adults. In all types of leukemia, the abnormal proliferation of bone marrow and blood cells interferes with the production of functionally healthy cells. Thus, anemia ensues in people with leukemia resulting in reduced ability to fight infections and clotting disorders. For most patients, the causes of leukemia and its subtypes are unclear partly due to diverse abnormalities and multiple risk factors. However, the genetic background interacting with environmental factors including exposure to high doses of radiation or carcinogenic agents, such as benzene; parental occupational exposures; and infections all contribute to a higher risk of developing leukemia [7].

The underlying molecular mechanisms mediating the pathophysiology of leukemia are not fully understood. Thus, deeper insights in the genetic basis of the disease and their influence on the progress of the disease and treatment response are crucial to discovering new prognostic markers and novel therapeutic targets that can open new doors in personalizing treatment. The focus of research for decades has been on the expression of messenger RNAs that code for proteins. Recently, there has been much research suggesting that protein-coding genes only cover a small proportion of the human transcriptome and that a more significant proportion of the human transcriptome (66%) is composed of long non-coding RNAs (long ncRNAs), antisense and micro RNAs (miRNAs), and pseudogenes [8]. Current evidence has shown that ncRNAs might act as a link between the genome and the environment by being an intricate player in the process of gene expression, contributing to the pathogenesis of various human diseases, including cancer. Several studies have documented the involvement of ncRNAs in differentiation, proliferation, and apoptosis of leukemic cells and their potential as a future prognostic biomarker.

In the current review, we discussed the characteristics and role of leukemia related non-coding RNAs. We provided a succinct overview of the current understanding of non-coding RNA expression patterns in different types of leukemia, the mechanisms that contribute to leukemia carcinogenesis, and their role in drug resistance. Deciphering the essential role of diverse non-coding RNAs may improvise the understanding of the underlying biological events, ultimately leading to the identification of novel therapeutic targets, opening new prospects for treatment, diagnosis, and prognostication of different types of leukemia.

Non-coding RNA networks and leukemia

Currently, there is an overpowering proof showing that transcriptional, posttranscriptional and translational controls, mediated by different non-coding RNAs, apply necessary pleiotropic activities on various highlights of leukemia science. This has opened space for disclosure and portrayal of non-coding RNAs as biomarkers in leukemia and prompted several investigations in this field over the last 10 years. The full picture of these unusually communicating non-coding RNAs in leukemia is slowly developing [9,10,11,12,13,14,15,16,17]. The vital role and underlying molecular mechanisms of non-coding RNAs and their therapeutic potential in leukemia are outlined in Table 1.

Table 1 Roles of ncRNAs implicated in leukemia

Characteristics of non-coding RNA networks

Latest proceedings in high-throughput sequencing for whole genomes and transcriptomes demonstrated that fewer than 2% of the entire human genome encodes proteins, whereas a large portion of the human genome, constituting at least 75%, encodes ncRNAs [74]. Currently, ncRNAs are classified according to transcript size into two broad categories, small (< 200 nucleotides; ncRNAs) and long (> 200 nucleotides; lncRNAs) non-coding RNAs (lncRNAs) (Fig. 1). The ncRNAs play a major role in the process of gene expression, RNA maturation, and protein synthesis [75,76,77]. With the emerging evidence, it has become quite evident that not only protein-coding mutations but variations within the noncoding genome are also responsible for various cancer etiologies [78, 79].

Fig. 1
figure 1

Classification of noncoding RNAs (ncRNAs). Noncoding RNAs are classified into small ncRNAs (< 200 nucleotides) or long ncRNAs (> 200 nucleotides) based on their length. Small ncRNAs are further classified into functional and regulatory noncoding RNAs while long ncRNAs are classified based on their structure, function and location

Long non-coding RNAs

lncRNAs are defined as transcripts with lengths exceeding 200 nucleotides that are not translated into protein [80, 81], and most of them are markedly expressed in differentiated tissues or particular cancer types [78]. RNA polymerase II is responsible for executing the transcription of lncRNAs, and generally, they are expressed in a tissue-specific manner [78, 82]. LncRNAs regulate several biological processes such as differentiation, development and biogenesis and multiple human disorders, including certain malignancies are associated with deregulation of lncRNAs. Deregulation of lncRNAs was demonstrated to be intrinsically connected with human illnesses, including different kinds of malignant growths [78, 82]. Because of this, lncRNAs have become a focal point of researchers, and practical explanations of the roles of lncRNAs are an evolving line of research. Usually, lncRNAs utilize various instruments to implement their functions at a cellular level. For example, lncRNAs can influence chromatin redesigning and methylation, act as a miRNA restraint sponge, and regulate protein complexes stability [76, 83, 84] (Fig. 2).

Fig. 2
figure 2

General mechanisms for functions of Long Noncoding RNAs. Nuclear lncRNAs are implicated in Epigenetic regulations, Transcriptional regulations, and splicing regulations while cytoplasmic lncRNAs are involved in mRNA stability, act as small regulatory RNA sponges, regulate mRNA translation and can also be small peptide producers

Several pieces of evidence have shown that some lncRNAs, for example, TARID, Kcnq1ot1, and AS1DHRS4, engage DNA methyltransferases to alter chromatin conformation or act to alter the position of nucleosome through the SWI/SNF complex as observed in SChLAP1 [85,86,87]. The histone methyltransferase poly-comb repressive complex-2 (PRC2) is a widely studied protein managed by ncRNAs and has shown as an intermediary target of lncRNAs [88]. PRC2 appears to play a role in inactivating chromatin through initiating the inhibitory H3K427me3 histone marks [88]. Also, chromatin alterations by specific lncRNAs, for example, HOTTIP and CCAT1, results in tweaking chromosome circling and influencing gene promoters [89, 90]. The lncRNA Firre was demonstrated to be crucial in maintaining inactivation of the X chromosome [91]. X-linked lncRNA Firre cohesion with the chromatin remodelers, CTCF and attachment, is one of the essential steps in the process and includes changing chromatin confirmation during the inactivation of X chromosome process. Subsequently, the inactive X chromosome is positioned close to the nucleolus and maintain H3K27me3 methylation [91]. Different lncRNAs have their distinct inhibitory roles regulated through the action of authoritative miRNAs, which can seize the biomolecules and diminish their potential to inhibit their targets [82].

The roles of miRNA in leukemia have been broadly explored in recent years, but the utilitarian roles of lncRNAs in such tumors are yet unclear. Numerous lncRNAs are deregulated in different sorts of malignant growths, including head and neck cancer [92]. Unmistakable expression profiles of lncRNA have been distinguished in leukemia [9, 33, 35, 45, 47,48,49, 93,94,95,96,97,98,99]. Some of these have been demonstrated to have well-understood jobs in the development and progression of leukemia, suggesting the vital use of lncRNAs as novel biomarkers and potential targets for the treatment of leukemia. Recent shreds of evidence have demonstrated that few lncRNAs play significant physiological roles and are essential for regulating different levels of gene expression [84, 100, 101]. While some of the lncRNAs act as oncogenes, others function as tumor suppressors, and they are involved in cellular processes, including the cell cycle and tumor invasion and metastasis [102]. For example, the lncRNA HOXA cluster antisense RNA2 (HOXA-AS2), which has been previously shown to have oncogenic properties in several human malignancies, was found to diminish glucocorticoid sensitivity in acute lymphoblastic leukemia through the HOXA3/EGFR/Ras/Raf/MEK/ERK pathway [33]. Likewise, exhaustive lncRNA expression profiling by RNA sequencing has uncovered that lncRNA RP11-342 M1.7, lncRNA CES1P1 and lncRNA AC008753.6 are both independent as well as in combination, serve as predictive factors for AML risk [35]. LncRNA LINP1 was found to regulate AML progression employing the HNF4alpha/AMPK/WNT5A signaling pathway [48]. miR-335-3p dysregulation, directed by the lncRNAs NEAT1 and MALAT1, is associated with a poor prognosis in childhood ALL. By and large, these discoveries provide a greater depth of understanding into the pathogenesis of a high-risk group of leukemias that can help clinicians explore the possibility of using lncRNAs for treatment.

Micro RNAs

Micro RNAs (miRNAs) are a subset of non-coding RNAs ~ 19–20 nt in length with 5′-phosphate and 3′-hydroxyl ends. The ribonuclease Dicer processes them from precursors having a characteristic hairpin secondary structure (Fig. 3). miRNAs were first discovered in Caenorhabditis elegans and have since been found in most eukaryotes, including humans [103,104,105]. According to the reports, human genome comprised of approximately 1–5% of miRNA, which is responsible for at least 30% of the protein-coding genes [106,107,108,109,110]. To date, 940 distinct miRNA molecules have been identified [111,112,113]. The knowledge about the specific targets and biological functions of miRNA molecules is still scarce, but their crucial role in the regulation of gene expression, controlling diverse cellular and metabolic pathways is well-evident [114,115,116,117,118,119]. As this field is still emerging, there are only a limited number of studies in the context of miRNAs in leukemia.

Fig. 3
figure 3

MicroRNA (miRNA) biogenesis and regulation of gene expression. The series of events includes the production of the primary miRNA (pri-miRNA) transcript by RNA polymerase II or III and cleavage of the pri-miRNA into a stem-loop structured miRNA precursor (pre-miRNA) by the microprocessor complex Drosha-DGCR8 (Pasha) in the nucleus. Then the pre-miRNA hairpin is exported from the nucleus by Exportin-5-Ran-GTP. In the cytoplasm, the RNase Dicer in complex with the double-stranded RNA-binding protein TRBP cleaves the pre-miRNA hairpin to its mature length. The functional strand of the mature miRNA is loaded together with Argonaute (Ago2) proteins into the RNA-induced silencing complex (RISC), where it guides the RISC to silence target mRNAs through mRNA cleavage, translational repression or deadenylation, whereas the passenger strand is degraded

While some of the miRNAs work as oncogenes, others work as tumor suppressors [120]. For instance, it has been shown that the balance between miR-194-5p and its target BCL2-associated transcription factor 1 (BCLAF1) is commonly deregulated in AML patients [18]. Also, miR-10a-5p was found to be overexpressed in relapsed AML cases [121]. Furthermore, the expression of miR-96 was downregulated in newly diagnosed AML and is associated with leukemic burden [122]. Collectively, these findings allow us to develop a better understanding of the underlying mechanisms of a high-risk group of leukemias that can assist clinicians in clarifying the function of miRNA and use this information to guide treatment.

Role of microRNA gene abnormalities in leukemia

Abnormal expression of miRNA has been reported in many malignancies, including stomach [123], brain [124], breast [125], lung [126], liver [127], colon [128], leukemia [129] and lymphoma [130]. Many studies have reported that microRNA function as a tumor suppressor or oncogene. In most of the tumors, the tumor -suppressing miRNAs are downregulated, whereas the oncogenic miRNAs are overexpressed. Jongen-Lavrencic et al., [131] reported that miR-155 is upregulated in hematopoietic stem cells carrying FLT3-ITD and nucleophosmin (NPM1) gene mutations of AML patients. Similarly, Lagos-Quintana et al., [132] in murine lymphocyte precursors reported increased expression of miR-155 that induces polyclonal lymphocytosis and develops high-grade lymphocytic leukemia. Also, in the case of a myeloproliferative disease, the overexpression of miR-155 was reported that leads to increased granulocyte-monocyte cells [122]. Fuster O et al., [133] suggested that abnormal expression of miR-155 signaling targets SHIP1 and CEBPB in AML patients, both of which are critical in granulopoiesis. Yamamoto et al., [134] reported that miR-133 in leukemic cells targets the Ecotropic viral integration site 1 (Evi1) which upregulated the drug sensitivity and suggested that miR-133 can be a potential therapeutic target for Evi1-overexpressing leukemia. In AML cell lines, Xiao et al. [135] reported elevated expression of miR-223 that inhibited proliferation and cell motility but promote cell apoptosis. Several studies reported that ectopic miR-223 overexpression decreased the tumorigenesis by controlling the G1/S cell cycle phase transition [136]. Lin X et al., [137] investigated that the miR-370 expression was decreased in pediatric AML patients which in turn contribute to the significant progression of the disease and it was suggested that the miR-370 expression could act as non-invasive diagnostic and, a prognostic marker for pediatric AML patients. Magee P et al. has reported [138] that abnormal expression of microRNAs induce chemoresistance that affects a variety of cancer types and he also determined that the forced expression of miR-22 and miR-193a leads to inhibition of leukemia progression. Liu X et al., [139] conducted experiments in leukemic cell lines HL60, NB4, and K562 and reported that the upregulation of miR-181a induces higher cell proliferation thereby increased cell cycling by targeting ATM. It has been investigated that the transfection of miR-128 increased the drug sensitivity, enhanced apoptosis in HL60 cell lines [140], whereas the DNA damage was tolerated; however, the molecular mechanism is yet to be elucidated. However, Volinia S et al., [141] reported that miR-128 to be overexpressed and upregulated in different malignancies, but its expression was decreased in AML cells carrying NPM1 mutations. Imatinib Resistance has been reported as a major hurdle for the treatment of chronic myeloid leukemia (CML). The miRNAs are involved in various processes from the development to drug resistance of tumors, including chronic myeloid leukemia (CML). Recent data suggested that miR-221-STAT5 axis played crucial roles in controlling the sensitivity of CML cells to imatinib [142]. Another recent finding reports that lncRNA MALAT1/miR-328 axis promotes the proliferation and imatinib resistance of CML cells, providing new perspectives for the future study of MALAT1 as a therapeutic target for CML [40]. In addition, miR-214 was associated with the imatinib resistance in CML patients by regulating ABCB1 expression [143].miR-30e has been shown to be directly targeting ABL mRNA and leads to decreased translation of ABL protein [144]. In K562 cells, the increased expression of miRNA-30e induces apoptosis and suppresses proliferation and sensitized the cells to imatinib treatment. miR-203 enhances the sensitivity of CML patients to imatinib and its expression was downregulated in bone marrow of CML patients [145].

Circular RNAs

Circular RNAs (circRNAs) are an abundant class of regulatory transcripts primarily derived from protein-coding exons and widely expressed across eukaryotic organisms, including Homo sapiens and Mus musculus [146,147,148,149,150]. They play an essential role in regulating gene expression [151] through forming covalently closed continuous loop structures with no exposed ends. CircRNAs are evolutionarily conserved, display a higher degree of relative stability in the cytoplasm and are often expressed in a tissue/developmental stage-specific trend [152]. Briefly, circRNAs are produced co-transcriptionally from precursor mRNA by back-splicing of RNA polymerase II transcribed genes and often expressed at only low levels. The biogenesis of circRNAs is regulated through cis and trans-acting regulatory elements that control splicing [153]. The structural form of most circRNAs is composed of multiple exons, and multiple circRNA isoforms can be expressed from a gene with the inclusion or exclusion of internal introns through alternative splicing [153,154,155].

Recent studies have shown that several circRNAs play important physiological and functional roles at multiple stages of the gene expression regulation cascade [84, 100, 101]. CircRNAs are known to be involved in post-transcriptional regulation by acting as decoys for binding of micro RNAs, reducing their cellular availability and resulting in the upregulation of their target mRNAs (Fig. 4). For example, circRNA ciRS-7, also known as CDR1as, is produced from the vertebrate cerebellar degeneration-related 1 (CDR1) antisense transcript and acts as an RNA sponge to repress miR-7 activity [148, 156]. Knockout mice of CDR1as show defects in sensorimotor gating [157] and knockdown of CDR1as expression results in a decrease of tumor growth and proliferation in cancer cell lines [158, 159]. Another circular RNA produced from the Sry gene has 16 binding sites for miR-138, and overexpression constructs of Sry circRNA attenuate the knockdown effects of miR-138 target mRNAs [156]. Indeed, multiple studies have remarkably demonstrated the potent sequestering effects of miRNA activity by circRNAs, making them excellent agents for competing endogenous RNA activity [148, 156, 160,161,162,163,164]. Increasing evidence also suggests that circular RNAs could perform other functional roles such as storage or sequestration of transcription factors and RNA binding proteins [165], microRNA transport [157] or encode functional proteins [166,167,168,169].

Fig. 4
figure 4

General mechanisms for functions of circular RNAs (circRNAs). circRNAs can function as a sponge for a miRNA/RBP keeping miRNA/RBP away (dashed arrows) from its mRNA targets, thus altering gene expression. Through interaction with U1 snRNP, exon-intron circRNAs (EIciRNAs) can interact with transcription complexes at host genes to induce their transcription

CircRNAs are altered in a variety of pathological conditions, which has stimulated significant interest in their role in human disease and cancer. There is emerging evidence that circRNAs show close association with many human diseases, including cancers – often but not always involving micro RNA (miRNA) intermediate. One study revealed hundreds of circRNAs being more abundant in blood than corresponding linear mRNAs, which suggests that circRNAs could be used as new biomarkers in standard clinical blood samples [170]. For instance, circ-CBFB was found to promote proliferation and inhibit apoptosis in CML by regulating the miR-607/FZD3/Wnt/beta-catenin pathway [171]. Additionally, circ_0009910 was found to be significantly upregulated in AML patients, and its high expression was shown to be associated with poor outcomes of AML patients [68]. Similarly, hsa_circ_0080145 was found to regulate CML cell proliferation by acting as a miR-29b sponge, and its knockdown was found to suppress CML cell proliferation [170] significantly. On the other hand, circRNAs circ_0132266 and hsa_circ_0004277 were found to be significantly downregulated in CLL and AML, respectively [63, 65].

We have identified multiple circular RNAs that are differentially expressed in metastatic versus primary ovarian tumors [172]. These circRNAs exhibits a robust expression pattern compared to their linear counterparts with higher power to distinguish between tumor subtypes. This may offer a more robust diagnostic marker of disease progression and prognosis. Our new results have indicated a substantial genetic control of the circular RNA expression that is mostly independent of the basal gene expression [173]. The power to distinguish between tumor subtypes along with an independent genetic control mechanism for their expression strongly points towards a functional and regulatory role for the circular RNA structures and their potential to contribute to disease pathogenicity. It is, therefore, worthwhile to investigate the mechanisms for biogenesis of circRNAs and their contribution to pathogenesis; this may lead to the development of new therapeutic interventions and biomarkers with diagnostic and prognostic capabilities.

Underlying mechanisms of chemoresistance regulated by ncRNAs in leukemia

As in many cancers, resistance to therapy is a significant problem in the treatment of leukemia patients. The most commonly used chemotherapeutic drugs like bendamustine, chlorambucil, and rituximab [174, 175] although show initial response, but later on patients acquire resistance to these therapeutic regimens, hence limiting their efficacy. Also, many leukemia patients show resistance before treatment (intrinsic resistance) and therefore do not even show initial response. While the molecular mechanisms for both intrinsic and acquired resistance are mostly unidentified, identification of unique targets and pathways involved are still an area of intense investigation. Though genetic and epigenetic modifications that result in dysregulation of multi drugs transporters, alterations of drug targets & metabolism of drugs, defects in apoptosis & autophagy machinery, disruption of redox system, increased DNA repair and increased stem cell populations. Have been identified as mediators of drug resistance, the exact mechanisms of drug resistance, cross-talk among different mechanisms and their regulation are still under investigation. Recently, studies have conclusively established the role of miRNAs in chemotherapeutic resistance in leukemia [176, 177]. These studies have shown the deregulation of many miRNAs and their association with resistance to chemotherapy. For example, miR-181a and miR-181b are downregulated in chronic lymphocytic leukemia (CLL) [138] and overexpression of these miRNAs sensitize CLL cells to fludarabine mediated cell death by targeting B-cell lymphoma − 2 (BCL − 2), myeloid cell leukemia-1 (MCL-1) and X-linked inhibitor of apoptosis protein (XIAP) [178]. Similarly, restoration of miR-181b sensitize leukemia cells to doxorubicin (DOX) and cytarabine (ara-C) by downregulating MCL-1 and high mobility group box-1 (HMGB1) expression [179]. On the contrary, ectopic overexpression of miR-125b in leukemia cells induced resistance to daunorubicin (DNR) and prevented apoptosis by downregulating G-protein-coupled receptor kinase 2 (GRK2) and p53 -upregulated modulator of apoptosis (PUMA) [180].

Chronic myeloid leukocyte (CML) is characterized by the Philadelphia (Ph) chromosome [181] with fusion protein breakpoint cluster region-Abelson murine leukemia (BCR-ABL) tyrosine kinase overexpression. Interestingly, Imatinib, an inhibitor of BCR-ABL, show improved therapeutic efficacy in Ph-positive CML patients [182]. Interestingly, downregulation of ABL targeting miRNA-30e was reported in CML cell lines and patient samples [182]. Furthermore, overexpression of miRNA-30e in K562 leukemia cells suppressed proliferation, induced apoptosis and sensitized them to Imatinib treatment. While miRNA-203 sensitizes CML cells to Imanitib and induces apoptosis [145], miRNA-486, on the other hand, promotes Imanitib resistance by targeting PTEN and FOXO1 [183].

While the intrinsic resistance is due to many factors, including miRNA in our case, acquired resistance by tumor cells is promoted in response to continuous drug treatment. DNR and ara-C (anthracyclines) are most commonly used and effective chemotherapeutic drugs for leukemia treatment [184]. Though the use of these drugs results in the complete remission of the disease, most of the patients relapse within 5 years [185, 186], while inefficient tumor cell targeting, mutagenic effects of the drug or selection of resistant clones might be the reasons for relapse and development of aggressive tumors, however the underlying mechanism(s) are still to be identified. These anthracyclines by intercalating into the DNA and targeting Topoisomerase II [187, 188] hinder replication [189]. Interestingly, Topoisomerase II is downregulated in drug-resistant AML subtypes [190, 191], thus making these tumors resistant to these drugs. The topoisomerase II cuts DNA strands and binds to the scaffold/matrix-associated protein region (S/MAR) to prevent or resolve DNA supercoils. Therefore, anthracycline treatment results in DNA double-strand breaks which can be temporarily fixed by non-homologous end joining leading to gene mutation and t4:11 is a common mutation that occurs at S/MAR in AML [192,193,194]. S/MARs by interacting with HDACs regulate expression of miRNAs like miR-93, miR-221, miR-17, let-7b and miR-17-92 cluster. While the dislocation or loss of S/MAR can modulate miRNAs expression [195], anthracyclines like daunorubicin can induce DNA damage associated with deregulation of miRNA expression in leukemia.

Though anthracyclines by modulating miRNA expression regulate cell proliferation and apoptosis, specific miRNAs modulate the DNA repair signaling pathway components resulting in the development of therapeutic resistance. In this connection, resistance to daunorubicin (DNR) has been associated with overexpression of miRNA-21 and its downregulation in resistant K562/DNR cells enhanced DNR cytotoxicity in vitro. Similarly, overexpression of miR-181a in HL60, NB4, and K562 cells by targeting ataxia telangiectasia mutated (ATM) increased proliferation [139]. Also, miR-128 by targeting Rad51 promoted DNA damage and sensitized AML OCI-AML3 and MV4–11 cells to oral nucleoside analog prodrug called sapacitabine [196]. Though upregulated in many cancers, miR-128 is downregulated in AML, especially carrying NPM1 mutations [141, 197]. However, ectopic overexpression of miR-128 in HL60 cells increased drug sensitivity and promoted apoptosis [140]. In addition to miR-128, other miRNAs such as miR-103, miR-107, and miR-506 have been found to target Rad51 in other cancers as well. More specifically, miRNA-125b is overexpressed in pediatric acute promyelocytic leukemia (APL) than in other subtypes of acute myelogenous leukemia (AML), and its exogenous expression in AML cells imparted DOX resistance [198].

ABC transporters are most important proteins promoting drug resistance in almost all the tumors. While the above mentioned miRNAs impart drug resistance, many other miRNAs that are involved in sensitizing cancer cells to therapeutic drugs by targeting ABC transporters are downregulated in cancer [199]. In this category, miR-326 was found to downregulate the ABC transporter ABCC144 in resistant HepG2 cells and sensitize them to chemotherapeutic drugs. In addition to ABCC144, miR-326 also negatively regulated other ABC family members such as ABCA2 and ABCA3, which are drug-resistance related genes [200]. However, the miR-326 expression is reported to be significantly downregulated in the multidrug resistance (MDR+) pediatric ALL patients compared to the (MRD-) group [27]. A recent study showed upregulation of miR-125b-2 cluster (Let-7c, miR-125b, and miR-99a) in leukemia patients with ETV6-RUNX1+ fusion gene expression.

Further studies showed that knockdown of miR-125b in REH ETV6-RUNX1+ cells result in increased sensitivity to staurosporine and doxorubicin treatment, while overexpression of miR-125b-2 cluster inhibited apoptosis and increased cell survival suggesting its therapeutic potential in pediatric ALL [201]. In a recent comprehensive study, the involvement of miRNAs in L-asparaginase (L-ASP), vincristine (VCR), prednisolone (PRED) and DNR resistance was investigated [202]. This study showed the involvement of miR-454 in resistance to L-ASP, miR-125b, miR-99a, & miR-100 to DNR and miR-125b to VCR resistance. Furthermore, over expression of miR-125b prevented VCR mediated apoptosis in vitro [202]. Interestingly, leukemia ETV6-RUNX1+ patients with high expression of miR-125b show resistance to VCR treatment. Like chemotherapeutic drugs, use of glucocorticoids (GCs) for clinical treatment of pediatric ALL is also limited by the development of resistance resulting in poor patient response. Involvement of miRNAs in resistance/sensitivity to GC treatment has recently been evaluated [203]. In a genome-wide study, while the expression of miR-335 was found to be downregulated in all pediatric ALL patients, its overexpression sensitized ALL cells to PRED treatment in vitro [204]. In addition to PRED, ALL cells with miR-335 overexpression showed resistance to other chemotherapeutic drugs with limited cell death [204]. Another miRNA, miR-210 is differentially expressed in various types of cancers including leukemia [205]. Using agomiR or antagomiR for miR-210 in LEH cells (to either increase or decrease the expression respectively) modulated the response to dexamethasone (DEX), L-ASP, VCR and DNR [205], suggesting that use of agomiR’s/antagomiR’s can be a novel alternative to overcome miRNA mediated therapeutic resistance in cancers including leukemia [205].

Role of non-coding RNAs in immune modulation in leukemia

Several ncRNAs, including miRNAs, lncRNAs and circRNAs have been implicated in the modulation of the immune system in various human malignancies, including leukemia. These ncRNAs can modulate immune system either directly by regulating the differentiation of immune cells or indirectly by regulating the expression of various signaling molecules, including NF-kB, c-Myc, p53 and Notch. In this section, we will discuss the available evidence on the role of ncRNAs in immune modulation and its implications in leukemia. Most leukemia are driven by genetic or epigenetic abnormalities in hematopoietic stem cells (HSCs) or progenitor cells, leading to differentiation arrest and increased proliferation and survival of immature blasts in the bone marrow. In one of the first studies on understanding the role of lncRNAs in early hematopoietic differentiation, RNA sequencing of HSCs led to the identification of two lncRNAs, lncHSC-1 and lncHSC-2 [206]. Their depletion resulted in altered myeloid differentiation, impaired self-renewal of HSCs and increased T cell differentiation [206]. These results indicate that lncRNAs can regulate HSC differentiation, and any deregulation in their expression might contribute to various hematological malignancies by altering the differentiation of various HSCs. Indeed, several ncRNAs have been found to contribute to leukemogenesis through immune modulation and altering cell differentiation. HOXA transcript antisense RNA, myeloid-specific 1 (HOTAIRM1) is a myeloid-specific long intergenic non-coding RNA (lincRNA), and it is upregulated during myeloid maturation [207]. Knockdown of HOTAIRM1 in the human acute promyelocytic leukemia (APL) cell line NB4 resulted in decreased granulocytic maturation [53]. HOTAIRM1 is known to regulate the expression of the HOX, CD11b and CD18 genes, which are required for myeloid cell differentiation [53]. Pathway analysis of HOTAIRM1 knockdown NB4 cells treated or untreated with all-trans retinoic acid (ATRA) revealed significant alterations in leukocyte mediated immunity, MHC class I protein complex, complement control module and regulation of leukocyte activation pathways [53]. Furthermore, HOTAIRM1 expression is also modulated by another transcription factor, PU.1, during granulocyte differentiation [208]. PU.1 is a master regulator of myeloid differentiation, while PU.1, along with IRF8, is known to control the fates of follicular (FO) and germinal centers (GO) B cells [209]. Double knockout of IRF8 and PU.1 in B cells has been shown to impair the development of FO and GC B cells [209]. This signifies that HOTAIRM1 can modulate tumor immunity in leukemia by interacting with other regulatory molecules. PU.1 is also known to drive the expression of lnc-DC, which is a lncRNA exclusively expressed in human dendritic cells (DCs) and is required for the differentiation of DCs [210]. Knockdown of lnc-DC resulted in impaired DC differentiation and function, and these effects were mediated by lnc-DC by regulating the posttranslational modification of a critical DC transcription factor, STAT3 [210]. Some of the proteins found to be altered after lnc-DC knockdown include those involved in antigen presentation (HLA-DR), cytokine secretion (IL-12) and T cell activation (CD40, CD80, and CD86). PU.1 also induces miR-23-27-24 cluster and plays a vital role in the regulation of immune cell lineage commitment [211].

Furthermore, this miRNA cluster regulates lymphoid cell differentiation and promotes myeloid lineage commitment and cell proliferation by directly targeting various lymphoid transcription factors, including Runx1 [211]. A recent study has identified a lincRNA, LINC00173, to be very specifically expressed in mature granulocytes [212]. Knockdown of LINC00173 in human CD34+ HSCs resulted in a defect in granulocytic differentiation and an increase in myeloid precursors in vitro [212]. Depletion of LINC00173 in NB4 leukemia cells, which carry an intrinsic block of granulocytic differentiation, resulted in reduced cell proliferation, signifying its role in early myelopoiesis [212]. Functional studies revealed the binding of LINC00173 with the EZH2 subunit of PRC2 [212]. X-inactive specific transcript (Xist) is another lncRNA reported in various human malignancies, including leukemia. Conditional knockout of Xist in murine hematopoietic cells resulted in myeloid leukemia and other impairments such as bone marrow dysfunction, lymphoid organomegaly and lymphoid infiltration of end organs [213]. Aforementioned examples emphasize the importance of ncRNAs in regulating immune cell differentiation, which is of great clinical relevance in leukemia.

The tumor suppressor p53 is known to induce the expression of lncRNA activator of enhancer domains (LED) in cancer [214]. The expression of LED is downregulated in leukemia, possibly due to promoter hypermethylation [214]. Another lncRNA, encoded from the first intron of the human p53 gene and known as lncRNAp53int1, is shown to be highly expressed in undifferentiated human myeloid leukemia cells [215]. However, expression of lncRNAp53int1 is significantly reduced during terminal differentiation of human leukemia cells into monocytes and macrophages [215]. Since several drugs have been used to induce differentiation of leukemia cells, targeting of lncRNAp53int1 could offer a newer therapeutic approach for the management of leukemias. Induction of p53 has also been shown to induce two other lncRNAs, nuclear enriched abundant transcript 1 (NEAT1) and lincRNA-p21, in primary human CLL [216]. The expression of NEAT1 is downregulated and seems to be regulated by PML-RARα in APL [217]. NEAT1 is also found to regulate myeloid differentiation in APL [217]. Recently, pharmacological activation of p53 has been shown to induce an immune-inflammatory response by activating NK cells, leading to suppression of leukemia growth [218]. However, p53 activation also results in the overexpression of PD-L1 in the surviving leukemia cells, promoting their immune escape [218]. All these evidences suggest a crucial role of p53 in regulating lncRNAs during immune modulation in leukemia.

Enhancer RNAs (eRNAs) are another class of lncRNAs and have been reported to be involved in immune modulation. Brazao et al. identified three lncRNA loci (LNCGme00432, LNCGme00344 and LNCGme00345), all of which are eRNAs, in a mouse model of B-ALL [219]. All of these eRNAs interact with PAX5, a transcription factor required for B-cell development and associated with the development of B-ALL, and are downstream of the B-cell lymphoma 11a (Bcl11a) gene [219]. Since the Bcl11a gene is required for VDJ recombination of immunoglobin genes and is also involved in B-cell development, a role of these eRNAs along with the PAX5 and Bcl11a genes in normal B-cell development and immune modulation in B-ALL cannot be ruled out.

In CLL, more than 50% of cases carry a deletion of the critical region at 13q14.3 [220, 221]. In addition to various tumor suppressor genes, miR-15a/16–1 and lncRNAs, deleted in lymphocytic leukemia 1 (DLEU1) and 2 (DLEU2), are also transcribed from this locus [222]. The miRNAs and lncRNAs have been reported to be deleted and epigenetically regulated in CLL [222, 223]. Interestingly, DLEU1 and DLEU2 are also known to regulate NF-kB activity through other NF-kβ regulating genes. Furthermore, the miR-15/16 family of genes is also known to induce NF-kβ activity [222] strongly. In CLL, NF-kB signaling is reported to be active, usually through interaction with the tumor microenvironment (TME), which leads to the survival of leukemia cells [224]. Another lncRNA, p50-associated COX-2, extragenic RNA (PACER), which is transcribed from the upstream region of the human COX-2 gene, regulates COX-2 expression by interacting with the repressive p50 subunit of NF-kβ, thereby functioning as a decoy lncRNA for NF-kB signaling [225]. NF-kB induced lncRNA, linc-Cox2, coactivates NF-kB, leading to induction of late-primary response genes in innate immune cells [226]. Since the NF-kβ family of transcription factors plays a crucial role in the regulation of tumor inflammation and immunity [227], we suggest that the NF-kβ as mentioned above regulated ncRNAs might also modulate immune system in leukemia.

Notch-regulated oncogenic lncRNA, leukemia-induced non-coding activator RNA-1 (LUNAR1), has been identified in T-cell acute lymphoblastic leukemia (T-ALL) [228]. Mechanistically, LUNAR1 regulates IGF signaling and induces IGF1R expression, leading to the survival of T-ALL cells [228]. The expression of LUNAR1 is upregulated in primary T-ALL cells, more so in Notch mutated samples, whereas its expression is suppressed upon Notch inhibition [228]. Another lncRNA, NOTCH1 associated lncRNA in T ALL (NALT), is also found to be associated with the Notch1 gene and functions as a transcription factor to activate Notch signaling and promote cell proliferation in pediatric T-ALL cells [229]. Role of Notch signaling in normal and effector immune cell differentiation is well established [230]. Furthermore, Notch can regulate various components of TME, including immune cells, fibroblasts, endothelial, and mesenchymal cells [230]. Since Notch signaling is also involved in human T-ALL [228, 229], we believe that Notch-regulated lncRNAs can potentially modulate immune system in leukemia.

Beta Globin Locus 3 (BGL3) is a lncRNA that regulates Bcr-Abl mediated cellular transformation in CML [57]. Bcr-Abl has been found to negatively regulated BGL3 expression through c-Myc-dependent DNA methylation in CML [57]. Interestingly, BGL3 acts as a competitive endogenous RNA (ceRNA), and it is targeted by many PTEN regulating miRNAs, including miR-17, miR-93, miR-20a, miR-20b, miR-106a and miR-106b [57]. It is well known that loss of PTEN in cancer cells leads to an immunosuppressive microenvironment through secretion of various immunosuppressive cytokines, recruitment of myeloid-derive suppressor cells (MDSCs) and regulatory T-cells (Tregs), and inhibition of CD8+ T-cell killing [231]. Hence, we speculate that BGL3 might also lead to immune modulation in leukemia through PTEN and PTEN-regulating miRNAs, although this needs to be experimentally proven. Colon cancer-associated transcript-1 (CCAT1) is a lncRNA that is known to be highly expressed in adult AML [153]. CCAT1 represses monocytic differentiation and promotes leukemia cell growth by upregulating oncogenic c-Myc and suppressing tumor suppressive miR-155 [153]. c-Myc is also known to induce lncRNA H19 expression in leukemia cells, thereby promoting cell proliferation and survival [232]. Plasmacytoma variant translocation 1 (PVT1) is another lncRNA that exerts its oncogenic effects by stabilizing the c-Myc protein in cancer [233]. Furthermore, in leukemia and other solid tumors, c-Myc is known to induce the expression of cluster of differentiation 47 (CD47), an innate immune regulator, and programmed death-ligand 1 (PD-L1), an adaptive immune checkpoint protein, involved in suppressing the antitumor immune response [234]. Hence, we believe that lncRNAs regulated by c-Myc might also modulate the immune response in leukemia.

Recent evidence also suggests a crucial role of circRNAs in immune modulation and leukemia development. The presence of fusion circRNAs (F-circRNAs) has been shown in PML/RARα positive APL and MLL/AF9 positive AML cells [58]. These F-circRNAs not only caused cellular transformation by activating PI3K and MAPK signaling but also contributed to leukemia cell proliferation, survival, progression and therapy resistance in vivo [58]. Since immune cells also regulates cell proliferation, survival and confer resistance to therapy, we believe that oncogenic F-circRNAs might also be involved in modulating the host immune system in leukemia, giving a survival advantage to leukemia cells. Because the presence of circRNAs has also been detected in extracellular vesicles [91], these circRNAs may modulate TME through cell-to-cell communication, although this is yet to be experimentally proven. Another circRNA, hsa_circ_0075001, has been detected in AML where its expression positively correlated with total NPM1 expression [60]. AML patients carrying a high expression of hsa_circ_0075001 had lower expression of components of the Toll-like receptor signaling pathway, suggesting that this circRNA might be involved in the modulation of the immune response in AML [60]. Another circRNA, circMYBL2, which is derived from the cell-cycle checkpoint gene MYBL2, has been reported to be highly expressed in FLT3-ITD mutation-positive AML patients [235]. Depletion of circMYBL2 inhibited proliferation and induced differentiation of FLT3-ITD AML cells in vitro and in vivo [235]. In a recent study of a comprehensive analysis of circRNA expression during hematopoiesis, the expression of circRNA was found to be highly cell-type specific during hematopoietic differentiation [236]. All these studies highlight the crucial role of circRNAs in immune modulation in leukemias.

Several miRNAs have been shown to modulate immune checkpoint proteins in various human malignancies, including leukemia. In AML, miR-34 regulates PD-L1 expression by targeting PD-L1 mRNA, thereby controlling PD-L1 specific T-cell apoptosis of human AML cells [85]. The miR-17-92 cluster, which encodes six miRNAs including 17, 18a, 19a, 20a, 19b-1, and 92–1, is also known to regulate T-cell responses in graft-versus-host disease (GVHD) post allogeneic bone marrow transplantation in mice [237]. This miRNA cluster has been found to promote CD4 T-cell activation, expansion, migration and Th1 differentiation while suppressing Th2 and Treg differentiation. Inhibition of miR-17 or miR-19b significantly inhibited alloreactive T-cell expansion and IFN-γ secretion, leading to prolonged survival in recipient mice with GVHD while preserving the graft-versus-leukemia effect [237]. Overexpression of miR-125a-5p has been shown to induce granulocytic differentiation, whereas miR-17-92 has the opposite effect in APL cells [238]. A recent study has identified overexpression of miR-708 in AML patients, which delayed HOXA9 mediated transformation in vivo by modulating myeloid differentiation [239]. The authors concluded that miR-708 is an indirect regulator of the HOX program during normal and impaired hematopoiesis [239].

Clinical significance of ncRNAs in leukemia

In the current exploratory genomic era, the cellular or extracellular level of noncoding RNAs (ncRNAs) are advancing for their roles in risk stratification, diagnosis, and prognosis. Biologically ncRNAs regulate different processes such as proliferation, apoptosis, stemness, and differentiation. The clinical significance of ncRNAs in leukemia broadly illustrates their capability for risk stratification, diagnosis, and prognosis [212, 240, 241]. The quantitative assessments of transcripts by highly sensitive assay (qPCR) for minimal residual disease detection make ncRNA as a suitable candidate biomarker. The residual transcript copies play a significant role in detecting minimal residual disease. The best analogy is BCR-ABL international scale detection for deep molecular and ultra-deep molecular response in Philadelphia positive leukemias.

The prerequisite for ncRNAs as biomarkers in leukemia is their aberrant expression in leukemic phenotype., A plethora of differential miRNA, lncRNA and circRNAs from high throughput data, supported the notion and met this primary concern. However, leukemia itself is a disease of heterogeneous cell population; therefore, precisely identifying the robust biomarker in variable data sets of different leukemia subtype is very challenging at the validation step. Furthermore, the ncRNA fine-tune the cellular homeostasis; therefore their regulatory function activated with a slight change in the oncogenic molecular thrust. The ncRNAs modulates and attempt to reconcile the abnormal molecular changes. Recently, three-lncRNA expression-based risk score was developed based on RNA-seq data for AML patients using two leading data repositories [Therapeutically Available Research to Generate Effective Treatments (TARGET) and The Cancer Genome Atlas (TCGA)]. According to prognosis modelling, which was developed based on survival data, the combination of the lncRNA risk score and cytogenetics risk group provided a higher prognostic value than any of the individual prognostic factor [61].

Acute myeloid leukemia is a heterogenous malignancy of defective stem cells with impaired proliferation and differentiation. Many regulatory ncRNAs largely regulate the deregulation, stemness, proliferation and differentiation. Various studies have proved that many deregulated miRNAs are correlated with acute leukemia as compared to control samples. Table 1 shows a list of significant ncRNAs (lncRNA and circRNAs) for their pathological and clinical significance in leukemia.

HOTAIRM1 is located between HOXA1 and HOXA2 gene cluster and regulate granulocytic differentiation in hematopoiesis. High HOTAIRM1 expression results in increased expression of HOXA4 gene expression and defective myelopoiesis. HOTAIRM1 knockdown experiments on NB4 cells correlated with low expression of HOXA1 and HOXA4 cluster genes and block the expression of CD11b and CD18 during granulopoiesis. HOTAIRM1 expression is activated by all-trans retinoic acid, which induces the differentiation of myeloid progenitor cells to granulocytes and mature myeloid cells [53]. HOTAIRM1 transcript also interacts and form complexes with transcripts of other key chromatin structure modulating proteins such as CBX1, PRC1 and PRC2 [242]. HOTAIRM1 was overexpressed in NPM1-mutated AML. Furthermore amongst, 215, intermediate cytogenetics risk group AML patients, high HOTAIRM1 expression was associated with inferior overall survival (OR: 2.04; P = 0.001) and disease-free survival (OR:2.56; P < 0.001) and a higher cumulative incidence of relapse (OR:1.67; P = 0.046). Furthermore, high expression of HOTAIRM1 was associated with poor survival outcome in the subgroup of NPM1 mutation-positive AML patients [54]. HOXA-AS2: HOXA cluster antisense RNA 2 (HOXA-AS2) located between HOXA3 and HOXA4 genes in the HOXA cluster. Like HOTAIR and HOTAIRM1, HOXA-AS2 regulates differentiation of myeloblasts to mature granulocytes and myeloid cells [243]. Dong et al. proved the important role of HOXA-AS2 in chemoresistance of myeloblast and the lncRNA HOXA-AS2 could act as a therapeutic target for overcoming resistance to chemotherapy in AML [96].

DLEU1 and DLEU2 lncRNA mapped on the frequently deleted region of chromosome 3q14.3 region in lymphoma and leukemia. DLEU2 lncRNA act as pre miRNA for 15a and 16–1 and both are involved in the pathogenesis of CLL through NF-kβ activity [222, 223]. LincRNA-p21 in CLL is associated with p53 gene repression; thereby, it acts as tumor suppressor gene, and this finding was confirmed in 68 CLL patients, 62 MM patients when compared with 36 healthy controls. The correlation of p53 repression through LincRNA-p21 makes it eligible therapeutic and prognostic marker in CLL patients [216, 244].

BGL3 lncRNA regulates the oncogenic expression of BCR-ABL fusion gene through c-Myc mediated signaling. The expression of BGL3 gene was inversely regulated through miR-17, miR-93, miR-20a, miR-20b, miR-106a, and miR-106b in Philadelphia positive ALL and CML patients [57, 245].

Non-coding microRNAs (miRNAs) are posttranscriptional and posttranslational regulators of the target genes and proteins, respectively. The expression and modulation of target genes is disease and tissue-specific. In leukemia, miRNA expression signature depends upon the disease subtype, cytogenetic risk group, age and molecular lesions like fusion genes or various mutations in a gene like FLT3, cKIT, NPM1, BCR-ABL, MLL rearrangement. The most frequently deregulated miRNAs in CML include miR-10a, miR-17/92, miR-150, miR-203, and miR-328. Oncogenic role of miR-9 was suggested by Chen et al. in the subgroup of AML patients with mixed lineage leukemia (MLL)-rearrangement [246]. However, Emmrich et al. suggested tumor suppressor role and expression was down-regulated in pediatric AML with t (8;21) translocation [247]. A similar finding was observed in Fu et al. that miR-9-1 was down-regulated in t (8;21) AML patients [248]. Many recent studies have compiled the biological and clinical significance of miRNAs in acute and chronic leukemia [10, 240, 241].

Like lncRNA and miRNA, circular RNAs (circRNAs) express as housekeeping, and regulatory RNAs. The mode of action of circRNA may be autocrine or paracrine; therefore, these circRNAs have been detected in various body fluids. The circRNAs are stable in different body fluids like saliva, urine, blood, and CSF. The basal level of various circRNAs is crucial to explore for understanding their clinical significance. In leukemia, ultra-deep genomic data is available, which enabled to explore different ncRNA entities for their diagnostic and prognostic significance. Various types of circRNAs have been characterized based on their position in the gene, the intron origin circRNA and exonic circRNAs, intergenic circRNAs, and exon-intron circRNAs. Although, a various study has supported the notion of differential circRNA expression profile in leukemia but the validation data from experimental studies is limited.. The origin of circRNAs has been associated with fusion genes in leukemia [249]. Isolated studies have shown the role of following cirRNAs, f-circPR, f-circM9, hsa_circ_0075001, circ-ANAPC7, circ-100,290, circPAN3, circ_0009910, circ-HIPK2, circ-DLEU2, has_cir_0004277, circPVT1 in AML [59, 71, 250,251,252]. In CML, the direct association of circBA9.3 with BCR-ABL tyrosine kinase activity was observed in CML patients. The high expression of circBA9.3 was associated with cell proliferation and inverse relations with apoptosis. Furthermore, the high expression was associated with relapse and disease progression suggesting the possible role of circBA9.3 as a potential therapeutic marker in CML [72].

Conclusions & future perspectives

The crucial role of ncRNAs in the gene regulatory networks and recent progress in the field of genomics and biotechnology has made them a favorable therapeutic targeting agent in cancer. lncRNAs and circRNAs act through various mechanisms as compared to miRNAs in cancer, and so targeting them can help in exploring more critical mechanisms involved in tumorigenesis. This review highlights the therapeutic potential of ncRNAs such as miRNAs, lncRNAs and circRNAs in leukemia and culminates the significance of these biomolecules as they improved the prognostic risk stratification in leukemia. The improvement in risk stratification has led to the generation of medical algorithms that can help in standardizing selection and treatment planning based on the molecular profile of the patient. These risk stratification schemes can be taken one step further by the inclusion of selected ncRNA expression profiles.

Additionally, by artificially modulating the expression of ncRNAs, the therapeutic sensitivity to conventional chemotherapy can be restored. In this regard, miRNAs have become the most extensively studied ncRNAs in leukemia because of their role as an oncogene and tumor suppressor in various cancers, including leukemia and their involvement in the regulation of post-transcriptional processes. The advanced genomic approaches, such as CRISPR-Cas9 technology is used to identify functionally relevant miRNA-mRNA target pairs that regulate leukemia (e.g., AML) cell line growth and will likely prove beneficial for preclinical models. Another approach is the use of miRNA mimics or modified miRNAs as RNA based drugs to target ncRNAs and mRNAs. Silencing of aberrant miRNAs can also be achieved by miRNA sponges and anti-miRNA oligonucleotides (AMOs). Finally, miRNA analysis through advanced next-generation sequencing will provide more details on the involvement of ncRNAs in the onset and progression of leukemia. For efficient miRNA-based therapy, improvised miRNA delivery vehicles with higher stability and less toxicity must be developed.

On the other hand, oncogenic lncRNAs can be targeted using siRNAs by packaging them in nanoparticle vectors for efficient targeting. In addition, high affinity or stability of antisense oligonucleotides can be achieved by synthetically modifying themto reduce the oncogenic lncRNAs by alternative splicing, modulation of RNA and protein interactions or by degrading them. Further, lentiviral vectors can be used as an efficient method for the transportation of RNA products into tissues as they aid in stable transfection by efficiently inserting the siRNA sequence into target cells.

Regarding the many roles of ncRNAs in cancer, there are still many challenges that must be resolved in order to improve the potential of ncRNAs as a potential therapeutic target in cancer. As the complex microenvironment of the cell makes the delivery of ncRNAs very challenging and difficult, the efficient delivery system with minimal toxicity is vital. It is suggested that the drug delivery can be improved by using two or more different carriers for targeting ncRNAs, for example combining nano designs with organ-specific response receptor. Moreover, in order to increase their bioavailability, different ways must be discovered to reduce RNA degradation. Although the field of ncRNAs is well studied, their role as a biomarker and as a therapeutic target in cancer is yet to be explored in detail. Many clinical trials are currently underway, and if some of the challenges mentioned above are addressed appropriately, then we would likely see ncRNAs emerging as a novel target for cancer therapy.

Availability of data and materials

Not applicable, please refer to the original reference.

Change history

  • 29 December 2020

    An amendment to this paper has been published and can be accessed via the original article.

Abbreviations

ALL:

Acute lymphoblastic leukemia

AML:

Acute myeloid leukemia

ATRA:

All-trans retinoid acid

BCLAF1:

BCL2-associated transcription factor 1

BGL3:

Beta Globin Locus 3

CD47:

Cluster of differentiation 47

ceRNA:

Endogenous RNA

circRNA:

Circular RNA

CLL:

Chronic lymphoblastic leukemia

CML:

Chronic myeloid leukemia

CRC:

Colorectal cancer

DCs:

Dendritic cells

DLEU1:

Deleted in lymphocytic leukemia 1

DLEU2:

Deleted in lymphocytic leukemia 2

eRNAs:

enhancer RNAs

GVHD:

Graft-versus-host disease

HOTAIRM1:

HOXA transcript antisense RNAs, myeloid-specific 1

HOXA-AS2:

HOXA cluster antisense RNA2

lncRNA:

long non-coding RNA

MDR:

Multidrug resistance

miRNA:

microRNA

MSC:

Mesenchymal stromal cells

ncRNA:

non-coding RNA

NEAT1:

Nuclear enriched abundant transcript 1

PD-L1:

Programmed death-ligand 1

PI3K:

Phosphoinositide-3 kinase

PRC2:

Poly-comb repressive complex-2

pre-miRNA:

precursor miRNAs

Pri-miRNA:

Primary miRNA

PVT1:

Plasmacytoma variant translocation 1

siRNAs:

small interfering RNAs

TME:

Tumor microenvironment

References

  1. Dohner H, Weisdorf DJ, Bloomfield CD. Acute Myeloid Leukemia. N Engl J Med. 2015;373(12):1136–52.

    PubMed  Google Scholar 

  2. Ferlay J, et al. GLOBOCAN 2012 v1. 0, cancer incidence and mortality worldwide: IARC CancerBase No. 11. Lyon: International agency for research on cancer, 2013; 2016.

    Google Scholar 

  3. Rodriguez-Abreu D, Bordoni A, Zucca E. Epidemiology of hematological malignancies. Ann Oncol. 2007;18(Suppl 1):i3–8.

    PubMed  Google Scholar 

  4. Weinberg RA. The Biology of Cancer: Second International Student Edition. New York: Garland Science; 2013.

  5. Swerdlow SH, International Agency for Research on Cancer. World Health Organization. WHO classification of tumours of haematopoietic and lymphoid tissues. Lyon: International Agency for Research on Cancer. 2008.

  6. Ward E, et al. Childhood and adolescent cancer statistics, 2014. CA Cancer J Clin. 2014;64(2):83–103.

    PubMed  Google Scholar 

  7. Chokkalingam AP, Buffler PA. Genetic susceptibility to childhood leukaemia. Radiat Prot Dosim. 2008;132(2):119–29.

    CAS  Google Scholar 

  8. Hubbard T, et al. The Ensembl genome database project. Nucleic Acids Res. 2002;30(1):38–41.

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Cruz-Miranda GM, et al. Long Non-Coding RNA and Acute Leukemia. Int J Mol Sci. 2019;20(3). https://doi.org/10.3390/ijms20030735.

  10. Liu Y, et al. Role of microRNAs, circRNAs and long noncoding RNAs in acute myeloid leukemia. J Hematol Oncol. 2019;12(1):51.

    PubMed  PubMed Central  Google Scholar 

  11. Handschuh L. Not only mutations matter: molecular picture of acute myeloid Leukemia emerging from transcriptome studies. J Oncol. 2019;2019:7239206.

    PubMed  PubMed Central  Google Scholar 

  12. Wang Y, et al. Long non-coding RNA LINC00899 as a novel serum biomarker for diagnosis and prognosis prediction of acute myeloid leukemia. Eur Rev Med Pharmacol Sci. 2018;22(21):7364–70.

    CAS  PubMed  Google Scholar 

  13. Wang Y, Zhou Q, Ma JJ. High expression of lnc-CRNDE presents as a biomarker for acute myeloid leukemia and promotes the malignant progression in acute myeloid leukemia cell line U937. Eur Rev Med Pharmacol Sci. 2018;22(3):763–70.

    CAS  PubMed  Google Scholar 

  14. Li J, Sun CK. Long noncoding RNA SNHG5 is up-regulated and serves as a potential prognostic biomarker in acute myeloid leukemia. Eur Rev Med Pharmacol Sci. 2018;22(11):3342–7.

    CAS  PubMed  Google Scholar 

  15. Lammens T, et al. Long non-coding RNAs in leukemia: biology and clinical impact. Curr Opin Hematol. 2017;24(4):353–8.

    CAS  PubMed  Google Scholar 

  16. Chen S, et al. Long non-coding RNAs: the novel diagnostic biomarkers for leukemia. Environ Toxicol Pharmacol. 2017;55:81–6.

    CAS  PubMed  Google Scholar 

  17. Zebisch A, et al. Therapeutic Resistance in Acute Myeloid Leukemia: The Role of Non-Coding RNAs. Int J Mol Sci. 2016;17(12):2080.

    PubMed Central  Google Scholar 

  18. Dell'Aversana C, et al. miR-194-5p/BCLAF1 deregulation in AML tumorigenesis. Leukemia. 2017;31(11):2315–25.

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Wan L, et al. MicroRNA-103 confers the resistance to long-treatment of adriamycin to human leukemia cells by regulation of COP1. J Cell Biochem. 2018;119(5):3843–52.

    CAS  PubMed  Google Scholar 

  20. Ferreira AF, et al. ApoptomiRs expression modulated by BCR-ABL is linked to CML progression and imatinib resistance. Blood Cells Mol Dis. 2014;53(1–2):47–55.

    CAS  PubMed  Google Scholar 

  21. Hu N, et al. MALAT1 knockdown inhibits proliferation and enhances cytarabine chemosensitivity by upregulating miR-96 in acute myeloid leukemia cells. Biomed Pharmacother. 2019;112:108720.

    CAS  PubMed  Google Scholar 

  22. Kotani A, et al. A novel mutation in the miR-128b gene reduces miRNA processing and leads to glucocorticoid resistance of MLL-AF4 acute lymphocytic leukemia cells. Cell Cycle. 2010;9(6):1037–42.

    CAS  PubMed  Google Scholar 

  23. Li Q, Song W, Wang J. TUG1 confers Adriamycin resistance in acute myeloid leukemia by epigenetically suppressing miR-34a expression via EZH2. Biomed Pharmacother. 2019;109:1793–801.

    CAS  PubMed  Google Scholar 

  24. Huang Y, et al. Identification of serum miR-34a as a potential biomarker in acute myeloid leukemia. Cancer Biomark. 2018;22(4):799–805.

    CAS  PubMed  Google Scholar 

  25. Zhang J, et al. Differential expression profiles and functional analysis of plasma miRNAs associated with chronic myeloid leukemia phases. Future Oncol. 2019;15(7):763–76.

    CAS  PubMed  Google Scholar 

  26. Zhang C, et al. Identification of miRNA-mRNA network associated with acute myeloid Leukemia survival. Med Sci Monit. 2017;23:4705–14.

    PubMed  PubMed Central  Google Scholar 

  27. Ghodousi ES, Rahgozar S. MicroRNA-326 and microRNA-200c: two novel biomarkers for diagnosis and prognosis of pediatric acute lymphoblastic leukemia. J Cell Biochem. 2018;119(7):6024–32.

    CAS  PubMed  Google Scholar 

  28. Liu Y, et al. Decreased microRNA-30a levels are associated with enhanced ABL1 and BCR-ABL1 expression in chronic myeloid leukemia. Leuk Res. 2013;37(3):349–56.

    PubMed  Google Scholar 

  29. El-Khazragy N, et al. miRNA-155 and miRNA-181a as prognostic biomarkers for pediatric acute lymphoblastic leukemia. J Cell Biochem. 2019;120(4):6315–21.

    CAS  PubMed  Google Scholar 

  30. Zhu X, et al. Functional studies of miR-130a on the inhibitory pathways of apoptosis in patients with chronic myeloid leukemia. Cancer Gene Ther. 2015;22(12):573–80.

    CAS  PubMed  Google Scholar 

  31. Zhou L, et al. microRNA125b promotes leukemia cell resistance to daunorubicin by inhibiting apoptosis. Mol Med Rep. 2014;9(5):1909–16.

    CAS  PubMed  Google Scholar 

  32. Zhou H, et al. Downregulation of miR-224 and let-7i contribute to cell survival and chemoresistance in chronic myeloid leukemia cells by regulating ST3GAL IV expression. Gene. 2017;626:106–18.

    CAS  PubMed  Google Scholar 

  33. Zhao Q, et al. TCF7L2 activated HOXA-AS2 decreased the glucocorticoid sensitivity in acute lymphoblastic leukemia through regulating HOXA3/EGFR/Ras/Raf/MEK/ERK pathway. Biomed Pharmacother. 2019;109:1640–9.

    CAS  PubMed  Google Scholar 

  34. Wang X, et al. Long non-coding RNA taurine-upregulated gene 1 correlates with poor prognosis, induces cell proliferation, and represses cell apoptosis via targeting aurora kinase a in adult acute myeloid leukemia. Ann Hematol. 2018;97(8):1375–89.

    CAS  PubMed  Google Scholar 

  35. Wang Y. Comprehensive long non-coding RNA expression profiling by RNA sequencing reveals potential biomarkers for acute myeloid leukemia risk. Cancer Biomark. 2019;26:93–108.

    CAS  PubMed  Google Scholar 

  36. Zhou X, et al. LncRNA MEG3 regulates Imatinib resistance in chronic myeloid Leukemia via suppressing MicroRNA-21. Biomol Ther (Seoul). 2017;25(5):490–6.

    CAS  Google Scholar 

  37. Yang L, et al. Overexpression of lncRNA PANDAR predicts adverse prognosis in acute myeloid leukemia. Cancer Manag Res. 2018;10:4999–5007.

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Yan H, et al. Long non-coding RNA GAS5 polymorphism predicts a poor prognosis of acute myeloid leukemia in Chinese patients via affecting hematopoietic reconstitution. Leuk Lymphoma. 2017;58(8):1948–57.

    CAS  PubMed  Google Scholar 

  39. Xiao Y, et al. lncRNA UCA1 contributes to Imatinib resistance by acting as a ceRNA against miR-16 in chronic myeloid Leukemia cells. DNA Cell Biol. 2017;36(1):18–25.

    CAS  PubMed  Google Scholar 

  40. Wen F, et al. LncRNA MALAT1 promotes cell proliferation and imatinib resistance by sponging miR-328 in chronic myelogenous leukemia. Biochem Biophys Res Commun. 2018;507(1–4):1–8.

    PubMed  Google Scholar 

  41. Zhang Y, Liu Y, Xu X. Knockdown of LncRNA-UCA1 suppresses chemoresistance of pediatric AML by inhibiting glycolysis through the microRNA-125a/hexokinase 2 pathway. J Cell Biochem. 2018;119(7):6296–308.

    CAS  PubMed  Google Scholar 

  42. Yu J, et al. Genomewide profiling of lncRNA expression patterns in patients with acute promyelocytic leukemia with differentiation therapy. Oncol Rep. 2018;40(3):1601–13.

    CAS  PubMed  Google Scholar 

  43. Wang H, et al. The role of long noncoding RNA HOTAIR in the acquired multidrug resistance to imatinib in chronic myeloid leukemia cells. Hematology. 2017;22(4):208–16.

    PubMed  Google Scholar 

  44. Sun J, et al. A novel antisense long noncoding RNA within the IGF1R gene locus is imprinted in hematopoietic malignancies. Nucleic Acids Res. 2014;42(15):9588–601.

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Chen C, et al. lncRNA-CCDC26, as a novel biomarker, predicts prognosis in acute myeloid leukemia. Oncol Lett. 2019;18(3):2203–11.

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Jia ZW, et al. Expression and clinical significance of LncRNA KCNQ1OT1 in patients with acute myeloid Leukemia. Zhongguo Shi Yan Xue Ye Xue Za Zhi. 2018;26(3):653–7.

  47. Li S, et al. Identification of novel lncRNAs involved in the pathogenesis of childhood acute lymphoblastic leukemia. Oncol Lett. 2019;17(2):2081–90.

    CAS  PubMed  Google Scholar 

  48. Shi J, et al. LncRNA LINP1 regulates acute myeloid leukemia progression via HNF4alpha/AMPK/WNT5A signaling pathway. Hematol Oncol. 2019;37:474–82.

    CAS  PubMed  Google Scholar 

  49. Peng L, Zhang Y, Xin H. lncRNA SNHG3 facilitates acute myeloid leukemia cell growth via the regulation of miR-758-3p/SRGN axis. J Cell Biochem. 2019;121(2):1023-31.

  50. de Souza Melo CP, et al. Long non-coding RNAs: biomarkers for acute leukaemia subtypes. Br J Haematol. 2016;173(2):318–20.

    Google Scholar 

  51. Peng W, Feng J. Long noncoding RNA LUNAR1 associates with cell proliferation and predicts a poor prognosis in diffuse large B-cell lymphoma. Biomed Pharmacother. 2016;77:65–71.

    CAS  PubMed  Google Scholar 

  52. Zhang L, Xu H-G, Lu C. A novel long non-coding RNA T-ALL-R-LncR1 knockdown and Par-4 cooperate to induce cellular apoptosis in T-cell acute lymphoblastic leukemia cells. Leuk Lymphoma. 2014;55(6):1373–82.

    CAS  PubMed  Google Scholar 

  53. Zhang X, Weissman SM, Newburger PE. Long intergenic non-coding RNA HOTAIRM1 regulates cell cycle progression during myeloid maturation in NB4 human promyelocytic leukemia cells. RNA Biol. 2014;11(6):777–87.

    PubMed  PubMed Central  Google Scholar 

  54. Díaz-Beyá M, et al. The lincRNA HOTAIRM1, located in the HOXA genomic region, is expressed in acute myeloid leukemia, impacts prognosis in patients in the intermediate-risk cytogenetic category, and is associated with a distinctive microRNA signature. Oncotarget. 2015;6(31):31613–27.

    PubMed  PubMed Central  Google Scholar 

  55. Zeng C, et al. Overexpression of the long non-coding RNA PVT1 is correlated with leukemic cell proliferation in acute promyelocytic leukemia. J Hematol Oncol. 2015;8:126.

    PubMed  PubMed Central  Google Scholar 

  56. Bhan A, Soleimani M, Mandal SS. Long noncoding RNA and Cancer: a new paradigm. Cancer Res. 2017;77(15):3965.

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Guo G, et al. A long noncoding RNA critically regulates Bcr-Abl-mediated cellular transformation by acting as a competitive endogenous RNA. Oncogene. 2015;34(14):1768–79.

    CAS  PubMed  Google Scholar 

  58. Guarnerio J, et al. Oncogenic role of fusion-circRNAs derived from Cancer-associated chromosomal translocations. Cell. 2016;165(2):289–302.

    CAS  PubMed  Google Scholar 

  59. LAbbate A, et al. MYC-containing amplicons in acute myeloid leukemia: genomic structures, evolution, and transcriptional consequences. Leukemia. 2018;32(10):2152–66.

    CAS  Google Scholar 

  60. Hirsch S, et al. Circular RNAs of the nucleophosmin (NPM1) gene in acute myeloid leukemia. Haematologica. 2017;102(12):2039–47.

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Li S, et al. Profiling and functional analysis of circular RNAs in acute promyelocytic leukemia and their dynamic regulation during all-trans retinoic acid treatment. Cell Death Dis. 2018;9(6):651.

    PubMed  PubMed Central  Google Scholar 

  62. Wu D-M, et al. Role of circular RNA DLEU2 in human acute myeloid Leukemia. Mol Cell Biol. 2018;38(20):e00259–18.

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Li W, et al. Characterization of hsa_circ_0004277 as a new biomarker for acute myeloid Leukemia via circular RNA profile and bioinformatics analysis. Int J Mol Sci. 2017;18(3):597.

    PubMed Central  Google Scholar 

  64. Xia L, et al. Circular RNA circ-CBFB promotes proliferation and inhibits apoptosis in chronic lymphocytic leukemia through regulating miR-607/FZD3/Wnt/beta-catenin pathway. Biochem Biophys Res Commun. 2018;503(1):385–90.

    CAS  PubMed  Google Scholar 

  65. Wu W, et al. Downregulation of circ_0132266 in chronic lymphocytic leukemia promoted cell viability through miR-337-3p/PML axis. Aging. 2019;11(11):3561–73.

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Shang J, et al. CircPAN3 mediates drug resistance in acute myeloid leukemia through the miR-153-5p/miR-183-5p-XIAP axis. Exp Hematol. 2019;70:42–54.e3.

    CAS  PubMed  Google Scholar 

  67. Shang J, et al. CircPAN3 contributes to drug resistance in acute myeloid leukemia through regulation of autophagy. Leuk Res. 2019;85:106198.

    CAS  PubMed  Google Scholar 

  68. Ping L, et al. Silencing of circ_0009910 inhibits acute myeloid leukemia cell growth through increasing miR-20a-5p. Blood Cell Mol Dis. 2019;75:41–7.

    Google Scholar 

  69. Ping L, et al. High circ_100053 predicts a poor outcome for chronic myeloid leukemia and is involved in imatinib resistance. Oncol Res. 2019. https://doi.org/10.3727/096504018X15412701483326.

  70. Liu J, et al. Global identification of circular RNAs in chronic myeloid leukemia reveals hsa_circ_0080145 regulates cell proliferation by sponging miR-29b. Biochem Biophys Res Commun. 2018;504(4):660–5.

    CAS  PubMed  Google Scholar 

  71. Chen H, et al. Circ-ANAPC7 is upregulated in acute myeloid Leukemia and appears to target the MiR-181 family. Cell Physiol Biochem. 2018;47(5):1998–2007.

    CAS  PubMed  Google Scholar 

  72. Pan Y, et al. CircBA9.3 supports the survival of leukaemic cells by up-regulating c-ABL1 or BCR-ABL1 protein levels. Blood Cells Mol Dis. 2018;73:38–44.

    CAS  PubMed  Google Scholar 

  73. Xiao J, et al. SKP2 siRNA inhibits the degradation of P27kip1 and down-regulates the expression of MRP in HL-60/a cells. Acta Biochim Biophys Sin Shanghai. 2009;41(8):699–708.

    CAS  PubMed  Google Scholar 

  74. Collins LJ, Penny D. The RNA infrastructure: dark matter of the eukaryotic cell? Trends Genet. 2009;25(3):120–8.

    CAS  PubMed  Google Scholar 

  75. Morris KV, Mattick JS. The rise of regulatory RNA. Nat Rev Genet. 2014;15:423.

    CAS  PubMed  PubMed Central  Google Scholar 

  76. Kung JTY, Colognori D, Lee JT. Long noncoding RNAs: past, present, and future. Genetics. 2013;193(3):651–69.

    CAS  PubMed  PubMed Central  Google Scholar 

  77. Palazzo AF, Lee ES. Non-coding RNA: what is functional and what is junk? Front Genet. 2015;6:2.

    PubMed  PubMed Central  Google Scholar 

  78. Iyer MK, et al. The landscape of long noncoding RNAs in the human transcriptome. Nat Genet. 2015;47(3):199–208.

    CAS  PubMed  PubMed Central  Google Scholar 

  79. Bartonicek N, Maag JLV, Dinger ME. Long noncoding RNAs in cancer: mechanisms of action and technological advancements. Mol Cancer. 2016;15(1):43.

    PubMed  PubMed Central  Google Scholar 

  80. Dhanoa JK, et al. Long non-coding RNA: its evolutionary relics and biological implications in mammals: a review. J Anim Sci Technol. 2018;60:25.

    CAS  PubMed  PubMed Central  Google Scholar 

  81. Prensner JR, Chinnaiyan AM. The emergence of lncRNAs in cancer biology. Cancer Discov. 2011;1(5):391–407.

    CAS  PubMed  PubMed Central  Google Scholar 

  82. Wilusz JE, Sunwoo H, Spector DL. Long noncoding RNAs: functional surprises from the RNA world. Genes Dev. 2009;23(13):1494–504.

    CAS  PubMed  PubMed Central  Google Scholar 

  83. Li T, et al. Molecular mechanisms of long noncoding RNAs on gastric cancer. Oncotarget. 2016;7(8):8601–12.

    PubMed  PubMed Central  Google Scholar 

  84. Chen L, et al. Long Non-Coding RNA CCAT1 acts as a competing endogenous RNA to regulate cell growth and differentiation in acute myeloid Leukemia. Mol Cell. 2016;39(4):330–6.

    CAS  Google Scholar 

  85. Merry CR, et al. DNMT1-associated long non-coding RNAs regulate global gene expression and DNA methylation in colon cancer. Hum Mol Genet. 2015;24(21):6240–53.

    CAS  PubMed  PubMed Central  Google Scholar 

  86. Lee RS, Roberts CWM. Linking the SWI/SNF complex to prostate cancer. Nat Genet. 2013;45(11):1268–9.

    CAS  PubMed  Google Scholar 

  87. Arab K, et al. Long noncoding RNA TARID directs demethylation and activation of the tumor suppressor TCF21 via GADD45A. Mol Cell. 2014;55(4):604–14.

    CAS  PubMed  Google Scholar 

  88. Davidovich C, Cech TR. The recruitment of chromatin modifiers by long noncoding RNAs: lessons from PRC2. RNA. 2015;21(12):2007–22.

    CAS  PubMed  PubMed Central  Google Scholar 

  89. Wang KC, et al. A long noncoding RNA maintains active chromatin to coordinate homeotic gene expression. Nature. 2011;472(7341):120–4.

    CAS  PubMed  PubMed Central  Google Scholar 

  90. Lai F, et al. Activating RNAs associate with mediator to enhance chromatin architecture and transcription. Nature. 2013;494(7438):497–501.

    CAS  PubMed  PubMed Central  Google Scholar 

  91. Yang F, et al. The lncRNA firre anchors the inactive X chromosome to the nucleolus by binding CTCF and maintains H3K27me3 methylation. Genome Biol. 2015;16(1):52.

    PubMed  PubMed Central  Google Scholar 

  92. Schmitt AM, Chang HY. Long noncoding RNAs in Cancer pathways. Cancer Cell. 2016;29(4):452–63.

    CAS  PubMed  PubMed Central  Google Scholar 

  93. Pouyanrad S, Rahgozar S, Ghodousi ES. Dysregulation of miR-335-3p, targeted by NEAT1 and MALAT1 long non-coding RNAs, is associated with poor prognosis in childhood acute lymphoblastic leukemia. Gene. 2019;692:35–43.

    CAS  PubMed  Google Scholar 

  94. El-Khazragy N, et al. The prognostic significance of the long non-coding RNAs "CCAT1, PVT1" in t (8;21) associated acute myeloid Leukemia. Gene. 2019;707:172–7.

    CAS  PubMed  Google Scholar 

  95. Tang P, et al. A 10-Long Non-Coding RNA-based expression signature as a potential biomarker for prognosis of acute myeloid Leukemia. Med Sci Monit. 2019;25:4999–5004.

    CAS  PubMed  PubMed Central  Google Scholar 

  96. Dong X, et al. Knockdown of Long noncoding RNA HOXA-AS2 suppresses Chemoresistance of acute myeloid Leukemia via the miR-520c-3p/S100A4 Axis. Cell Physiol Biochem. 2018;51(2):886–96.

    CAS  PubMed  Google Scholar 

  97. Miller CR, et al. The long noncoding RNA, treRNA, decreases DNA damage and is associated with poor response to chemotherapy in chronic lymphocytic leukemia. Oncotarget. 2017;8(16):25942–54.

    PubMed  PubMed Central  Google Scholar 

  98. Hughes JM, et al. C/EBPalpha-p30 protein induces expression of the oncogenic long non-coding RNA UCA1 in acute myeloid leukemia. Oncotarget. 2015;6(21):18534–44.

    PubMed  PubMed Central  Google Scholar 

  99. Fang K, et al. A distinct set of long non-coding RNAs in childhood MLL-rearranged acute lymphoblastic leukemia: biology and epigenetic target. Hum Mol Genet. 2014;23(12):3278–88.

    CAS  PubMed  Google Scholar 

  100. Conn SJ, et al. The RNA Binding Protein Quaking Regulates Formation of circRNAs. Cell. 2015;160(6):1125–34.

    CAS  PubMed  Google Scholar 

  101. You X, et al. Neural circular RNAs are derived from synaptic genes and regulated by development and plasticity. Nat Neurosci. 2015;18(4):603–10.

    CAS  PubMed  PubMed Central  Google Scholar 

  102. Li X, et al. Long noncoding RNAs in head and neck cancer. Oncotarget. 2017;8(6):10726–40.

    PubMed  Google Scholar 

  103. Perron MP, Provost P. Protein interactions and complexes in human microRNA biogenesis and function. Front Biosci. 2008;13:2537–47.

    CAS  PubMed  PubMed Central  Google Scholar 

  104. Wightman B, Ha I, Ruvkun G. Posttranscriptional regulation of the heterochronic gene lin-14 by lin-4 mediates temporal pattern formation in C. elegans. Cell. 1993;75(5):855–62.

    CAS  PubMed  Google Scholar 

  105. Lee RC, Feinbaum RL, Ambros V. The C. elegans heterochronic gene lin-4 encodes small RNAs with antisense complementarity to lin-14. Cell. 1993;75(5):843–54.

    CAS  PubMed  Google Scholar 

  106. Rajewsky N. L (ou) sy miRNA targets? Nat Struct Mol Biol. 2006;13(9):754–5.

    CAS  PubMed  Google Scholar 

  107. Rajewsky N. microRNA target predictions in animals. Nat Genet. 2006;38(Suppl):S8–13.

    CAS  PubMed  Google Scholar 

  108. Liu W, Mao SY, Zhu WY. Impact of tiny miRNAs on cancers. World J Gastroenterol. 2007;13(4):497–502.

    CAS  PubMed  PubMed Central  Google Scholar 

  109. Berezikov E, et al. Phylogenetic shadowing and computational identification of human microRNA genes. Cell. 2005;120(1):21–4.

    CAS  PubMed  Google Scholar 

  110. Stanczyk J, et al. Altered expression of MicroRNA in synovial fibroblasts and synovial tissue in rheumatoid arthritis. Arthritis Rheum. 2008;58(4):1001–9.

    PubMed  Google Scholar 

  111. Griffiths-Jones S. The microRNA registry. Nucleic Acids Res. 2004;32(Database issue):D109–11.

    CAS  PubMed  PubMed Central  Google Scholar 

  112. Griffiths-Jones S, et al. miRBase: tools for microRNA genomics. Nucleic Acids Res. 2008;36(Database issue):D154–8.

    CAS  PubMed  Google Scholar 

  113. Cullen BR. Transcription and processing of human microRNA precursors. Mol Cell. 2004;16(6):861–5.

    CAS  PubMed  Google Scholar 

  114. Esau C, et al. miR-122 regulation of lipid metabolism revealed by in vivo antisense targeting. Cell Metab. 2006;3(2):87–98.

    CAS  PubMed  Google Scholar 

  115. Krutzfeldt J, et al. Silencing of microRNAs in vivo with 'antagomirs'. Nature. 2005;438(7068):685–9.

    PubMed  Google Scholar 

  116. Zhao Y, Samal E, Srivastava D. Serum response factor regulates a muscle-specific microRNA that targets Hand2 during cardiogenesis. Nature. 2005;436(7048):214–20.

    CAS  PubMed  Google Scholar 

  117. Garzon R, et al. MicroRNA fingerprints during human megakaryocytopoiesis. Proc Natl Acad Sci U S A. 2006;103(13):5078–83.

    CAS  PubMed  PubMed Central  Google Scholar 

  118. Monticelli S, et al. MicroRNA profiling of the murine hematopoietic system. Genome Biol. 2005;6(8):R71.

    PubMed  PubMed Central  Google Scholar 

  119. Ambros V. The functions of animal microRNAs. Nature. 2004;431(7006):350–5.

    CAS  PubMed  Google Scholar 

  120. Lee YS, Dutta A. MicroRNAs: small but potent oncogenes or tumor suppressors. Curr Opin Investig Drugs. 2006;7(6):560–4.

    CAS  PubMed  Google Scholar 

  121. Zhi Y, et al. Serum level of miR-10-5p as a prognostic biomarker for acute myeloid leukemia. Int J Hematol. 2015;102(3):296–303.

    CAS  PubMed  Google Scholar 

  122. Zhao J, et al. Prognostic value of miR-96 in patients with acute myeloid leukemia. Diagn Pathol. 2014;9:76.

    PubMed  PubMed Central  Google Scholar 

  123. Shrestha S, et al. A systematic review of microRNA expression profiling studies in human gastric cancer. Cancer Med. 2014;3(4):878–88.

    CAS  PubMed  PubMed Central  Google Scholar 

  124. Swartling FJ, et al. Deregulated proliferation and differentiation in brain tumors. Cell Tissue Res. 2015;359(1):225–54.

    CAS  PubMed  Google Scholar 

  125. Takahashi RU, Miyazaki H, Ochiya T. The roles of MicroRNAs in breast Cancer. Cancers (Basel). 2015;7(2):598–616.

    CAS  Google Scholar 

  126. Xu YM, et al. Regulation of miRNAs affects radiobiological response of lung cancer stem cells. Biomed Res Int. 2015;2015:851841.

    PubMed  PubMed Central  Google Scholar 

  127. Nicolaidou V, Koufaris C. MicroRNA responses to environmental liver carcinogens: Biological and clinical significance. Clin Chim Acta. 2015;445:25–33.

    CAS  PubMed  Google Scholar 

  128. Rodriguez-Montes JA, Menendez Sanchez P. Role of micro-RNA in colorectal cancer screening. Cir Esp. 2014;92(10):654–8.

    PubMed  Google Scholar 

  129. Bottoni A, Calin GA. MicroRNAs as main players in the pathogenesis of chronic lymphocytic leukemia. Microrna. 2014;2(3):158–64.

    PubMed  Google Scholar 

  130. Wang W, et al. MicroRNA profiling of follicular lymphoma identifies microRNAs related to cell proliferation and tumor response. Haematologica. 2012;97(4):586–94.

    CAS  PubMed  PubMed Central  Google Scholar 

  131. Jongen-Lavrencic M, et al. MicroRNA expression profiling in relation to the genetic heterogeneity of acute myeloid leukemia. Blood. 2008;111(10):5078–85.

    CAS  PubMed  Google Scholar 

  132. Lagos-Quintana M, et al. Identification of tissue-specific microRNAs from mouse. Curr Biol. 2002;12(9):735–9.

    CAS  PubMed  Google Scholar 

  133. Fuster O, et al. Adverse prognostic value of MYBL2 overexpression and association with microRNA-30 family in acute myeloid leukemia patients. Leuk Res. 2013;37(12):1690–6.

    CAS  PubMed  Google Scholar 

  134. Yamamoto H, et al. miR-133 regulates Evi1 expression in AML cells as a potential therapeutic target. Sci Rep. 2016;6:19204.

    CAS  PubMed  PubMed Central  Google Scholar 

  135. Xiao Y, Su C, Deng T. miR-223 decreases cell proliferation and enhances cell apoptosis in acute myeloid leukemia via targeting FBXW7. Oncol Lett. 2016;12(5):3531–6.

    CAS  PubMed  PubMed Central  Google Scholar 

  136. Pulikkan JA, et al. Cell-cycle regulator E2F1 and microRNA-223 comprise an autoregulatory negative feedback loop in acute myeloid leukemia. Blood. 2010;115(9):1768–78.

    CAS  PubMed  PubMed Central  Google Scholar 

  137. Lin X, et al. Serum MicroRNA-370 as a potential diagnostic and prognostic biomarker for pediatric acute myeloid leukemia. Int J Clin Exp Pathol. 2015;8(11):14658–66.

    CAS  PubMed  PubMed Central  Google Scholar 

  138. Magee P, Shi L, Garofalo M. Role of microRNAs in chemoresistance. Ann Transl Med. 2015;3(21):332.

    PubMed  PubMed Central  Google Scholar 

  139. Liu X, et al. miR-181a promotes G1/S transition and cell proliferation in pediatric acute myeloid leukemia by targeting ATM. J Cancer Res Clin Oncol. 2016;142(1):77–87.

    CAS  PubMed  Google Scholar 

  140. Hugo S, et al. Effect of miR-128 in DNA damage of HL-60 acute myeloid Leukemia cells. Curr Pharm Biotechnol. 2014;15(5):492–502.

    Google Scholar 

  141. Volinia S, et al. A microRNA expression signature of human solid tumors defines cancer gene targets. Proc Natl Acad Sci U S A. 2006;103(7):2257–61.

    CAS  PubMed  PubMed Central  Google Scholar 

  142. Jiang X, et al. MicroRNA-221 sensitizes chronic myeloid leukemia cells to imatinib by targeting STAT5. Leuk Lymphoma. 2019;60(7):1709–20.

    CAS  PubMed  Google Scholar 

  143. Jin J, et al. Decreased expression of microRNA-214 contributes to imatinib mesylate resistance of chronic myeloid leukemia patients by upregulating ABCB1 gene expression. Exp Ther Med. 2018;16(3):1693–700.

    PubMed  PubMed Central  Google Scholar 

  144. Rokah OH, et al. Downregulation of miR-31, miR-155, and miR-564 in chronic myeloid leukemia cells. PLoS One. 2012;7(4):e35501.

    CAS  PubMed  Google Scholar 

  145. Li Y, et al. Inhibition of BCR/ABL protein expression by miR-203 sensitizes for imatinib mesylate. PLoS One. 2013;8(4):e61858.

    CAS  PubMed  PubMed Central  Google Scholar 

  146. Salzman J, et al. Cell-type specific features of circular RNA expression. PLoS Genet. 2013;9(9):e1003777.

    CAS  PubMed  PubMed Central  Google Scholar 

  147. Jeck WR, et al. Circular RNAs are abundant, conserved, and associated with ALU repeats. RNA. 2013;19(2):141–57.

    CAS  PubMed  PubMed Central  Google Scholar 

  148. Memczak S, et al. Circular RNAs are a large class of animal RNAs with regulatory potency. Nature. 2013;495(7441):333–8.

    CAS  PubMed  Google Scholar 

  149. Danan M, et al. Transcriptome-wide discovery of circular RNAs in archaea. Nucleic Acids Res. 2012;40(7):3131–42.

    CAS  PubMed  Google Scholar 

  150. Wang PL, et al. Circular RNA is expressed across the eukaryotic tree of life. PLoS One. 2014;9(6):e90859.

    PubMed  PubMed Central  Google Scholar 

  151. Huang S, et al. The emerging role of circular RNAs in transcriptome regulation. Genomics. 2017;109(5):401–7.

    CAS  PubMed  Google Scholar 

  152. Ebbesen KK, Kjems J, Hansen TB. Circular RNAs: identification, biogenesis and function. Biochim Biophys Acta. 2016;1859(1):163–8.

    CAS  PubMed  Google Scholar 

  153. Chen L-L. The biogenesis and emerging roles of circular RNAs. Nat Rev Mol Cell Biol. 2016;17(4):205–11.

    CAS  PubMed  Google Scholar 

  154. Barrett SP, Salzman J. Circular RNAs: analysis, expression and potential functions. Development. 2016;143(11):1838.

    CAS  PubMed  PubMed Central  Google Scholar 

  155. Vicens Q, Westhof E. Biogenesis of circular RNAs. Cell. 2014;159(1):13–4.

    CAS  PubMed  Google Scholar 

  156. Hansen TB, et al. Natural RNA circles function as efficient microRNA sponges. Nature. 2013;495(7441):384–8.

    CAS  PubMed  Google Scholar 

  157. Piwecka M, et al. Loss of a mammalian circular RNA locus causes miRNA deregulation and affects brain function. Science. 2017;357(6357):eaam8526.

    PubMed  Google Scholar 

  158. Tang W, et al. Silencing CDR1as inhibits colorectal cancer progression through regulating microRNA-7. Onco Targets Ther. 2017;10:2045–56.

    CAS  PubMed  PubMed Central  Google Scholar 

  159. Yu L, et al. The circular RNA Cdr1as act as an oncogene in hepatocellular carcinoma through targeting miR-7 expression. PLoS One. 2016;11(7):e0158347.

    PubMed  PubMed Central  Google Scholar 

  160. Guo JU, et al. Expanded identification and characterization of mammalian circular RNAs. Genome Biol. 2014;15(7):409.

    PubMed  PubMed Central  Google Scholar 

  161. Li J, et al. Circular RNAs in cancer: novel insights into origins, properties, functions and implications. Am J Cancer Res. 2015;5(2):472–80.

    CAS  PubMed  PubMed Central  Google Scholar 

  162. Du WW, et al. Foxo3 circular RNA retards cell cycle progression via forming ternary complexes with p21 and CDK2. Nucleic Acids Res. 2016;44(6):2846–58.

    PubMed  PubMed Central  Google Scholar 

  163. Han D, et al. Circular RNA circMTO1 acts as the sponge of microRNA-9 to suppress hepatocellular carcinoma progression. Hepatology. 2017;66(4):1151–64.

    CAS  PubMed  Google Scholar 

  164. Jeck WR, Sharpless NE. Detecting and characterizing circular RNAs. Nat Biotechnol. 2014;32(5):453–61.

    CAS  PubMed  PubMed Central  Google Scholar 

  165. Hentze MW, Preiss T. Circular RNAs: splicing's enigma variations. EMBO J. 2013;32(7):923–5.

    CAS  PubMed  PubMed Central  Google Scholar 

  166. Yang Y, et al. Extensive translation of circular RNAs driven by N (6)-methyladenosine. Cell Res. 2017;27(5):626–41.

    CAS  PubMed  PubMed Central  Google Scholar 

  167. Pamudurti NR, et al. Translation of CircRNAs. Mol Cell. 2017;66(1):9–21.e7.

    CAS  PubMed  PubMed Central  Google Scholar 

  168. Wang Y, Wang Z. Efficient backsplicing produces translatable circular mRNAs. RNA. 2015;21(2):172–9.

    CAS  PubMed  PubMed Central  Google Scholar 

  169. Legnini I, et al. Circ-ZNF609 is a circular RNA that can be translated and functions in Myogenesis. Mol Cell. 2017;66(1):22–37.e9.

    CAS  PubMed  PubMed Central  Google Scholar 

  170. Memczak S, et al. Identification and characterization of circular RNAs as a new class of putative biomarkers in human blood. PLoS One. 2015;10(10):e0141214.

    PubMed  PubMed Central  Google Scholar 

  171. Xia L, et al. Circular RNA circ-CBFB promotes proliferation and inhibits apoptosis in chronic lymphocytic leukemia through regulating miR-607/FZD3/Wnt/β-catenin pathway. Biochem Biophys Res Commun. 2018;503(1):385–90.

    CAS  PubMed  Google Scholar 

  172. Ahmed I, et al. Altered expression pattern of circular RNAs in primary and metastatic sites of epithelial ovarian carcinoma. Oncotarget. 2016;7(24):36366–81.

    PubMed  PubMed Central  Google Scholar 

  173. Ahmed I, et al. Identification of human genetic variants controlling circular RNA expression. RNA. 2019;25(12):1765–78.

    CAS  PubMed  PubMed Central  Google Scholar 

  174. Chang JE, Kahl BS. Bendamustine for treatment of chronic lymphocytic leukemia. Expert Opin Pharmacother. 2012;13(10):1495–505.

    CAS  PubMed  Google Scholar 

  175. Foà R, et al. Chlorambucil plus rituximab with or without maintenance rituximab as first-line treatment for elderly chronic lymphocytic leukemia patients. Am J Hematol. 2014;89(5):480–6.

    PubMed  Google Scholar 

  176. Zenz T, et al. miR-34a as part of the resistance network in chronic lymphocytic leukemia. Blood. 2009;113(16):3801–8.

    CAS  PubMed  Google Scholar 

  177. Asslaber D, et al. microRNA-34a expression correlates with MDM2 SNP309 polymorphism and treatment-free survival in chronic lymphocytic leukemia. Blood. 2010;115(21):4191–7.

    CAS  PubMed  Google Scholar 

  178. Zhu D-X, et al. miR-181a/b significantly enhances drug sensitivity in chronic lymphocytic leukemia cells via targeting multiple anti-apoptosis genes. Carcinogenesis. 2012;33(7):1294–301.

    CAS  PubMed  Google Scholar 

  179. Lu F, et al. miR-181b increases drug sensitivity in acute myeloid leukemia via targeting HMGB1 and MCL-1. Int J Oncol. 2014;45:383–92.

    CAS  PubMed  Google Scholar 

  180. Zhou L, et al. microRNA-125b promotes leukemia cell resistance to daunorubicin by inhibiting apoptosis. Mol Med Rep. 2014;9:1909–16.

    CAS  PubMed  Google Scholar 

  181. Si W, et al. The role and mechanisms of action of microRNAs in cancer drug resistance. Clin Epigenetics. 2019;11(1):25.

    PubMed  PubMed Central  Google Scholar 

  182. Hershkovitz-Rokah O, et al. MiR-30e induces apoptosis and sensitizes K562 cells to imatinib treatment via regulation of the BCR–ABL protein. Cancer Lett. 2015;356(2, Part B):597–605.

    CAS  PubMed  Google Scholar 

  183. Wang L-S, et al. MicroRNA-486 regulates normal erythropoiesis and enhances growth and modulates drug response in CML progenitors. Blood. 2015;125(8):1302–13.

    CAS  PubMed  PubMed Central  Google Scholar 

  184. Löwenberg B, et al. High-dose Daunorubicin in older patients with acute myeloid Leukemia. N Engl J Med. 2009;361(13):1235–48.

    PubMed  Google Scholar 

  185. Döhner H, et al. Diagnosis and management of acute myeloid leukemia in adults: recommendations from an international expert panel, on behalf of the European LeukemiaNet. Blood. 2010;115(3):453–74.

    PubMed  Google Scholar 

  186. Estey E. AML in older patients: are we making progress? Best Pract Res Clin Haematol. 2009;22(4):529–36.

    PubMed  Google Scholar 

  187. D'Ugo E, et al. Identification of scaffold/matrix attachment region in recurrent site of woodchuck hepatitis virus integration. DNA Cell Biol. 1998;17(6):519–27.

    CAS  PubMed  Google Scholar 

  188. Sumer H, et al. A rapid method of genomic array analysis of scaffold/matrix attachment regions (S/MARs) identifies a 2.5-Mb region of enhanced scaffold/matrix attachment at a human neocentromere. Genome Res. 2003;13(7):1737–43.

    CAS  PubMed  PubMed Central  Google Scholar 

  189. Barone G, et al. Intercalation of Daunomycin into stacked DNA Base pairs. DFT Study of an Anticancer Drug. J Biomol Struct Dyn. 2008;26:115–30.

    CAS  PubMed  Google Scholar 

  190. Chikamori K, et al. Downregulation of topoisomerase IIβ in myeloid leukemia cell lines leads to activation of apoptosis following all-trans retinoic acid-induced differentiation/growth arrest. Leukemia. 2006;20(10):1809–18.

    CAS  PubMed  Google Scholar 

  191. Hermanson DL, et al. Overexpression of mcl-1 confers multidrug resistance, whereas topoisomerase IIβ downregulation introduces mitoxantrone-specific drug resistance in acute myeloid leukemia. Mol Pharmacol. 2013;84(2):236–43.

    CAS  PubMed  PubMed Central  Google Scholar 

  192. Heidenreich E, et al. Non-homologous end joining as an important mutagenic process in cell cycle-arrested cells. EMBO J. 2003;22(9):2274–83.

    CAS  PubMed  PubMed Central  Google Scholar 

  193. Hensel JP, et al. Breakpoints of t (4;11) translocations in the human MLL and AF4 genes in ALL patients are preferentially clustered outside of high-affinity matrix attachment regions. J Cell Biochem. 2001;82(2):299–309.

    CAS  PubMed  Google Scholar 

  194. Domer P, et al. Molecular analysis of 13 cases of MLL/11q23 secondary acute leukemia and identification of topoisomerase II consensus-binding sequences near the chromosomal breakpoint of a secondary leukemia with the t (4;11). Leukemia. 1995;9:1305–12.

    CAS  PubMed  Google Scholar 

  195. Chavali PL, Funa K, Chavali S. Cis-regulation of microRNA expression by scaffold/matrix-attachment regions. Nucleic Acids Res. 2011;39(16):6908–18.

    CAS  PubMed  PubMed Central  Google Scholar 

  196. Lai T-H, et al. HDAC inhibition induces MicroRNA-182, which targets RAD51 and impairs HR repair to sensitize cells to Sapacitabine in acute myelogenous Leukemia. Clin Cancer Res. 2016;22(14):3537–49.

    CAS  PubMed  PubMed Central  Google Scholar 

  197. Garzon R, et al. Distinctive microRNA signature of acute myeloid leukemia bearing cytoplasmic mutated nucleophosmin. Proc Natl Acad Sci U S A. 2008;105(10):3945–50.

    CAS  PubMed  PubMed Central  Google Scholar 

  198. Zhang H, et al. Upregulation of microRNA-125b contributes to leukemogenesis and increases drug resistance in pediatric acute promyelocytic leukemia. Mol Cancer. 2011;10:108.

    CAS  PubMed  PubMed Central  Google Scholar 

  199. Kibria G, Hatakeyama H, Harashima H. Cancer multidrug resistance: mechanisms involved and strategies for circumvention using a drug delivery system. Arch Pharm Res. 2014;37(1):4–15.

    CAS  PubMed  Google Scholar 

  200. de Oliveira JC, et al. Differential MiRNA expression in childhood acute lymphoblastic leukemia and association with clinical and biological features. Leuk Res. 2012;36(3):293–8.

    PubMed  Google Scholar 

  201. Gefen N, et al. Hsa-mir-125b-2 is highly expressed in childhood ETV6/RUNX1 (TEL/AML1) leukemias and confers survival advantage to growth inhibitory signals independent of p53. Leukemia. 2010;24(1):89–96.

    CAS  PubMed  Google Scholar 

  202. Schotte D, et al. MicroRNA characterize genetic diversity and drug resistance in pediatric acute lymphoblastic leukemia. Haematologica. 2011;96(5):703.

    PubMed  PubMed Central  Google Scholar 

  203. Wang H, et al. The effects of microRNAs on glucocorticoid responsiveness. J Cancer Res Clin Oncol. 2017;143(6):1005–11.

    CAS  PubMed  Google Scholar 

  204. Yan J, et al. Deregulated MIR335 that targets MAPK1 is implicated in poor outcome of paediatric acute lymphoblastic leukaemia. Br J Haematol. 2013;163(1):93–103.

    CAS  PubMed  Google Scholar 

  205. Mei Y, et al. Effect of microRNA-210 on prognosis and response to chemotherapeutic drugs in pediatric acute lymphoblastic leukemia. Cancer Sci. 2014;105(4):463–72.

    CAS  PubMed  PubMed Central  Google Scholar 

  206. Luo M, et al. Long non-coding RNAs control hematopoietic stem cell function. Cell Stem Cell. 2015;16(4):426–38.

    CAS  PubMed  PubMed Central  Google Scholar 

  207. Zhang X, et al. A myelopoiesis-associated regulatory intergenic noncoding RNA transcript within the human HOXA cluster. Blood. 2009;113(11):2526–34.

    CAS  PubMed  PubMed Central  Google Scholar 

  208. Wei S, et al. PU.1 controls the expression of long noncoding RNA HOTAIRM1 during granulocytic differentiation. J Hematol Oncol. 2016;9(1):44.

    PubMed  PubMed Central  Google Scholar 

  209. Wang H, et al. Transcription factors IRF8 and PU.1 are required for follicular B cell development and BCL6-driven germinal center responses. Proc Natl Acad Sci U S A. 2019;116(19):9511–20.

    CAS  PubMed  PubMed Central  Google Scholar 

  210. Wang P, et al. The STAT3-binding long noncoding RNA lnc-DC controls human dendritic cell differentiation. Science (New York, NY). 2014;344(6181):310–3.

    CAS  Google Scholar 

  211. Hirschberger S, Hinske LC, Kreth S. MiRNAs: dynamic regulators of immune cell functions in inflammation and cancer. Cancer Lett. 2018;431:11–21.

    CAS  PubMed  Google Scholar 

  212. Schwarzer A, et al. The non-coding RNA landscape of human hematopoiesis and leukemia. Nat Commun. 2017;8:218. https://doi.org/10.1038/s41467-017-00212-4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  213. Yildirim E, et al. Xist RNA is a potent suppressor of hematologic cancer in mice. Cell. 2013;152(4):727–42.

    CAS  PubMed  Google Scholar 

  214. Léveillé N, et al. Genome-wide profiling of p53-regulated enhancer RNAs uncovers a subset of enhancers controlled by a lncRNA. Nat Commun. 2015;6:6520.

    PubMed  Google Scholar 

  215. Reisman D, et al. Evidence for a role of a lncRNA encoded from the p53 tumor suppressor gene in maintaining the undifferentiated state of human myeloid leukemias. Gene Reports. 2016;5:45–50.

    Google Scholar 

  216. Blume CJ, et al. p53-dependent non-coding RNA networks in chronic lymphocytic leukemia. Leukemia. 2015;29(10):2015–23.

    CAS  PubMed  Google Scholar 

  217. Zeng C, et al. Inhibition of long non-coding RNA NEAT1 impairs myeloid differentiation in acute promyelocytic leukemia cells. BMC Cancer. 2014;14:693.

    PubMed  PubMed Central  Google Scholar 

  218. Hayashi Y, et al. Antitumor immunity augments the therapeutic effects of p53 activation on acute myeloid leukemia. Nat Commun. 2019;10:4869. https://doi.org/10.1038/s41467-019-12555-1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  219. Brazão TF, et al. Long noncoding RNAs in B-cell development and activation. Blood. 2016;128(7):e10–9.

    PubMed  PubMed Central  Google Scholar 

  220. Liu Y, et al. Cloning of two candidate tumor suppressor genes within a 10 kb region on chromosome 13q14, frequently deleted in chronic lymphocytic leukemia. Oncogene. 1997;15(20):2463–73.

    CAS  PubMed  Google Scholar 

  221. Stilgenbauer S, et al. Expressed sequences as candidates for a novel tumor suppressor gene at band 13q14 in B-cell chronic lymphocytic leukemia and mantle cell lymphoma. Oncogene. 1998;16(14):1891–7.

    CAS  PubMed  Google Scholar 

  222. Garding A, et al. Epigenetic upregulation of lncRNAs at 13q14.3 in leukemia is linked to the In Cis downregulation of a gene cluster that targets NF-kB. PLoS Genet. 2013;9(4):e1003373.

    CAS  PubMed  PubMed Central  Google Scholar 

  223. Mertens D, et al. Chronic lymphocytic leukemia and 13q14: miRs and more. Leuk Lymphoma. 2009;50(3):502–5.

    CAS  PubMed  Google Scholar 

  224. Herishanu Y, et al. The lymph node microenvironment promotes B-cell receptor signaling, NF-kappaB activation, and tumor proliferation in chronic lymphocytic leukemia. Blood. 2011;117(2):563–74.

    CAS  PubMed  PubMed Central  Google Scholar 

  225. Krawczyk M, Emerson BM. p50-associated COX-2 extragenic RNA (PACER) activates COX-2 gene expression by occluding repressive NF-κB complexes. eLife. 2014;3:e01776.

    PubMed  PubMed Central  Google Scholar 

  226. Hu G, et al. LincRNA-Cox2 Promotes Late Inflammatory Gene Transcription in Macrophages through Modulating SWI/SNF-Mediated Chromatin Remodeling. J Immunol (Baltimore, Md : 1950). 2016;196(6):2799–808.

    CAS  Google Scholar 

  227. Hoesel B, Schmid JA. The complexity of NF-κB signaling in inflammation and cancer. Mol Cancer. 2013;12:86.

    CAS  PubMed  PubMed Central  Google Scholar 

  228. Trimarchi T, et al. Genome-wide mapping and characterization of notch-regulated long noncoding RNAs in acute leukemia. Cell. 2014;158(3):593–606.

    CAS  PubMed  PubMed Central  Google Scholar 

  229. Wang Y, et al. LncRNA NALT interaction with NOTCH1 promoted cell proliferation in pediatric T cell acute lymphoblastic leukemia. Sci Rep. 2015;5:13749. https://doi.org/10.1038/srep13749.

    Article  PubMed  PubMed Central  Google Scholar 

  230. Janghorban M, et al. Notch signaling as a regulator of the tumor immune response: to target or not to target? Front Immunol. 2018;9:1649.

    PubMed  PubMed Central  Google Scholar 

  231. Piro G, et al. Revising PTEN in the era of immunotherapy: new perspectives for an old story. Cancers. 2019;11(10):1525.

    CAS  PubMed Central  Google Scholar 

  232. Guo G, et al. High expression of long non-coding RNA H19 is required for efficient tumorigenesis induced by Bcr-Abl oncogene. FEBS Lett. 2014;588(9):1780–6.

    CAS  PubMed  Google Scholar 

  233. Tseng Y-Y, et al. PVT1 dependence in cancer with MYC copy-number increase. Nature. 2014;512(7512):82–6.

    CAS  PubMed  PubMed Central  Google Scholar 

  234. Casey SC, et al. MYC regulates the antitumor immune response through CD47 and PD-L1. Science. 2016;352(6282):227.

    CAS  PubMed  PubMed Central  Google Scholar 

  235. Sun Y-M, et al. circMYBL2, a circRNA from MYBL2, regulates FLT3 translation by recruiting PTBP1 to promote FLT3-ITD AML progression. Blood. 2019;134(18):1533–46.

    CAS  PubMed  PubMed Central  Google Scholar 

  236. Nicolet BP, et al. Circular RNA expression in human hematopoietic cells is widespread and cell-type specific. Nucleic Acids Res. 2018;46(16):8168–80.

    CAS  PubMed  PubMed Central  Google Scholar 

  237. Wu Y, et al. MicroRNA-17-92 controls T-cell responses in graft-versus-host disease and leukemia relapse in mice. Blood. 2015;126(11):1314–23.

    CAS  PubMed  PubMed Central  Google Scholar 

  238. Dakir E-H, Mollinedo F. Genome-wide miRNA profiling and pivotal roles of miRs 125a-5p and 17-92 cluster in human neutrophil maturation and differentiation of acute myeloid leukemia cells. Oncotarget. 2019;10(51):5313–31.

    PubMed Central  Google Scholar 

  239. Schneider E, et al. MicroRNA-708 is a novel regulator of the Hoxa9 program in myeloid cells. Leukemia. 2019. https://doi.org/10.1038/s41375-019-0651-1.

  240. Wang X, et al. Noncoding RNAs and their functional involvement in regulation of chronic myeloid leukemia. Brief Funct Genomics. 2015;15(3):239–48.

    PubMed  Google Scholar 

  241. Yeh C-H, Moles R, Nicot C. Clinical significance of microRNAs in chronic and acute human leukemia. Mol Cancer. 2016;15(1):37.

    PubMed  PubMed Central  Google Scholar 

  242. Wang XQD, Dostie J. Reciprocal regulation of chromatin state and architecture by HOTAIRM1 contributes to temporal collinear HOXA gene activation. Nucleic Acids Res. 2017;45(3):1091–104.

    CAS  PubMed  Google Scholar 

  243. Zhao H, et al. HOX antisense lincRNA HOXA-AS2 is an apoptosis repressor in all trans retinoic acid treated NB4 promyelocytic leukemia cells. J Cell Biochem. 2013;114(10):2375–83.

    CAS  PubMed  PubMed Central  Google Scholar 

  244. Isin M, et al. Investigation of circulating lncRNAs in B-cell neoplasms. Clin Chim Acta. 2014;431:255–9.

    CAS  PubMed  Google Scholar 

  245. Xia T, et al. Long noncoding RNA FER1L4 suppresses cancer cell growth by acting as a competing endogenous RNA and regulating PTEN expression. Sci Rep. 2015;5:13445.

    CAS  PubMed  PubMed Central  Google Scholar 

  246. Chen P, et al. miR-9 is an essential oncogenic microRNA specifically overexpressed in mixed lineage leukemia-rearranged leukemia. Proc Natl Acad Sci U S A. 2013;110(28):11511–6.

    CAS  PubMed  PubMed Central  Google Scholar 

  247. Emmrich S, et al. miR-9 is a tumor suppressor in pediatric AML with t(8;21). Leukemia. 2014;28(5):1022–32.

    CAS  PubMed  Google Scholar 

  248. Fu L, et al. A minicircuitry of microRNA-9-1 and RUNX1-RUNX1T1 contributes to leukemogenesis in t(8;21) acute myeloid leukemia. Int J Cancer. 2017;140(3):653–61.

    CAS  PubMed  Google Scholar 

  249. Bonizzato A, et al. CircRNAs in hematopoiesis and hematological malignancies. Blood Cancer J. 2016;6(10):e483.

    CAS  PubMed  PubMed Central  Google Scholar 

  250. Prendergast M, Schoenrock LD. Malar augmentation: patient classification and placement. Arch Otolaryngol Head Neck Surg. 1989;115(8):964–9.

    CAS  PubMed  Google Scholar 

  251. Guarnerio J, et al. Oncogenic role of fusion-circRNAs derived from Cancer-associated chromosomal translocations. Cell. 2016;166(4):1055–6.

    CAS  PubMed  Google Scholar 

  252. Okcanoğlu TB, Gündüz C. Circular RNAs in leukemia. Biomed Reports. 2019;0(0):1–5.

    Google Scholar 

Download references

Acknowledgements

MH is supported by Sidra Medicine institutional funding. AAB is supported by Sidra Medicine internal grant (SIRF_20046) and SU is supported by Medical Research Centre grants (grant# 16102/6, #16354/16). The authors would like to express their gratitude to Dr. Vineeta Tanwar (Research Scientist, Ohio State University, Ohio, Columbus, USA) for help in English editing and valuable suggestions to improve the quality of the manuscript.

Funding

The authors declare that no funding support was received for this study.

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization, AAB, SYN, MH and SU; writing—original draft preparation, AAB, SN, IA, RM, SKS, LZ, IE, SK, KSP, AQ, SK and SU; writing—review and editing, AAB, SN, MH, MK, WER, HZ, MK and SU; Revision of manuscript, AAB, SN, IA, RM, SKS, LZ, IE, SK, KSP, AQ,SK and SU supervision, AAB, SU, MH, SN, MK, WER and SU. All authors have read and approved the final version of the manuscript.

Corresponding authors

Correspondence to Mohammad Haris or Shahab Uddin.

Ethics declarations

Ethics approval and consent to participate

Not applicable, neither ethics approval was required for this review nor involvement of patients.

Consent for publication

All authors consent to publication.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bhat, A.A., Younes, S.N., Raza, S.S. et al. Role of non-coding RNA networks in leukemia progression, metastasis and drug resistance. Mol Cancer 19, 57 (2020). https://doi.org/10.1186/s12943-020-01175-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12943-020-01175-9

Keywords