Acute myeloid leukemia (AML) is a deadly hematologic malignancy characterized by the uncontrolled growth of immature myeloid cells. Over the past several decades, we have learned a tremendous amount regarding the genetic aberrations that govern disease development in AML. Among these are genes that encode noncoding RNAs, including the microRNA (miRNA) family. miRNAs are evolutionarily conserved small noncoding RNAs that display important physiological effects through their posttranscriptional regulation of messenger RNA targets. Over the past decade, studies have identified miRNAs as playing a role in nearly all aspects of AML disease development, including cellular proliferation, survival, and differentiation. These observations have led to the study of miRNAs as biomarkers of disease, and efforts to therapeutically manipulate miRNAs to improve disease outcome in AML are ongoing. Although much has been learned regarding the importance of miRNAs in AML disease initiation and progression, there are many unanswered questions and emerging facets of miRNA biology that add complexity to their roles in AML. Moving forward, answers to these questions will provide a greater level of understanding of miRNA biology and critical insights into the many translational applications for these small regulatory RNAs in AML.

MicroRNA (miRNAs) are small noncoding RNAs (∼20-24 nucleotides) that play vital roles in posttranscriptional gene regulation through repression of target messenger RNAs (mRNAs).1  miRNA-encoding genes in the nucleus are transcribed into primary miRNA transcripts, which then undergo a number of processing steps in the nucleus and cytoplasm to generate the mature miRNA molecule. The mature miRNA is loaded into the RNA-induced silencing complex (RISC), and this miRNA-RISC complex targets the 3′ untranslated region (UTR) of specific mRNAs on the basis of sequence complementarity, resulting in reduced protein outputs through mechanisms involving decreased mRNA stability and reduced translation.2 

miRNAs are now recognized to play roles in nearly all physiological processes and have been implicated in a number of human diseases including cancer, where miRNAs can act as either oncogenes (oncomiRs) or tumor suppressors.3  Hematologic malignancies are no exception, as dysregulated miRNA expression contributes to blood cancers from many different hematopoietic lineages.4  This review will focus on recent advances in understanding the role of miRNAs in a hematologic malignancy with a particularly high rate of mortality, acute myeloid leukemia (AML).

AML is a heterogeneous disease characterized by the increased proliferation and survival of immature myeloid cells and is the result of a number of genetic abnormalities, including mutations and chromosomal rearrangements.5  Early studies characterizing the role of miRNAs in AML focused on identifying AML-specific miRNA expression patterns. Distinctive miRNA profiles were identified for many cytogenetic subtypes of AML,6-8  as well as for several specific mutations in cytogenetically normal AML, including mutations in NPM1, FLT3, and CEBPA.9-13  miRNA expression profiles also correlate with prognosis,12,14  highlighting the potential importance of miRNAs in this disease. However, although miRNAs are enriched in leukemia-associated genomic alterations, only ∼100 are expressed above background level,15  suggesting that only a subset of miRNAs have functional effects in AML.

Beyond dysregulated miRNA expression profiles, it is now well accepted that miRNAs can function as either oncomiRs or tumor suppressors in many subtypes of AML, affecting a broad range of leukemic processes, including proliferation, survival, differentiation, self-renewal, epigenetic regulation, in vivo disease progression, and chemotherapy resistance8,15-57  (Table 1). miRNAs impact leukemic development and progression through collaboration with known oncogenes or tumor suppressors, either by directly targeting them on the mRNA level or by working in concert with these proteins to promote malignancy. To illustrate these concepts, we have summarized findings for selected miRNAs consistently found to play a role in AML in Table 1, and we discuss some of the more novel aspects of miRNA biology in AML below, as a greater understanding of miRNA biology will enable more strategic design of therapies in the future.

Table 1.

Selected miRNAs involved in AML pathogenesis

miRNAExpression changeMechanism of dysregulationConfirmed targetsFunctional effect of expressionReferences
OncomiRTumor suppressor
miR-9 ↑ MLL-rearranged AML, ↓ t(8;21), ↓ EVI1-induced AML Promoter targeted by MLL-fusion proteins; EVI1 hypermethylates promoter RHOH, RYBP, HMGA2, LIN28B, FOXO1, FOXO3 ↑ Proliferation, ↑ survival, ↑ leukemogenesis in mice (MLL rearranged) ↓ Proliferation, ↑ monocytic differentiation (t(8;21) and EVI1+) 16-18 
miR-17-92 cluster ↑ In LSCs in MLL-associated AML Activated by MYC; epigenetically activated by MLL-fusion proteins; genomic amplification in MLL-rearranged AML P21 ↑ Proliferation, ↑ survival, ↓ differentiation, ↑ self-renewal, ↑ colony-forming capacity, ↑ leukemogenesis in mice  19, 20 
miR-22 ↑ MDS/MDS-derived AML, ↓ de novo AML Downregulated via TET1/GFI1/EZH2/SIN3A-mediated epigenetic repression and DNA copy-number loss; increased with loss of PU.1 TET2, CRTC1, FLT3, MYCBP ↑ Proliferation, ↑ survival, ↓ differentiation, ↑ self-renewal, overexpression leads to myeloid malignancy in mice (MDS/MDS-derived AML) ↓ AML blast cell growth, ↑ differentiation, ↓ leukemic progression in mice (de novo AML) 21-23 
miR-29b ↓ Various subtypes of AML Downregulated via loss of CEBPA; chromosome 7q deletions; MYC represses; NF-kB represses MCL-1, CXXC6, CDK6, AKT2, CCND2, SP1, DNMT3A, DNMT3B  ↓ Cell growth, ↑ apoptosis, ↓ leukemic progression in vivo, ↓ KIT activation, prevents global DNA hypermethylation 24-27, 56 
miR-125b ↑ t(2;11)(p21;q23) AML, ↑ MDS/MDS-derived AML, ↑ pediatric AML Increased by t(2;11)(p21;q23) translocation LIN28A, IRF4 ↑ Proliferation, ↑ production of myeloid progenitors, ↑ self-renewal, overexpression leads to AML in mice  28-32 
miR-126 ↑ t(8;21) AML, ↑ in LSCs of CN-AML Promoter demethylated in t(8;21) AML PLK2, ADAM9, ILK, GOLPH3, CDK3, TOM1, ERRFI1, SPRED1, FZD7 ↑ Proliferation, ↑ survival, ↓ differentiation, ↑ colony-forming ability, ↑ LSC maintenance and self-renewal, ↓ cell cycling, ↑ quiescence, ↑ chemotherapy resistance  8, 33-36 
miR-146a ↓ Various subtypes of AML, ↓ 5q syndrome MDS/MDS-derived AML Deletion in del(5q) MDS/MDS-derived AML TRAF6, IRAK1, TIRAP  ↓ Proliferation, ↓ survival, ↓ NF-kB activation, deletion leads to myeloproliferation in mice 15, 37-41 
miR-155 ↑ CN-AML (highest in FLT3-ITD+ AML) Targeted by STAT5 and NF-kB in FLT3-ITD+ AML; upregulated by MLL-fusion genes via MEIS1 CEBPB, SHIP1, PU.1 ↑ Proliferation, ↑ survival, overexpression leads to myeloproliferative neoplasm in mice, confers negative prognosis in CN-AML, no effect in MLL-AF9 mouse model of leukemia  42-46 
miR-193a ↓ Various subtypes of AML (lowest in t(8;21) AML) Epigenetically silenced by AML1/ETO AML1/ETO, DNMT3A, HDAC3, KIT, CCND1, MDM2  ↓ Cell growth, ↑ apoptosis, ↑ differentiation, ↓ cell cycling, ↓ KIT expression 47, 48 
miR-196b ↑ MLL-associated AML Activated by MLL-fusion proteins; co-expressed with HOXA9 in MLL-rearranged leukemia HOXA9, MEIS1, FAS ↑ Proliferation, ↑ survival, ↓ differentiation, ↓ replating potential, ↑ MLL-AF9–induced leukemic progression in mice  49, 50 
miR-223 ↓ t(8;21) AML, ↓ various subtypes of AML Targeted by transcription factors CEBPA (activates) and E2F1 (represses); epigenetically silenced by AML1/ETO E2F1, MEF2C, FBXW7  ↓ Proliferation, ↑ apoptosis, ↑ differentiation/granulopoiesis 51-54 
miRNAExpression changeMechanism of dysregulationConfirmed targetsFunctional effect of expressionReferences
OncomiRTumor suppressor
miR-9 ↑ MLL-rearranged AML, ↓ t(8;21), ↓ EVI1-induced AML Promoter targeted by MLL-fusion proteins; EVI1 hypermethylates promoter RHOH, RYBP, HMGA2, LIN28B, FOXO1, FOXO3 ↑ Proliferation, ↑ survival, ↑ leukemogenesis in mice (MLL rearranged) ↓ Proliferation, ↑ monocytic differentiation (t(8;21) and EVI1+) 16-18 
miR-17-92 cluster ↑ In LSCs in MLL-associated AML Activated by MYC; epigenetically activated by MLL-fusion proteins; genomic amplification in MLL-rearranged AML P21 ↑ Proliferation, ↑ survival, ↓ differentiation, ↑ self-renewal, ↑ colony-forming capacity, ↑ leukemogenesis in mice  19, 20 
miR-22 ↑ MDS/MDS-derived AML, ↓ de novo AML Downregulated via TET1/GFI1/EZH2/SIN3A-mediated epigenetic repression and DNA copy-number loss; increased with loss of PU.1 TET2, CRTC1, FLT3, MYCBP ↑ Proliferation, ↑ survival, ↓ differentiation, ↑ self-renewal, overexpression leads to myeloid malignancy in mice (MDS/MDS-derived AML) ↓ AML blast cell growth, ↑ differentiation, ↓ leukemic progression in mice (de novo AML) 21-23 
miR-29b ↓ Various subtypes of AML Downregulated via loss of CEBPA; chromosome 7q deletions; MYC represses; NF-kB represses MCL-1, CXXC6, CDK6, AKT2, CCND2, SP1, DNMT3A, DNMT3B  ↓ Cell growth, ↑ apoptosis, ↓ leukemic progression in vivo, ↓ KIT activation, prevents global DNA hypermethylation 24-27, 56 
miR-125b ↑ t(2;11)(p21;q23) AML, ↑ MDS/MDS-derived AML, ↑ pediatric AML Increased by t(2;11)(p21;q23) translocation LIN28A, IRF4 ↑ Proliferation, ↑ production of myeloid progenitors, ↑ self-renewal, overexpression leads to AML in mice  28-32 
miR-126 ↑ t(8;21) AML, ↑ in LSCs of CN-AML Promoter demethylated in t(8;21) AML PLK2, ADAM9, ILK, GOLPH3, CDK3, TOM1, ERRFI1, SPRED1, FZD7 ↑ Proliferation, ↑ survival, ↓ differentiation, ↑ colony-forming ability, ↑ LSC maintenance and self-renewal, ↓ cell cycling, ↑ quiescence, ↑ chemotherapy resistance  8, 33-36 
miR-146a ↓ Various subtypes of AML, ↓ 5q syndrome MDS/MDS-derived AML Deletion in del(5q) MDS/MDS-derived AML TRAF6, IRAK1, TIRAP  ↓ Proliferation, ↓ survival, ↓ NF-kB activation, deletion leads to myeloproliferation in mice 15, 37-41 
miR-155 ↑ CN-AML (highest in FLT3-ITD+ AML) Targeted by STAT5 and NF-kB in FLT3-ITD+ AML; upregulated by MLL-fusion genes via MEIS1 CEBPB, SHIP1, PU.1 ↑ Proliferation, ↑ survival, overexpression leads to myeloproliferative neoplasm in mice, confers negative prognosis in CN-AML, no effect in MLL-AF9 mouse model of leukemia  42-46 
miR-193a ↓ Various subtypes of AML (lowest in t(8;21) AML) Epigenetically silenced by AML1/ETO AML1/ETO, DNMT3A, HDAC3, KIT, CCND1, MDM2  ↓ Cell growth, ↑ apoptosis, ↑ differentiation, ↓ cell cycling, ↓ KIT expression 47, 48 
miR-196b ↑ MLL-associated AML Activated by MLL-fusion proteins; co-expressed with HOXA9 in MLL-rearranged leukemia HOXA9, MEIS1, FAS ↑ Proliferation, ↑ survival, ↓ differentiation, ↓ replating potential, ↑ MLL-AF9–induced leukemic progression in mice  49, 50 
miR-223 ↓ t(8;21) AML, ↓ various subtypes of AML Targeted by transcription factors CEBPA (activates) and E2F1 (represses); epigenetically silenced by AML1/ETO E2F1, MEF2C, FBXW7  ↓ Proliferation, ↑ apoptosis, ↑ differentiation/granulopoiesis 51-54 

Table represents selected miRNAs involved in AML disease progression, with a focus on miRNAs with clinical evidence of dysregulation, as well as in vivo and in vitro functional evidence of role in leukemogenesis.

CN, cytogenetically normal; LSC, leukemic stem cell.

Alterations in miRNA expression can occur through a variety of mechanisms in AML (Figure 1). Copy number alterations (CNAs), which include deletions22,25  and amplifications,20  can drastically alter miRNA expression. However, acquired CNAs specifically targeting miRNAs may be relatively rare in AML. By using a combination of comparative genomic hybridization and whole-genome sequencing, researchers found that 18% of patients had CNAs involving miRNA genes, with a single CNA affecting up to 121 miRNAs.58  However, these CNAs always contained one or more protein-coding genes, suggesting that the miRNA genes involved in these CNAs may be passenger alterations. miRNAs may also be aberrantly expressed when located in oncogenic genomic locations, which occurs through chromosomal translocations28  or overexpression of nearby protein-coding genes.49 

Figure 1.

Mechanisms of dysregulated miRNA expression in AML. miRNA dysregulation can contribute to the development of AML. Thus far, numerous mechanisms by which miRNAs become dysregulated in AML have been identified, including (1) deletions leading to decreased miRNA expression, (2) improper expression because of close proximity to an oncogenic genomic region created as a result of either a translocation event or overexpression of a neighboring protein-coding gene, (3) copy number amplifications leading to increased miRNA expression, (4) epigenetic alterations affecting miRNA expression, (5) miRNA promoter regions being aberrantly targeted by dysregulated transcription factors or oncoproteins, and (6) dysregulated miRNA processing leading to altered levels of mature miRNAs.

Figure 1.

Mechanisms of dysregulated miRNA expression in AML. miRNA dysregulation can contribute to the development of AML. Thus far, numerous mechanisms by which miRNAs become dysregulated in AML have been identified, including (1) deletions leading to decreased miRNA expression, (2) improper expression because of close proximity to an oncogenic genomic region created as a result of either a translocation event or overexpression of a neighboring protein-coding gene, (3) copy number amplifications leading to increased miRNA expression, (4) epigenetic alterations affecting miRNA expression, (5) miRNA promoter regions being aberrantly targeted by dysregulated transcription factors or oncoproteins, and (6) dysregulated miRNA processing leading to altered levels of mature miRNAs.

Close modal

The most common mechanisms by which miRNA expression becomes dysregulated in AML are epigenetic alterations and via targeting by dysregulated transcription factors or oncogenic fusion proteins. These two mechanisms are not always distinct, as epigenetic alterations to miRNA loci often occur via dysregulated transcription factors or oncoproteins.18,47  There is some evidence that alterations in miRNA expression in cancer can be the result of dysregulated miRNA processing59,60 ; however, it is unclear whether this occurs in AML.

Although mutations in the mature miRNA sequence would likely have no effect on expression levels, these mutations could change mRNA target specificity and dramatically alter phenotypic effects in AML. In perhaps the most thorough AML sequencing effort to date, The Cancer Genome Atlas group reported that miR-142-3p was the only miRNA bearing recurrent somatic single nucleotide variants in its mature strand that could alter binding to targets (4 of 187).13,61  Only 7 miRNA single nucleotide variant mutations were discovered in the 187 samples analyzed, indicating that these are rare events. However, while mutations in the miRNA sequence itself are uncommon, polymorphisms in the mRNA 3′ UTR miRNA binding site may happen more frequently and could predispose patients to AML by altering miRNA regulation of specific genes.62  Taken together, it seems that aberrant miRNA levels that are observed in AML are largely driven by altered transcription of miRNA primary transcripts, which suggests that targeting of key transcription factors or epigenetic regulators may be one way to restore proper miRNA expression in AML.

It is now well established that miRNAs play a variety of critical roles in AML, in which they can either promote or inhibit tumor cell biology. However, these advances have yet to make a clinical impact. Here we will highlight the efforts being made toward moving miRNA research in AML to the clinic and focus on the potential for using miRNAs as disease biomarkers, as well as advances in miRNA-targeting therapeutic strategies in AML.

miRNAs as AML biomarkers

Perhaps the most encouraging clinical application of miRNA research to date is the potential use of miRNAs as disease biomarkers in AML.63  When a patient initially presents with leukemia, proper classification is critical to determining the correct treatment. However, a small number of leukemias are difficult to identify as myeloid or lymphoid, thus making treatment decisions challenging. miRNA expression profiling can help classify acute leukemias of ambiguous lineage as either AML or acute lymphoblastic leukemia,64  with 1 group claiming that as few as 2 miRNAs can be used to discriminate between acute lymphoblastic leukemia and AML at an accuracy of >95%.65  As mentioned earlier, specific subtypes and mutant drivers of AML are associated with distinctive miRNA expression profiles, again suggesting that miRNAs could be useful in the initial classification of disease.

Beyond classification, miRNA expression profiles may provide important prognostic information. Several groups have reported that miRNA expression at diagnosis adds relevant prognostic information in patients with AML and can even predict survival in some cases.66-68  It was also recently reported that miRNA expression can predict progression of myelodysplastic syndrome (MDS) to AML.69 

A major issue with patients receiving treatment for AML is the persistence of a small number of leukemic blasts in the bone marrow after intensive chemotherapy known as minimal residual disease (MRD), which can eventually give rise to leukemia relapse. Because the appearance of leukemic blasts in circulation often occurs late in the relapse process, a number of highly sensitive polymerase chain reaction– and flow cytometry–based methods for detection of blast nucleic acid or protein products have been developed for monitoring MRD,70  as recognizing MRD before patient relapse could allow for preemptive therapy.71  A number of groups have proposed screening circulating miRNAs as an inexpensive, noninvasive, and sensitive option to monitor for MRD, because the serum expression of miRNAs changes after standard chemotherapy,72  and patients with AML have a distinctive serum miRNA expression profile compared with healthy controls.73,74  These early results are promising, because a specific AML-associated miRNA serum profile could not only be used to track MRD after chemotherapy, but could also potentially provide an important screening tool for early detection of de novo AML in the clinic, as alterations in serum miRNA profiles may precede the entry of leukemic blasts into the periphery. However, there has been a lack of concordance between these individual studies, suggesting that more work is needed on larger AML cohorts with more rigorous study design to validate these initial findings.

Advances in miRNA-based therapeutics

As the list of miRNAs and their mRNA targets that are relevant in AML disease progression continues to grow, therapeutic manipulation of these miRNAs becomes more enticing. It is easy to imagine delivering locked nucleic acid (LNA) oligonucleotide inhibitors to target known oncomiRs in AML or delivery of synthetic miRNA mimics that act as tumor suppressors. These approaches have exciting therapeutic potential, because miRNAs are endogenous molecules that often repress multiple targets, either in the same pathway or by affecting a common biological process. Thus, resistance to miRNA-based therapies through target site mutation would be unlikely.

A good example of the effectiveness of an miRNA-based therapeutic in AML was recently demonstrated with targeted delivery of miR-29b via transferrin-conjugated lipid nanoparticles both in vitro and in mice engrafted with human AML cell lines.75  Delivery of miR-29b led to decreased leukemic cell growth and improved survival in the AML xenograft mouse model, which was attributed to miR-29b downregulating CDK6, SP1, FLT3, DNMTs, and KIT, either directly or indirectly. These target genes affect a variety of cellular processes in AML, and this study highlights the ability of 1 miRNA-based treatment to target many pathways simultaneously. Several studies involving the use of miR-based therapeutics have shown encouraging results in preclinical in vitro and animal models,22,33,75-77  the results of which are summarized in Table 2.

Table 2.

Preclinical studies of miRNA-based therapeutics

TherapyDelivery methodTargetsEfficacy of deliveryIn vitro resultsIn vivo resultsReferences
miR-22 mimic G7 poly(amidoamine) dendrimer nanoparticles CRTC1, FLT3, MYCBP N/A N/A ↑ Survival in MV4-11 xenotransplantation mouse models;
↑ Survival (40% cure rate) in primary mouse AML (MLL-AF9 and AE9a) transplant models 
22 
miR-29b mimic Transferrin-conjugated anionic lipid-based nanoparticle DNMT3A, DNMT3B, DNMT1, SP1, CDK6, FLT3, KIT >100-fold ↑ in vitro ↓ Cell proliferation and ↓ colony formation in Kasumi-1, OCI-AML3, and MV4-11 cells ↑ Survival and ↓ splenomegaly in NSG mice xenografted with MV4-11 cells 69 
   20-fold ↑ in vivo ↓ Cell viability in primary AML samples; pretreatment ↑ efficacy of decitabine treatment Pretreatment ↑ efficacy of decitabine treatment 
miR-21/miR-196b antagomiRs Naked antagomiR delivered via implanted osmotic pumps N/A 60%-80% ↓ of miRNAs in peripheral WBCs in vivo ↓ Colony formation in HOXA9, Nup98-HOXA9, or MLL-AF9–transduced mouse Lin BM cells Cured MLL-AF9 transplanted primary mouse AML; ↑ survival in combination with induction chemotherapy in MLL-AF9 xenotransplantation model 71 
miR-126 antagomiR Transferrin or CD45.2- conjugated anionic lipid-based nanoparticle MMP7, CHD7, JAG1 80% ↓ in primary AML blasts in vitro; 50% ↓ in BM and spleen in vivo ↓ Long-term colony formation, ↓ self-renewal capacity, and ↓ quiescence of LSCs ↑ Survival in NSG mice engrafted with human AML primary blasts and MLL FLT3-ITD mouse model 33 
miR-181a mimic Transferrin-conjugated anionic lipid-based nanoparticle KRAS, NRAS, MAPK1 2.6-fold ↑ in BM and 35-fold ↑ in spleen in vivo ↓ Cell proliferation and ↓ colony forming in KG1a, MV4-11, OCI-AML3 cell lines; ↑ apoptosis in primary AML blasts ↑ Survival and ↓ splenomegaly in NSG mice xenografted with MV4-11 cells 70 
TherapyDelivery methodTargetsEfficacy of deliveryIn vitro resultsIn vivo resultsReferences
miR-22 mimic G7 poly(amidoamine) dendrimer nanoparticles CRTC1, FLT3, MYCBP N/A N/A ↑ Survival in MV4-11 xenotransplantation mouse models;
↑ Survival (40% cure rate) in primary mouse AML (MLL-AF9 and AE9a) transplant models 
22 
miR-29b mimic Transferrin-conjugated anionic lipid-based nanoparticle DNMT3A, DNMT3B, DNMT1, SP1, CDK6, FLT3, KIT >100-fold ↑ in vitro ↓ Cell proliferation and ↓ colony formation in Kasumi-1, OCI-AML3, and MV4-11 cells ↑ Survival and ↓ splenomegaly in NSG mice xenografted with MV4-11 cells 69 
   20-fold ↑ in vivo ↓ Cell viability in primary AML samples; pretreatment ↑ efficacy of decitabine treatment Pretreatment ↑ efficacy of decitabine treatment 
miR-21/miR-196b antagomiRs Naked antagomiR delivered via implanted osmotic pumps N/A 60%-80% ↓ of miRNAs in peripheral WBCs in vivo ↓ Colony formation in HOXA9, Nup98-HOXA9, or MLL-AF9–transduced mouse Lin BM cells Cured MLL-AF9 transplanted primary mouse AML; ↑ survival in combination with induction chemotherapy in MLL-AF9 xenotransplantation model 71 
miR-126 antagomiR Transferrin or CD45.2- conjugated anionic lipid-based nanoparticle MMP7, CHD7, JAG1 80% ↓ in primary AML blasts in vitro; 50% ↓ in BM and spleen in vivo ↓ Long-term colony formation, ↓ self-renewal capacity, and ↓ quiescence of LSCs ↑ Survival in NSG mice engrafted with human AML primary blasts and MLL FLT3-ITD mouse model 33 
miR-181a mimic Transferrin-conjugated anionic lipid-based nanoparticle KRAS, NRAS, MAPK1 2.6-fold ↑ in BM and 35-fold ↑ in spleen in vivo ↓ Cell proliferation and ↓ colony forming in KG1a, MV4-11, OCI-AML3 cell lines; ↑ apoptosis in primary AML blasts ↑ Survival and ↓ splenomegaly in NSG mice xenografted with MV4-11 cells 70 

This table summarizes the results from selected pre-clinical miRNA-based therapies, including information such as method of delivery, miRNA targets (direct and indirect), efficacy of delivery, and in vivo/in vitro results.

BM, bone marrow; N/A, not applicable; WBC, white blood cell.

Another underexplored area of miRNA-based therapy is the possibility of repurposing existing drugs known to influence miRNA levels by targeting the pathways that regulate miRNA expression. MLN4924 (Pevonedistat), a drug known to reduce nuclear factor κB (NF-κB) activation that is currently being evaluated in clinical trials, was recently shown to decrease the levels of oncogenic miR-155 in FLT3-ITD+ AML cell lines, leading to decreased leukemic phenotypes both in vitro and in vivo.78  miRNA-based therapeutics may also be efficacious when used in combination with existing chemotherapeutics. Manipulation of miRNA expression levels can increase AML responsiveness to standard chemotherapeutic regimens.34,75,79-81 

Although several studies have implicated miRNAs and their putative targets as being clinically actionable, the vast majority of these studies have yet to achieve clinical relevance, with the first therapies targeting miRNAs just entering clinical trials within the last few years.82,83  One miRNA-based therapy in clinical trials that shows promise is treatment of hepatitis C virus by miR-122, in which researchers found that patients treated with an LNA inhibitor of mature miR-122 have reductions in hepatitis C virus RNA levels in a dose-dependent manner.84  This study provides proof of principle that should encourage future endeavors of this kind in the cancer arena. Although the list of miRNA-based therapeutics entering clinical trials continues to grow, to the best of our knowledge, no miRNA-based therapies have made their way to clinical trials specifically for the treatment of AML.

A major barrier preventing the development of miRNA-based therapies is the lack of more efficient and specific delivery methods, because synthetic miRNAs or oligonucleotide inhibitors are degraded rapidly in circulation and have limited cellular uptake and specificity. To further complicate matters, delivery of drugs to the bone marrow is difficult, and higher doses are often required to elicit a therapeutic effect. This highlights the importance of developing novel targeting techniques for more effective delivery. Consequently, there are many new approaches being explored for improved delivery of miRNA-based therapies, including liposomes, nanoparticles, LNAs with increased stability, peptide-based inhibitors, and several other creative approaches,85  some of which are highlighted in Table 2. Efficient and specific delivery of miRNA mimics or antagonists to the proper cell types in vivo is a key step toward unlocking the therapeutic potential of manipulating miRNA function to combat AML.

Going forward, there remain several aspects of miRNA biology that need further investigation to fully grasp how miRNAs function within AML cells. This includes a continued effort to better answer certain questions that have been raised in the past and work in novel areas that have recently emerged. Several of these areas and how they relate to AML are described below.

Improper regulation of inflammatory pathways leads to AML

A newly appreciated mechanism by which miRNAs promote malignancy is through their impacts on classical inflammatory pathways. It has long been known that there is a strong link between inflammation and cancer; however, the mechanisms governing this association are still largely unclear. NF-κB is critical in initiating inflammatory responses, and dysregulation of this critical transcription factor is heavily integrated into cancer biology because of its role in promoting proliferation and survival.86 

Several groups have shown that dysregulation of certain miRNAs can disrupt normal NF-κB signaling that results in cancerous transformation,87,88  including in myeloid malignancies.39  Because miR-146a is a negative regulator of NF-κB signaling, and overactivation of NF-κB is involved in malignant transformation, one could predict that loss of miR-146a might lead to the development of hematopoietic cancers. Indeed, a miR-146a deficiency has been shown to result in the development of both lymphoid and myeloid malignancies in an age-dependent manner.39,41  Chromosome 5q deletions, which are common in MDS progressing to AML, leads to loss of miR-145 and miR-146a because they are both encoded on the long arm of chromosome 5.37  The loss of these miRNAs leads to myeloproliferation and eventual progression to AML in mice as a result of increased NF-κB signaling,37,40  as miR-145 and miR-146a target TIRAP, IRAK1, and TRAF6, known activators of NF-κB. Targeted inhibition of IRAK1 has significant activity against MDS/AML cells in vitro and in xenograft mouse models,89  suggesting that targeting of these traditional innate immune pathways may have clinical efficacy.

Not only does miRNA regulation of NF-κB signaling seem to be important in AML progression, but there is also evidence that NF-κB activates miRNA expression to promote leukemic phenotypes.90  Additionally, there could be some contribution from the bone marrow microenvironment. Activation of inflammatory signaling in mesenchymal cells was recently found to drive development of an MDS preleukemic condition in mice.91  A further understanding of how alterations to these miRNA-regulated classical inflammatory pathways can promote AML progression will be an interesting new area of miRNA research in the future.

miRNAs play context-dependent roles in AML

An interesting aspect of miRNA biology in AML is that a miRNA can have opposing roles, depending on the disease context. For example, miR-9 was identified as being specifically upregulated in MLL-rearranged AML, in which it plays an oncogenic role in promoting leukemogenesis in the presence of MLL-AF9.16  However, other studies have found miR-9 to play a tumor suppressive role in AML, including in pediatric AML with t(8;21), in which miR-9 overexpression reduced leukemic growth and induced monocytic differentiation in human AML cells and in xenotransplantation mouse models,17  as well as in EVI1-induced AML, in which miR-9 is epigenetically silenced leading to decreased apoptosis and myelopoiesis.18 

There are many other examples of miRNAs displaying context-dependent roles in AML. miR-155 seems to have no phenotypic effect in MLL-rearranged AML,44  but it is consistently found to play an oncogenic role in FLT3-ITD–driven AML pathogenesis.42,45  miR-126 plays different roles and regulates different targets in normal vs malignant hematopoietic stem cells35  and, interestingly, both overexpression and knockout of miR-126 promote leukemogenesis.34  These studies highlight how the influence of an miRNA in AML can be dependent on the underlying genetic abnormalities that drive disease or cell type of expression.

A variety of potential explanations for context-dependent discrepancies have been proposed, including RNA-binding protein regulation of miRNA binding to 3′ UTRs or differential splicing to include or exclude a given 3′ UTR.92  In addition, recent findings suggest that mutant proteins in AML can alter miRNA-mRNA interactions.93  Perhaps the most plausible explanation would be differences in the mRNA target availability for the miRNA, because AML driven by independent mutations would have distinct transcriptional profiles (Figure 2). However, most studies that find context-specific roles for miRNAs in AML do not go on to explore the mechanistic basis underlying this phenomenon, and more work in this area is needed to better understand context-dependent miRNA functions.

Figure 2.

miRNAs play context-dependent roles in AML. A model for context-dependent effects of a specific miRNA given different transcriptional backgrounds between 2 distinct AML-driver mutations, mutation A and mutation B. Mutation A leads to the transcription of mRNA A, B, and C, whereas mutation B drives the transcription of mRNA X, Y, and Z. All mRNAs have predicted targeting by the example miRNA. The miRNA depicted in mutation A and mutation B is the same hypothetical miRNA.

Figure 2.

miRNAs play context-dependent roles in AML. A model for context-dependent effects of a specific miRNA given different transcriptional backgrounds between 2 distinct AML-driver mutations, mutation A and mutation B. Mutation A leads to the transcription of mRNA A, B, and C, whereas mutation B drives the transcription of mRNA X, Y, and Z. All mRNAs have predicted targeting by the example miRNA. The miRNA depicted in mutation A and mutation B is the same hypothetical miRNA.

Close modal

Varying levels of miRNA expression may have opposing effects

One potential explanation for the context-dependent effects observed for miRNA dysregulation in AML is that different levels of miRNA expression may have vastly different effects on host-cell phenotypes. miR-125b is one of the more interesting known oncomiRs because it plays a role in promoting both myeloid and lymphoid malignancies. Overexpression in mice has been shown to cause both lymphoproliferative and myeloproliferative disorders and, ultimately, a frank malignancy in these compartments.29,30,94,95  Recently, some light was shed on the dual nature of miR-125b in promoting hematologic malignancy in which miR-125b was found to selectively induce either myeloid or lymphoid leukemia based on the level and time course of miR-125b overexpression.31 

miR-155 is another well-studied oncogenic miRNA in AML, and overexpression correlates with a poor prognosis.43  However, evidence has emerged that miR-155 may play a role as a tumor suppressor in certain contexts.96  A recent study examined the role of miR-155 in AML more closely to help resolve these opposing effects. The study found that when overexpressing miR-155 in 3 different murine models of AML (HoxA9/Meis1, MLL-ENL, MLL-AF9) to an intermediate level (∼5- to 10-fold above control), miR-155 displayed oncogenic function, leading to increased proliferation and enhanced colony-forming potential.97  This was in contrast to miR-155 high levels (>10-fold above control), in which miR-155 acted as a tumor suppressor by repressing colony formation and proliferation, establishing a dose-dependent effect of miR-155 in these AML mouse models. The study did confirm that the intermediate miR-155 expression levels were a better representation of what was seen in their pediatric AML data set (increased ∼one- to sevenfold), values that were consistent with miR-155 expression in other AML data sets,12,45  suggesting that miR-155 likely has a predominately oncogenic effect in human AML. The study highlights the importance of considering the level of expression in various model systems when studying miRNAs in AML.

5p vs 3p transcripts

A long-standing question regarding miRNAs is the significance of differential use of 5p vs 3p miRNA transcripts. 5p and 3p miRNAs are encoded by the same genomic region and are both contained within the initial transcript and mature miRNA duplex before 1 of them is chosen as the active or guide strand (miR) and the other as the passenger strand (miR*). The passenger strand is then typically degraded and traditionally thought not to have a functional role. Interestingly, the 2 miRNA strands each have unique seed sequences and therefore do not share the same mRNA target spectrum. This means that the biological processes and pathways being regulated by any given pri-miRNA transcript could be vastly different depending on the selection of either the 5p or 3p strand as the guide. There are several examples of 5p and 3p transcripts from the same duplex having distinct biological functions.98,99 

In a small percentage of cases, the passenger strand can be stabilized at the same level as the active strand, and may even exhibit important physiological function in myeloid cells.100  A recent report identified a specific passenger strand, miR-9*, that is not detectable in normal myeloid cells but is expressed in 59% of AML cases, and expression levels correlated with prognosis.101  This group also found that miR-9* expression levels had prognostic value, because patients with high miR-9* expression vs low expression were associated with positive patient outcome. Although the mechanisms behind passenger strand–retained expression are still largely unknown, there is some evidence that posttranscriptional modifications to the RNA duplex and differential expression of various RISC components could play an important role in strand selection.102,103  Whether these processes are dysregulated in AML remains to be determined.

A variety of miRNA sources and noncanonical targeting

Although the traditional dogma of miRNA biology states that miRNAs are encoded from their own genes, go through a distinct processing pathway, and then repress mRNA targets via binding the 3′ UTR, there is mounting evidence that this canonical biogenesis pathway might not be exclusive. It is now understood that the mature miRNA can come from a variety of sources, including from the 5p or 3p transcript, long noncoding RNAs (lncRNAs),104  small nucleolar RNAs,105,106  or spliced from introns107  (Figure 3). Identifying and characterizing miRNAs generated from these nontraditional sources may be challenging, but it is key to comprehensively understanding the breadth of small RNAs in AML.

Figure 3.

Alternate miRNA sources and noncanonical targeting. A schematic depicting the varying sources identified for production of mature miRNAs, including the 5′ or 3′ strand of the traditional miRNA hairpin structure, processing of other noncoding RNAs, such as lncRNAs, small nucleolar RNAs (snoRNAs), miRNAs transcribed from the same gene but having different mature miRNA sequences (isomiRs), and miRNA spliced from introns (miRtrons). Nontraditional miRNA targets beyond the 3′ UTR are also depicted, including promoter regions of protein-coding genes, the 5′ UTR, the RNA coding sequence, and protein.

Figure 3.

Alternate miRNA sources and noncanonical targeting. A schematic depicting the varying sources identified for production of mature miRNAs, including the 5′ or 3′ strand of the traditional miRNA hairpin structure, processing of other noncoding RNAs, such as lncRNAs, small nucleolar RNAs (snoRNAs), miRNAs transcribed from the same gene but having different mature miRNA sequences (isomiRs), and miRNA spliced from introns (miRtrons). Nontraditional miRNA targets beyond the 3′ UTR are also depicted, including promoter regions of protein-coding genes, the 5′ UTR, the RNA coding sequence, and protein.

Close modal

Functional effects exhibited by miRNAs are often attributed to a handful of targets predicted by seed sequence complementarity in the 3′ UTR. But miRNAs can also bind and repress targets without predicted binding sites in their 3′ UTRs, referred to as noncanonical targets. Researchers found that when they pulled down the RISC complex to identify mRNA targets loaded in an miR-155–specific manner, ∼40% of the targets identified were noncanonical targets.108  This was explained, in part, by laxity in the seed matching of miRNAs and mRNA 3′UTRs but could also be explained by the concept of isomiRs, which are miRNAs transcribed from the same gene but having different mature miRNA sequences, found extensively in a murine model of leukemia.109  These variants are a result of posttranscriptional modifications, including “errors” in miRNA processing, nucleotide addition to the 3′ end, and nucleotide substitution.110,111  Beyond isomiRs, there is also some evidence that miRNAs can bind to promoter regions of DNA,112  5′ UTRs,113  the mRNA coding sequence,114,115  and even proteins.116  Thus far, the evidence for noncanonical targeting playing a functional role in AML is lacking, but it could be a more prominent mode of miRNA function than we realize.

Transfer of miRNAs in exosomes alters leukemic phenotypes

Recently, miRNAs have been found within extracellular vesicles, including exosomes that are produced by the multivesicular body pathway.117  Both primary and malignant cells can release miRNAs in exosomes, which can be taken up by certain recipient cells where they deliver their miRNA cargo in a functionally relevant manner.118,119  Although evidence for the functional role of exosomally transferred miRNAs in AML is somewhat limited, preliminary studies indicate this could be a paradigm-shifting field of study.

Early work showed that both primary AML cells and AML cell lines do in fact release exosomes containing miRNAs.120  Moreover, these exosomes contain an miRNA population that is compositionally distinct from the miRNA population of the host cell,120  suggesting that there is specificity with the loading of miRNAs into exosomes. Other functional studies have revealed that miRNAs can be transferred in exosomes from AML cells to both stromal and normal hematopoietic cells and alter their function in a manner that promotes leukemic phenotypes.121,122  Interestingly, exosomes from extramedullary tumors have been shown to alter the bone marrow niche, suggesting that miRNA-containing exosomes can home to the bone marrow and alter function independent of cell contact.121  In a recent study,122  authors found that exosomes containing miR-150 and miR-155 released from AML cells suppressed normal hematopoietic stem cell proliferation and differentiation through inhibition of c-MYB, thus perpetuating a malignant phenotype by directly altering hematopoietic stem cell biology.

Such studies provide evidence that miRNAs secreted in exosomes are a novel form of intercellular communication that may play vital regulatory roles in suppressing normal hematopoiesis and disrupting the hematopoietic niche to promote leukemic cell outgrowth (Figure 4). Analysis of the miRNA content of exosomes has even been suggested as a novel biomarker for the detection of AML, because blast-derived exosomes can be isolated from circulation before the appearance of circulating blast cells in a xenograft mouse model.74  Further study of the machinery required for specific miRNA loading into exosomes and uptake by recipient cells is critical to manipulating this system in a manner that will better test the relevance of exosomal miRNAs in AML.

Figure 4.

Exosomally transferred miRNAs alter leukemic phenotypes. A model depicting the different possibilities of exosomal miRNA transfer in the bone marrow, including (1) transfer from AML blast to normal hematopoietic cells (NHCs), (2) NHC to AML blast, (3) AML blast to AML blast, (4) bone marrow (BM) stromal cells to AML blast, (5) AML blast to BM stromal cells, (6) entrance of extramedullary produced exosomes into the hematopoietic niche, and (7) AML blast-derived exosomes leaving the BM and entering circulation.

Figure 4.

Exosomally transferred miRNAs alter leukemic phenotypes. A model depicting the different possibilities of exosomal miRNA transfer in the bone marrow, including (1) transfer from AML blast to normal hematopoietic cells (NHCs), (2) NHC to AML blast, (3) AML blast to AML blast, (4) bone marrow (BM) stromal cells to AML blast, (5) AML blast to BM stromal cells, (6) entrance of extramedullary produced exosomes into the hematopoietic niche, and (7) AML blast-derived exosomes leaving the BM and entering circulation.

Close modal

lncRNAs can interfere with miRNA function in AML

lncRNAs are a distinct class of noncoding RNAs much longer than mature miRNAs (>200 nucleotides) that have been observed to exhibit a variety of functions, but are typically involved in regulating gene expression,123  including miRNA genes.124  It has recently been learned that lncRNA dysregulation in AML can alter the function of specific miRNAs leading to skewed disease phenotypes (Figure 5).

Figure 5.

lncRNAs can interfere with miRNA function in AML. A picture depicting the ways lncRNAs can affect miRNA biology in AML, including (1) lncRNAs serving as a source for mature miRNA production, (2) lncRNAs acting as miRNA sponges, binding miRNAs to prevent them from repressing their target mRNAs, and (3) altering miRNA gene transcription.

Figure 5.

lncRNAs can interfere with miRNA function in AML. A picture depicting the ways lncRNAs can affect miRNA biology in AML, including (1) lncRNAs serving as a source for mature miRNA production, (2) lncRNAs acting as miRNA sponges, binding miRNAs to prevent them from repressing their target mRNAs, and (3) altering miRNA gene transcription.

Close modal

HOTAIRM1 is a lncRNA located in the HOXA genomic region that was recently found to impact prognosis in various subtypes of AML and was associated with a distinct miRNA expression profile.125  Additional work on lncRNA HOTAIRM1 revealed that it was acting to sequester autophagy-regulating miRNAs miR-20a, miR-106b, and miR-125b, which affected the degradation of PML-RARA in acute promyelocytic leukemia.126  Another lncRNA, HOTAIR1, was also found to act as a miRNA sponge in AML, competitively binding miR-193a, which leads to increased c-KIT expression and ultimately confers a poor prognosis.127  Other examples of lncRNAs acting to competitively bind miRNAs in myeloid malignancy have started to emerge.128,129 

As mentioned above, lncRNAs can provide a nontraditional source for mature miRNA production. It was recently learned that the pri-miR-223 transcript is actually a functional lncRNA in AML that displays tumor suppressive functions by sponging oncogenic miRNAs miR-125a and miR-125b.104  Interestingly, while miR-223 is processed out of this lncRNA, miR-223 and lncRNA-223 are expressed at different levels, and these 2 noncoding RNAs have distinct functions in the myeloid lineage. It is unclear how prevalent this phenomenon is in AML, but there is additional evidence that lncRNAs can serve as precursors for miRNAs in T-cell lymphomas.130,131  A better understanding of the ways in which lncRNAs regulate miRNAs in AML could shed more light on their biology in this setting.

miRNAs are now widely regarded as playing a critical role in AML pathogenesis. Specific miRNA expression profiles can help classify subtype, determine prognosis, and predict response to treatment in AML, but the use of miRNAs as biomarkers is not yet routine practice. Therapies targeting miRNAs in AML have shown promise in preclinical models but have not made the leap to human clinical trials, which will require improvements in our delivery methods.

The continued development of advanced genomic approaches, including CRISPR-Cas9 technology, will allow us to more quickly identify and efficiently study relevant miRNAs and their targets in AML. Indeed, genome-wide CRISPR-Cas9 screening has been used to identify functionally relevant miRNA-mRNA target pairs that regulate AML cell line growth132  and will likely be extended to additional preclinical models of AML.

Many complexities and mysteries of miRNA biology remain, but their solutions will substantially improve our understanding of how miRNAs function in AML. In some cases, novel aspects of miRNA biology have already shed additional light on how AML cells are regulated, whereas in other cases, emerging mechanisms have yet to be explored in AML despite their potential to help us understand how miRNAs influence this deadly leukemia. Addressing long-standing questions and exploring emerging concepts of miRNA biology will provide important insights into how miRNAs function in AML, and only through an improved understanding of these mechanisms can we better exploit miRNAs therapeutically to improve disease outcomes in the clinic.

The authors acknowledge Rachel Merrill for assistance with figure preparation.

This work was supported by the National Institutes of Health (NIH) under the Ruth L. Kirschstein National Research Service Award NIH 5T32DK007115 from the National Institute of Diabetes and Digestive and Kidney Diseases (J.A.W.); the NIH National Cancer Institute under Award Number F30CA217027 (J.A.W.); the NIH New Innovator Award DP2GM111099-01 from the National Institute of General Medical Sciences (R.M.O.); Gabrielle's Angel Foundation Award (R.M.O.); and the Leukemia & Lymphoma Society Scholar Award (R.M.O.).

Contribution: J.A.W. and R.M.O. reviewed the literature and wrote the review.

Conflict-of-interest disclosure: The authors declare no competing financial interests.

Correspondence: Ryan M. O’Connell, Department of Pathology, University of Utah, 15 North Medical Dr East, JMRB 4280B, Salt Lake City, UT 84112; e-mail: ryan.oconnell@path.utah.edu.

1.
Bartel
DP
.
MicroRNAs: genomics, biogenesis, mechanism, and function
.
Cell
.
2004
;
116
(
2
):
281
-
297
.
2.
Bartel
DP
.
MicroRNAs: target recognition and regulatory functions
.
Cell
.
2009
;
136
(
2
):
215
-
233
.
3.
Garzon
R
,
Calin
GA
,
Croce
CM
.
MicroRNAs in cancer
.
Annu Rev Med
.
2009
;
60
:
167
-
179
.
4.
Fabbri
M
,
Garzon
R
,
Andreeff
M
,
Kantarjian
HM
,
Garcia-Manero
G
,
Calin
GA
.
MicroRNAs and noncoding RNAs in hematological malignancies: molecular, clinical and therapeutic implications
.
Leukemia
.
2008
;
22
(
6
):
1095
-
1105
.
5.
Ferrara
F
,
Schiffer
CA
.
Acute myeloid leukaemia in adults
.
Lancet
.
2013
;
381
(
9865
):
484
-
495
.
6.
Dixon-McIver
A
,
East
P
,
Mein
CA
, et al
.
Distinctive patterns of microRNA expression associated with karyotype in acute myeloid leukaemia
.
PLoS One
.
2008
;
3
(
5
):
e2141
.
7.
Jongen-Lavrencic
M
,
Sun
SM
,
Dijkstra
MK
,
Valk
PJ
,
Löwenberg
B
.
MicroRNA expression profiling in relation to the genetic heterogeneity of acute myeloid leukemia
.
Blood
.
2008
;
111
(
10
):
5078
-
5085
.
8.
Li
Z
,
Lu
J
,
Sun
M
, et al
.
Distinct microRNA expression profiles in acute myeloid leukemia with common translocations
.
Proc Natl Acad Sci USA
.
2008
;
105
(
40
):
15535
-
15540
.
9.
Garzon
R
,
Garofalo
M
,
Martelli
MP
, et al
.
Distinctive microRNA signature of acute myeloid leukemia bearing cytoplasmic mutated nucleophosmin
.
Proc Natl Acad Sci USA
.
2008
;
105
(
10
):
3945
-
3950
.
10.
Marcucci
G
,
Maharry
K
,
Radmacher
MD
, et al
.
Prognostic significance of, and gene and microRNA expression signatures associated with, CEBPA mutations in cytogenetically normal acute myeloid leukemia with high-risk molecular features: a Cancer and Leukemia Group B Study
.
J Clin Oncol
.
2008
;
26
(
31
):
5078
-
5087
.
11.
Marcucci
G
,
Radmacher
MD
,
Maharry
K
, et al
.
MicroRNA expression in cytogenetically normal acute myeloid leukemia
.
N Engl J Med
.
2008
;
358
(
18
):
1919
-
1928
.
12.
Garzon
R
,
Volinia
S
,
Liu
CG
, et al
.
MicroRNA signatures associated with cytogenetics and prognosis in acute myeloid leukemia
.
Blood
.
2008
;
111
(
6
):
3183
-
3189
.
13.
Cancer Genome Atlas Research Network
,
Ley
TJ
,
Miller
C
, et al
.
Genomic and epigenomic landscapes of adult de novo acute myeloid leukemia
.
N Engl J Med
.
2013
;
368
(
22
):
2059
-
2074
.
14.
de Leeuw
DC
,
Verhagen
HJ
,
Denkers
F
, et al
.
MicroRNA-551b is highly expressed in hematopoietic stem cells and a biomarker for relapse and poor prognosis in acute myeloid leukemia
.
Leukemia
.
2016
;
30
(
3
):
742
-
746
.
15.
Starczynowski
DT
,
Morin
R
,
McPherson
A
, et al
.
Genome-wide identification of human microRNAs located in leukemia-associated genomic alterations
.
Blood
.
2011
;
117
(
2
):
595
-
607
.
16.
Chen
P
,
Price
C
,
Li
Z
, et al
.
miR-9 is an essential oncogenic microRNA specifically overexpressed in mixed lineage leukemia-rearranged leukemia
.
Proc Natl Acad Sci USA
.
2013
;
110
(
28
):
11511
-
11516
.
17.
Emmrich
S
,
Katsman-Kuipers
JE
,
Henke
K
, et al
.
miR-9 is a tumor suppressor in pediatric AML with t(8;21)
.
Leukemia
.
2014
;
28
(
5
):
1022
-
1032
.
18.
Senyuk
V
,
Zhang
Y
,
Liu
Y
, et al
.
Critical role of miR-9 in myelopoiesis and EVI1-induced leukemogenesis
.
Proc Natl Acad Sci USA
.
2013
;
110
(
14
):
5594
-
5599
.
19.
Wong
P
,
Iwasaki
M
,
Somervaille
TC
, et al
.
The miR-17-92 microRNA polycistron regulates MLL leukemia stem cell potential by modulating p21 expression
.
Cancer Res
.
2010
;
70
(
9
):
3833
-
3842
.
20.
Mi
S
,
Li
Z
,
Chen
P
, et al
.
Aberrant overexpression and function of the miR-17-92 cluster in MLL-rearranged acute leukemia
.
Proc Natl Acad Sci USA
.
2010
;
107
(
8
):
3710
-
3715
.
21.
Song
SJ
,
Ito
K
,
Ala
U
, et al
.
The oncogenic microRNA miR-22 targets the TET2 tumor suppressor to promote hematopoietic stem cell self-renewal and transformation
.
Cell Stem Cell
.
2013
;
13
(
1
):
87
-
101
.
22.
Jiang
X
,
Hu
C
,
Arnovitz
S
, et al
.
miR-22 has a potent anti-tumour role with therapeutic potential in acute myeloid leukaemia
.
Nat Commun
.
2016
;
7
:
11452
.
23.
Shen
C
,
Chen
MT
,
Zhang
XH
, et al
.
The PU.1-modulated microRNA-22 is a regulator of monocyte/macrophage differentiation and acute myeloid leukemia
.
PLoS Genet
.
2016
;
12
(
9
):
e1006259
.
24.
Gong
JN
,
Yu
J
,
Lin
HS
, et al
.
The role, mechanism and potentially therapeutic application of microRNA-29 family in acute myeloid leukemia
.
Cell Death Differ
.
2014
;
21
(
1
):
100
-
112
.
25.
Garzon
R
,
Heaphy
CE
,
Havelange
V
, et al
.
MicroRNA 29b functions in acute myeloid leukemia
.
Blood
.
2009
;
114
(
26
):
5331
-
5341
.
26.
Liu
S
,
Wu
LC
,
Pang
J
, et al
.
Sp1/NFkappaB/HDAC/miR-29b regulatory network in KIT-driven myeloid leukemia
.
Cancer Cell
.
2010
;
17
(
4
):
333
-
347
.
27.
Eyholzer
M
,
Schmid
S
,
Wilkens
L
,
Mueller
BU
,
Pabst
T
.
The tumour-suppressive miR-29a/b1 cluster is regulated by CEBPA and blocked in human AML
.
Br J Cancer
.
2010
;
103
(
2
):
275
-
284
.
28.
Bousquet
M
,
Quelen
C
,
Rosati
R
, et al
.
Myeloid cell differentiation arrest by miR-125b-1 in myelodysplastic syndrome and acute myeloid leukemia with the t(2;11)(p21;q23) translocation
.
J Exp Med
.
2008
;
205
(
11
):
2499
-
2506
.
29.
Chaudhuri
AA
,
So
AY
,
Mehta
A
, et al
.
Oncomir miR-125b regulates hematopoiesis by targeting the gene Lin28A
.
Proc Natl Acad Sci USA
.
2012
;
109
(
11
):
4233
-
4238
.
30.
O’Connell
RM
,
Chaudhuri
AA
,
Rao
DS
,
Gibson
WS
,
Balazs
AB
,
Baltimore
D
.
MicroRNAs enriched in hematopoietic stem cells differentially regulate long-term hematopoietic output
.
Proc Natl Acad Sci USA
.
2010
;
107
(
32
):
14235
-
14240
.
31.
So
AY
,
Sookram
R
,
Chaudhuri
AA
, et al
.
Dual mechanisms by which miR-125b represses IRF4 to induce myeloid and B-cell leukemias
.
Blood
.
2014
;
124
(
9
):
1502
-
1512
.
32.
Zhang
H
,
Luo
XQ
,
Zhang
P
, et al
.
MicroRNA patterns associated with clinical prognostic parameters and CNS relapse prediction in pediatric acute leukemia
.
PLoS One
.
2009
;
4
(
11
):
e7826
.
33.
Dorrance
AM
,
Neviani
P
,
Ferenchak
GJ
, et al
.
Targeting leukemia stem cells in vivo with antagomiR-126 nanoparticles in acute myeloid leukemia
.
Leukemia
.
2015
;
29
(
11
):
2143
-
2153
.
34.
Li
Z
,
Chen
P
,
Su
R
, et al
.
Overexpression and knockout of miR-126 both promote leukemogenesis
.
Blood
.
2015
;
126
(
17
):
2005
-
2015
.
35.
Lechman
ER
,
Gentner
B
,
Ng
SW
, et al
.
miR-126 Regulates Distinct Self-Renewal Outcomes in Normal and Malignant Hematopoietic Stem Cells
.
Cancer Cell
.
2016
;
29
(
2
):
214
-
228
.
36.
de Leeuw
DC
,
Denkers
F
,
Olthof
MC
, et al
.
Attenuation of microRNA-126 expression that drives CD34+38- stem/progenitor cells in acute myeloid leukemia leads to tumor eradication
.
Cancer Res
.
2014
;
74
(
7
):
2094
-
2105
.
37.
Starczynowski
DT
,
Kuchenbauer
F
,
Argiropoulos
B
, et al
.
Identification of miR-145 and miR-146a as mediators of the 5q- syndrome phenotype
.
Nat Med
.
2010
;
16
(
1
):
49
-
58
.
38.
Varney
ME
,
Niederkorn
M
,
Konno
H
, et al
.
Loss of Tifab, a del(5q) MDS gene, alters hematopoiesis through derepression of Toll-like receptor-TRAF6 signaling
.
J Exp Med
.
2015
;
212
(
11
):
1967
-
1985
.
39.
Zhao
JL
,
Rao
DS
,
Boldin
MP
,
Taganov
KD
,
O’Connell
RM
,
Baltimore
D
.
NF-kappaB dysregulation in microRNA-146a-deficient mice drives the development of myeloid malignancies
.
Proc Natl Acad Sci USA
.
2011
;
108
(
22
):
9184
-
9189
.
40.
Fang
J
,
Barker
B
,
Bolanos
L
, et al
.
Myeloid malignancies with chromosome 5q deletions acquire a dependency on an intrachromosomal NF-κB gene network
.
Cell Reports
.
2014
;
8
(
5
):
1328
-
1338
.
41.
Boldin
MP
,
Taganov
KD
,
Rao
DS
, et al
.
miR-146a is a significant brake on autoimmunity, myeloproliferation, and cancer in mice
.
J Exp Med
.
2011
;
208
(
6
):
1189
-
1201
.
42.
Gerloff
D
,
Grundler
R
,
Wurm
AA
, et al
.
NF-κB/STAT5/miR-155 network targets PU.1 in FLT3-ITD-driven acute myeloid leukemia
.
Leukemia
.
2015
;
29
(
3
):
535
-
547
.
43.
Marcucci
G
,
Maharry
KS
,
Metzeler
KH
, et al
.
Clinical role of microRNAs in cytogenetically normal acute myeloid leukemia: miR-155 upregulation independently identifies high-risk patients
.
J Clin Oncol
.
2013
;
31
(
17
):
2086
-
2093
.
44.
Schneider
E
,
Staffas
A
,
Röhner
L
, et al
.
MicroRNA-155 is upregulated in MLL-rearranged AML but its absence does not affect leukemia development
.
Exp Hematol
.
2016
;
44
(
12
):
1166
-
1171
.
45.
Wallace
JA
,
Kagele
DA
,
Eiring
AM
, et al
.
miR-155 promotes FLT3-ITD-induced myeloproliferative disease through inhibition of the interferon response
.
Blood
.
2017
;
129
(
23
):
3074
-
3086
.
46.
O’Connell
RM
,
Rao
DS
,
Chaudhuri
AA
, et al
.
Sustained expression of microRNA-155 in hematopoietic stem cells causes a myeloproliferative disorder
.
J Exp Med
.
2008
;
205
(
3
):
585
-
594
.
47.
Li
Y
,
Gao
L
,
Luo
X
, et al
.
Epigenetic silencing of microRNA-193a contributes to leukemogenesis in t(8;21) acute myeloid leukemia by activating the PTEN/PI3K signal pathway
.
Blood
.
2013
;
121
(
3
):
499
-
509
.
48.
Gao
XN
,
Lin
J
,
Li
YH
, et al
.
MicroRNA-193a represses c-kit expression and functions as a methylation-silenced tumor suppressor in acute myeloid leukemia
.
Oncogene
.
2011
;
30
(
31
):
3416
-
3428
.
49.
Li
Z
,
Huang
H
,
Chen
P
, et al
.
miR-196b directly targets both HOXA9/MEIS1 oncogenes and FAS tumour suppressor in MLL-rearranged leukaemia
.
Nat Commun
.
2012
;
3
:
688
.
50.
Popovic
R
,
Riesbeck
LE
,
Velu
CS
, et al
.
Regulation of mir-196b by MLL and its overexpression by MLL fusions contributes to immortalization
.
Blood
.
2009
;
113
(
14
):
3314
-
3322
.
51.
Pulikkan
JA
,
Dengler
V
,
Peramangalam
PS
, et al
.
Cell-cycle regulator E2F1 and microRNA-223 comprise an autoregulatory negative feedback loop in acute myeloid leukemia
.
Blood
.
2010
;
115
(
9
):
1768
-
1778
.
52.
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
-
3536
.
53.
Gentner
B
,
Pochert
N
,
Rouhi
A
, et al
.
MicroRNA-223 dose levels fine tune proliferation and differentiation in human cord blood progenitors and acute myeloid leukemia
. Exp Hematol.
2015
;43(10):858-868.e7.
54.
Fazi
F
,
Racanicchi
S
,
Zardo
G
, et al
.
Epigenetic silencing of the myelopoiesis regulator microRNA-223 by the AML1/ETO oncoprotein
.
Cancer Cell
.
2007
;
12
(
5
):
457
-
466
.
55.
Han
YC
,
Park
CY
,
Bhagat
G
, et al
.
microRNA-29a induces aberrant self-renewal capacity in hematopoietic progenitors, biased myeloid development, and acute myeloid leukemia
.
J Exp Med
.
2010
;
207
(
3
):
475
-
489
.
56.
Garzon
R
,
Liu
S
,
Fabbri
M
, et al
.
MicroRNA-29b induces global DNA hypomethylation and tumor suppressor gene reexpression in acute myeloid leukemia by targeting directly DNMT3A and 3B and indirectly DNMT1
.
Blood
.
2009
;
113
(
25
):
6411
-
6418
.
57.
Cheng
J
,
Guo
S
,
Chen
S
, et al
.
An extensive network of TET2-targeting MicroRNAs regulates malignant hematopoiesis
.
Cell Reports
.
2013
;
5
(
2
):
471
-
481
.
58.
Ramsingh
G
,
Jacoby
MA
,
Shao
J
, et al
.
Acquired copy number alterations of miRNA genes in acute myeloid leukemia are uncommon
.
Blood
.
2013
;
122
(
15
):
e44
-
e51
.
59.
Merritt
WM
,
Lin
YG
,
Han
LY
, et al
.
Dicer, Drosha, and outcomes in patients with ovarian cancer
.
N Engl J Med
.
2008
;
359
(
25
):
2641
-
2650
.
60.
Nakamura
T
,
Canaani
E
,
Croce
CM
.
Oncogenic All1 fusion proteins target Drosha-mediated microRNA processing
.
Proc Natl Acad Sci USA
.
2007
;
104
(
26
):
10980
-
10985
.
61.
Thol
F
,
Scherr
M
,
Kirchner
A
, et al
.
Clinical and functional implications of microRNA mutations in a cohort of 935 patients with myelodysplastic syndromes and acute myeloid leukemia
.
Haematologica
.
2015
;
100
(
4
):
e122
-
e124
.
62.
Dzikiewicz-Krawczyk
A
,
Macieja
A
,
Mały
E
, et al
.
Polymorphisms in microRNA target sites modulate risk of lymphoblastic and myeloid leukemias and affect microRNA binding
.
J Hematol Oncol
.
2014
;
7
:
43
.
63.
Grasedieck
S
,
Sorrentino
A
,
Langer
C
, et al
.
Circulating microRNAs in hematological diseases: principles, challenges, and perspectives
.
Blood
.
2013
;
121
(
25
):
4977
-
4984
.
64.
de Leeuw
DC
,
van den Ancker
W
,
Denkers
F
, et al
.
MicroRNA profiling can classify acute leukemias of ambiguous lineage as either acute myeloid leukemia or acute lymphoid leukemia
.
Clin Cancer Res
.
2013
;
19
(
8
):
2187
-
2196
.
65.
Mi
S
,
Lu
J
,
Sun
M
, et al
.
MicroRNA expression signatures accurately discriminate acute lymphoblastic leukemia from acute myeloid leukemia
.
Proc Natl Acad Sci USA
.
2007
;
104
(
50
):
19971
-
19976
.
66.
Díaz-Beyá
M
,
Brunet
S
,
Nomdedéu
J
, et al
;
Cooperative AML group CETLAM (Grupo Cooperativo Para el Estudio y Tratamiento de las Leucemias Agudas y Mielodisplasias)
.
MicroRNA expression at diagnosis adds relevant prognostic information to molecular categorization in patients with intermediate-risk cytogenetic acute myeloid leukemia
.
Leukemia
.
2014
;
28
(
4
):
804
-
812
.
67.
Marcucci
G
,
Mrózek
K
,
Radmacher
MD
,
Garzon
R
,
Bloomfield
CD
.
The prognostic and functional role of microRNAs in acute myeloid leukemia
.
Blood
.
2011
;
117
(
4
):
1121
-
1129
.
68.
Sun
SM
,
Rockova
V
,
Bullinger
L
, et al
.
The prognostic relevance of miR-212 expression with survival in cytogenetically and molecularly heterogeneous AML
.
Leukemia
.
2013
;
27
(
1
):
100
-
106
.
69.
Guo
Y
,
Strickland
SA
,
Mohan
S
, et al
.
MicroRNAs and tRNA-derived fragments predict the transformation of myelodysplastic syndromes to acute myeloid leukemia
.
Leuk Lymphoma
.
2017
;
58
(
9
):
1
-
15
.
70.
Hourigan
CS
,
Gale
RP
,
Gormley
NJ
,
Ossenkoppele
GJ
,
Walter
RB
.
Measurable residual disease testing in acute myeloid leukaemia
.
Leukemia
.
2017
;
31
(
7
):
1482
-
1490
.
71.
Yin
JA
,
O’Brien
MA
,
Hills
RK
,
Daly
SB
,
Wheatley
K
,
Burnett
AK
.
Minimal residual disease monitoring by quantitative RT-PCR in core binding factor AML allows risk stratification and predicts relapse: results of the United Kingdom MRC AML-15 trial
.
Blood
.
2012
;
120
(
14
):
2826
-
2835
.
72.
Koutova
L
,
Sterbova
M
,
Pazourkova
E
, et al
.
The impact of standard chemotherapy on miRNA signature in plasma in AML patients
.
Leuk Res
.
2015
;
39
(
12
):
1389
-
1395
.
73.
Zhi
F
,
Cao
X
,
Xie
X
, et al
.
Identification of circulating microRNAs as potential biomarkers for detecting acute myeloid leukemia
.
PLoS One
.
2013
;
8
(
2
):
e56718
.
74.
Hornick
NI
,
Huan
J
,
Doron
B
, et al
.
Serum exosome microRNA as a minimally-invasive early biomarker of AML
.
Sci Rep
.
2015
;
5
:
11295
.
75.
Huang
X
,
Schwind
S
,
Yu
B
, et al
.
Targeted delivery of microRNA-29b by transferrin-conjugated anionic lipopolyplex nanoparticles: a novel therapeutic strategy in acute myeloid leukemia
.
Clin Cancer Res
.
2013
;
19
(
9
):
2355
-
2367
.
76.
Huang
X
,
Schwind
S
,
Santhanam
R
, et al
.
Targeting the RAS/MAPK pathway with miR-181a in acute myeloid leukemia
.
Oncotarget
.
2016
;
7
(
37
):
59273
-
59286
.
77.
Velu
CS
,
Chaubey
A
,
Phelan
JD
, et al
.
Therapeutic antagonists of microRNAs deplete leukemia-initiating cell activity
.
J Clin Invest
.
2014
;
124
(
1
):
222
-
236
.
78.
Khalife
J
,
Radomska
HS
,
Santhanam
R
, et al
.
Pharmacological targeting of miR-155 via the NEDD8-activating enzyme inhibitor MLN4924 (Pevonedistat) in FLT3-ITD acute myeloid leukemia
.
Leukemia
.
2015
;
29
(
10
):
1981
-
1992
.
79.
Shah
NM
,
Zaitseva
L
,
Bowles
KM
,
MacEwan
DJ
,
Rushworth
SA
.
NRF2-driven miR-125B1 and miR-29B1 transcriptional regulation controls a novel anti-apoptotic miRNA regulatory network for AML survival
.
Cell Death Differ
.
2015
;
22
(
4
):
654
-
664
.
80.
Blum
W
,
Garzon
R
,
Klisovic
RB
, et al
.
Clinical response and miR-29b predictive significance in older AML patients treated with a 10-day schedule of decitabine
.
Proc Natl Acad Sci USA
.
2010
;
107
(
16
):
7473
-
7478
.
81.
Lu
F
,
Zhang
J
,
Ji
M
, et al
.
miR-181b increases drug sensitivity in acute myeloid leukemia via targeting HMGB1 and Mcl-1
.
Int J Oncol
.
2014
;
45
(
1
):
383
-
392
.
82.
Li
Z
,
Rana
TM
.
Therapeutic targeting of microRNAs: current status and future challenges
.
Nat Rev Drug Discov
.
2014
;
13
(
8
):
622
-
638
.
83.
Bouchie
A
.
First microRNA mimic enters clinic
.
Nat Biotechnol
.
2013
;
31
(
7
):
577
.
84.
Janssen
HL
,
Reesink
HW
,
Lawitz
EJ
, et al
.
Treatment of HCV infection by targeting microRNA
.
N Engl J Med
.
2013
;
368
(
18
):
1685
-
1694
.
85.
Chen
Y
,
Gao
DY
,
Huang
L
.
In vivo delivery of miRNAs for cancer therapy: challenges and strategies
.
Adv Drug Deliv Rev
.
2015
;
81
:
128
-
141
.
86.
Karin
M
.
NF-kappaB as a critical link between inflammation and cancer
.
Cold Spring Harb Perspect Biol
.
2009
;
1
(
5
):
a000141
.
87.
Iliopoulos
D
,
Hirsch
HA
,
Struhl
K
.
An epigenetic switch involving NF-kappaB, Lin28, Let-7 MicroRNA, and IL6 links inflammation to cell transformation
.
Cell
.
2009
;
139
(
4
):
693
-
706
.
88.
Xiang
M
,
Birkbak
NJ
,
Vafaizadeh
V
, et al
.
STAT3 induction of miR-146b forms a feedback loop to inhibit the NF-κB to IL-6 signaling axis and STAT3-driven cancer phenotypes
.
Sci Signal
.
2014
;
7
(
310
):
ra11
.
89.
Rhyasen
GW
,
Bolanos
L
,
Fang
J
, et al
.
Targeting IRAK1 as a therapeutic approach for myelodysplastic syndrome
.
Cancer Cell
.
2013
;
24
(
1
):
90
-
104
.
90.
Zhang
YC
,
Ye
H
,
Zeng
Z
,
Chin
YE
,
Huang
YN
,
Fu
GH
.
The NF-κB p65/miR-23a-27a-24 cluster is a target for leukemia treatment
.
Oncotarget
.
2015
;
6
(
32
):
33554
-
33567
.
91.
Zambetti
NA
,
Ping
Z
,
Chen
S
, et al
.
Mesenchymal inflammation drives genotoxic stress in hematopoietic stem cells and predicts disease evolution in human pre-leukemia
.
Cell Stem Cell
.
2016
;
19
(
5
):
613
-
627
.
92.
O’Connell
RM
,
Baltimore
D
.
MicroRNAs and hematopoietic cell development
.
Curr Top Dev Biol
.
2012
;
99
:
145
-
174
.
93.
Chiu
YC
,
Tsai
MH
,
Chou
WC
, et al
.
Prognostic significance of NPM1 mutation-modulated microRNA-mRNA regulation in acute myeloid leukemia
.
Leukemia
.
2016
;
30
(
2
):
274
-
284
.
94.
Bousquet
M
,
Harris
MH
,
Zhou
B
,
Lodish
HF
.
MicroRNA miR-125b causes leukemia
.
Proc Natl Acad Sci USA
.
2010
;
107
(
50
):
21558
-
21563
.
95.
Ooi
AG
,
Sahoo
D
,
Adorno
M
,
Wang
Y
,
Weissman
IL
,
Park
CY
.
MicroRNA-125b expands hematopoietic stem cells and enriches for the lymphoid-balanced and lymphoid-biased subsets
.
Proc Natl Acad Sci USA
.
2010
;
107
(
50
):
21505
-
21510
.
96.
Palma
CA
,
Al Sheikha
D
,
Lim
TK
, et al
.
MicroRNA-155 as an inducer of apoptosis and cell differentiation in acute myeloid leukaemia
.
Mol Cancer
.
2014
;
13
:
79
.
97.
Narayan
N
,
Morenos
L
,
Phipson
B
, et al
.
Functionally distinct roles for different miR-155 expression levels through contrasting effects on gene expression, in acute myeloid leukaemia
.
Leukemia
.
2017
;31(4):808-829.
98.
Jiang
L
,
Huang
Q
,
Zhang
S
, et al
.
Hsa-miR-125a-3p and hsa-miR-125a-5p are downregulated in non-small cell lung cancer and have inverse effects on invasion and migration of lung cancer cells
.
BMC Cancer
.
2010
;
10
:
318
.
99.
Almeida
MI
,
Nicoloso
MS
,
Zeng
L
, et al
.
Strand-specific miR-28-5p and miR-28-3p have distinct effects in colorectal cancer cells
. Gastroenterology.
2012
;142(4):886-896.e9.
100.
Kuchenbauer
F
,
Mah
SM
,
Heuser
M
, et al
.
Comprehensive analysis of mammalian miRNA* species and their role in myeloid cells
.
Blood
.
2011
;
118
(
12
):
3350
-
3358
.
101.
Nowek
K
,
Sun
SM
,
Dijkstra
MK
, et al
.
Expression of a passenger miR-9* predicts favorable outcome in adults with acute myeloid leukemia less than 60 years of age
.
Leukemia
.
2016
;
30
(
2
):
303
-
309
.
102.
Meijer
HA
,
Smith
EM
,
Bushell
M
.
Regulation of miRNA strand selection: follow the leader?
Biochem Soc Trans
.
2014
;
42
(
4
):
1135
-
1140
.
103.
Ha
M
,
Kim
VN
.
Regulation of microRNA biogenesis
.
Nat Rev Mol Cell Biol
.
2014
;
15
(
8
):
509
-
524
.
104.
Mangiavacchi
A
,
Sorci
M
,
Masciarelli
S
, et al
.
The miR-223 host non-coding transcript linc-223 induces IRF4 expression in acute myeloid leukemia by acting as a competing endogenous RNA
.
Oncotarget
.
2016
;
7
(
37
):
60155
-
60168
.
105.
Brameier
M
,
Herwig
A
,
Reinhardt
R
,
Walter
L
,
Gruber
J
.
Human box C/D snoRNAs with miRNA like functions: expanding the range of regulatory RNAs
.
Nucleic Acids Res
.
2011
;
39
(
2
):
675
-
686
.
106.
Ender
C
,
Krek
A
,
Friedländer
MR
, et al
.
A human snoRNA with microRNA-like functions
.
Mol Cell
.
2008
;
32
(
4
):
519
-
528
.
107.
Ruby
JG
,
Jan
CH
,
Bartel
DP
.
Intronic microRNA precursors that bypass Drosha processing
.
Nature
.
2007
;
448
(
7149
):
83
-
86
.
108.
Loeb
GB
,
Khan
AA
,
Canner
D
, et al
.
Transcriptome-wide miR-155 binding map reveals widespread noncanonical microRNA targeting
.
Mol Cell
.
2012
;
48
(
5
):
760
-
770
.
109.
Kuchenbauer
F
,
Morin
RD
,
Argiropoulos
B
, et al
.
In-depth characterization of the microRNA transcriptome in a leukemia progression model
.
Genome Res
.
2008
;
18
(
11
):
1787
-
1797
.
110.
Ebhardt
HA
,
Tsang
HH
,
Dai
DC
,
Liu
Y
,
Bostan
B
,
Fahlman
RP
.
Meta-analysis of small RNA-sequencing errors reveals ubiquitous post-transcriptional RNA modifications
.
Nucleic Acids Res
.
2009
;
37
(
8
):
2461
-
2470
.
111.
Wyman
SK
,
Knouf
EC
,
Parkin
RK
, et al
.
Post-transcriptional generation of miRNA variants by multiple nucleotidyl transferases contributes to miRNA transcriptome complexity
.
Genome Res
.
2011
;
21
(
9
):
1450
-
1461
.
112.
Place
RF
,
Li
LC
,
Pookot
D
,
Noonan
EJ
,
Dahiya
R
.
MicroRNA-373 induces expression of genes with complementary promoter sequences
.
Proc Natl Acad Sci USA
.
2008
;
105
(
5
):
1608
-
1613
.
113.
Lytle
JR
,
Yario
TA
,
Steitz
JA
.
Target mRNAs are repressed as efficiently by microRNA-binding sites in the 5′ UTR as in the 3′ UTR
.
Proc Natl Acad Sci USA
.
2007
;
104
(
23
):
9667
-
9672
.
114.
Tay
Y
,
Zhang
J
,
Thomson
AM
,
Lim
B
,
Rigoutsos
I
.
MicroRNAs to Nanog, Oct4 and Sox2 coding regions modulate embryonic stem cell differentiation
.
Nature
.
2008
;
455
(
7216
):
1124
-
1128
.
115.
Forman
JJ
,
Legesse-Miller
A
,
Coller
HA
.
A search for conserved sequences in coding regions reveals that the let-7 microRNA targets Dicer within its coding sequence
.
Proc Natl Acad Sci USA
.
2008
;
105
(
39
):
14879
-
14884
.
116.
Eiring
AM
,
Harb
JG
,
Neviani
P
, et al
.
miR-328 functions as an RNA decoy to modulate hnRNP E2 regulation of mRNA translation in leukemic blasts
.
Cell
.
2010
;
140
(
5
):
652
-
665
.
117.
Valadi
H
,
Ekström
K
,
Bossios
A
,
Sjöstrand
M
,
Lee
JJ
,
Lötvall
JO
.
Exosome-mediated transfer of mRNAs and microRNAs is a novel mechanism of genetic exchange between cells
.
Nat Cell Biol
.
2007
;
9
(
6
):
654
-
659
.
118.
Alexander
M
,
Hu
R
,
Runtsch
MC
, et al
.
Exosome-delivered microRNAs modulate the inflammatory response to endotoxin
.
Nat Commun
.
2015
;
6
:
7321
.
119.
Melo
SA
,
Sugimoto
H
,
O’Connell
JT
, et al
.
Cancer exosomes perform cell-independent microRNA biogenesis and promote tumorigenesis
.
Cancer Cell
.
2014
;
26
(
5
):
707
-
721
.
120.
Huan
J
,
Hornick
NI
,
Shurtleff
MJ
, et al
.
RNA trafficking by acute myelogenous leukemia exosomes
.
Cancer Res
.
2013
;
73
(
2
):
918
-
929
.
121.
Huan
J
,
Hornick
NI
,
Goloviznina
NA
, et al
.
Coordinate regulation of residual bone marrow function by paracrine trafficking of AML exosomes
.
Leukemia
.
2015
;
29
(
12
):
2285
-
2295
.
122.
Hornick
NI
,
Doron
B
,
Abdelhamed
S
, et al
.
AML suppresses hematopoiesis by releasing exosomes that contain microRNAs targeting c-MYB
.
Sci Signal
.
2016
;
9
(
444
):
ra88
.
123.
Rinn
JL
,
Chang
HY
.
Genome regulation by long noncoding RNAs
.
Annu Rev Biochem
.
2012
;
81
:
145
-
166
.
124.
Kallen
AN
,
Zhou
XB
,
Xu
J
, et al
.
The imprinted H19 lncRNA antagonizes let-7 microRNAs
.
Mol Cell
.
2013
;
52
(
1
):
101
-
112
.
125.
Díaz-Beyá
M
,
Brunet
S
,
Nomdedéu
J
, et al
;
Cooperative AML group CETLAM
.
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
-
31627
.
126.
Chen
ZH
,
Wang
WT
,
Huang
W
, et al
.
The lncRNA HOTAIRM1 regulates the degradation of PML-RARA oncoprotein and myeloid cell differentiation by enhancing the autophagy pathway
.
Cell Death Differ
.
2017
;
24
(
2
):
212
-
224
.
127.
Xing
CY
,
Hu
XQ
,
Xie
FY
, et al
.
Long non-coding RNA HOTAIR modulates c-KIT expression through sponging miR-193a in acute myeloid leukemia
.
FEBS Lett
.
2015
;
589
(
15
):
1981
-
1987
.
128.
Chen
L
,
Wang
W
,
Cao
L
,
Li
Z
,
Wang
X
.
Long non-coding RNA CCAT1 acts as a competing endogenous RNA to regulate cell growth and differentiation in acute myeloid leukemia
.
Mol Cells
.
2016
;
39
(
4
):
330
-
336
.
129.
Guo
G
,
Kang
Q
,
Zhu
X
, 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
-
1779
.
130.
Huppi
K
,
Volfovsky
N
,
Runfola
T
, et al
.
The identification of microRNAs in a genomically unstable region of human chromosome 8q24
.
Mol Cancer Res
.
2008
;
6
(
2
):
212
-
221
.
131.
Beck-Engeser
GB
,
Lum
AM
,
Huppi
K
,
Caplen
NJ
,
Wang
BB
,
Wabl
M
.
Pvt1-encoded microRNAs in oncogenesis
.
Retrovirology
.
2008
;
5
:
4
.
132.
Wallace
J
,
Hu
R
,
Mosbruger
TL
, et al
.
Genome-Wide CRISPR-Cas9 Screen Identifies MicroRNAs That Regulate Myeloid Leukemia Cell Growth
.
PLoS One
.
2016
;
11
(
4
):
e0153689
.
Sign in via your Institution