MicroRNAs are small RNA molecules that modulate protein expression by degrading mRNA or repressing translation. They have been shown to play important roles in hematopoiesis, including embryonic stem cell differentiation, erythropoiesis, granulocytopoiesis/monocytopoiesis, lymphopoiesis, and megakaryocytopoiesis. miR-150 and miR-155 play divergent roles in megakaryocytopoiesis, with the former promoting development of megakaryocytes at the expense of erythrocytes and the latter causing a reduction in megakaryocyte colony formation. Platelets also contain fully functional miRNA machinery, and certain miRNA levels in platelets have been found to coordinate with reactivity to specific agonists and to pathologic states. This review will cover the current state of knowledge of miRNAs in megakaryocytes and platelets and the exciting possibilities for future research.

The linear string of nucleotides that forms a genome has a simplicity that is incongruous with the sometimes overwhelming complexity of organism anatomy and physiology. Some order is achieved by considering messenger RNA (mRNA), linking the proteins responsible for organism intricacy with specific genomic locations. But this “coding gene centric” viewpoint does not easily explain the large amount of noncoding DNA sequence, representing approximately 70% of the genome.1  Although protein coding genes may substantively contribute to the increased complexity of Caenorhabditis elegans compared with Escherichia coli (a 439% increase in gene number), differences in protein coding genes do not seem to sufficiently explain the greater complexity between Homo sapiens and C elegans (an increase of only 21%). On the other hand, unlike C elegans, humans have 21-fold more transcribed, noncoding sequence than transcribed, coding sequence, whereas C elegans has only 1.25-fold.2  This intriguing association between organism complexity and transcribed, noncoding sequence suggests a role for noncoding RNA, and the last decade has clearly established an important role for noncoding RNA in biology. Noncoding RNAs are classified as large (> 200 bp) or small (< 200 bp). Small regulatory RNAs include microRNAs (miRNAs), endogenous small interfering RNAs, piwi-interacting RNAs, small nucleolar RNAs, and others.3,4  The best studied noncoding RNA are miRNAs. miRNAs were discovered in C elegans in 19935,6  and have been shown to play important roles in developmental biology, cellular stress, circadian rhythm, and immunology, as well as numerous disease states, including Alzheimer disease, cancer, and heart failure. Abundant evidence demonstrates a critical role for miRNAs in normal human hematopoiesis, and dysregulated miRNA biology has been associated with and shown to cause numerous hematologic diseases (Table 1). This review will focus on the role of miRNAs in megakaryocytopoiesis and platelet biology.

Table 1

miRNA dysregulation and hematologic disease

Hematologic disorderImplicated miRReferences
Acute leukemia miR-125b-2, miR-29b, miR-181a, miR-204 61,86,,89  
Chronic leukemia miR-15a, miR-16-1, miR-155, miR-29b, miR-181a, miR-328 58,90,,,,,96  
Myeloproliferative disorders miR-150, let-7a, miR-182, miR-26b, miR-143, miR-145, miR-223,miR-26b, miR-30b, miR-30c 31,65  
Myelodysplasia miR-145, miR-146a 51  
lymphoma miR-21, miR-155, miR-17-92 cluster 97,99  
Hematologic disorderImplicated miRReferences
Acute leukemia miR-125b-2, miR-29b, miR-181a, miR-204 61,86,,89  
Chronic leukemia miR-15a, miR-16-1, miR-155, miR-29b, miR-181a, miR-328 58,90,,,,,96  
Myeloproliferative disorders miR-150, let-7a, miR-182, miR-26b, miR-143, miR-145, miR-223,miR-26b, miR-30b, miR-30c 31,65  
Myelodysplasia miR-145, miR-146a 51  
lymphoma miR-21, miR-155, miR-17-92 cluster 97,99  

miRNAs are 21 to 23 nucleotide regulatory RNAs expressed in multicellular organisms, from viruses to plants to humans.7  There are at least 400 human miRNAs, although some predictions estimate more than 1000.8,9  miRNAs regulate most (> 60%) mammalian protein coding genes primarily by repressing gene expression.10  Some miRNAs are expressed ubiquitously, but many are tissue and/or developmental stage–specific.11  Cell miRNA content is highly variable from 1 copy to 10 000 copies.12 

Approximately 30% of miRNA genes are located in intergenic regions, and approximately 70% are located within introns or exons of protein coding genes. Figure 1A illustrates the canonical understanding of intergenic miRNA biogenesis, although exceptions have been described to some of these steps. miRNAs located within introns are transcribed as part of the surrounding coding gene's mRNA, whereas miRNAs located intergenically use an independent transcriptional apparatus. RNA polymerase II is responsible for most miRNA transcription, although RNA polymerase III has been identified transcribing miRNAs located within repetitive elements.13  A several-kilobase primary transcript (pri-miRNA) is then capped and polyadenylated, and forms a hairpin structure.14,15  The pri-miRNA is cleaved into an approximately 60- to 70-bp pre-miRNA by Drosha, an RNase type III endonuclease, which is complexed with DiGeorge syndrome critical region 8 (DGCR8). The pre-miRNA is transported out of the nucleus via exportin 5. In the cytoplasm, the 3′ overhang of the pre-miRNA is recognized by the Dicer-TAR RNA-binding protein (TRBP) complex. Dicer is another RNase type III endonuclease, which generates the miRNA duplex for many miRNAs. The strands separate and the mature miRNA associates with a macromolecular complex called the RNA-induced silencing complex (RISC), which guides the miRNA to its mRNA target. The process of miRNA generation is regulated at both transcriptional and posttranscriptional levels, occasionally composed of positive feed-forward or negative feedback circuitry, in which the miRNA targets a transcriptional activator or repressor of itself. The defining RISC protein is Argonaute 2 (Ago2), which catalyzes mRNA cleavage. Notably, Ago2-dependent, dicer-independent miRNA processing pathways have been identified.16,17 

Figure 1

miRNA biogenesis and function. (A) The canonical miRNA biosynthesis pathway. Drosha indicates an RNase type III endonuclease; DGCR8, DiGeorge syndrome critical region 8; Dicer, another RNase type III endonuclease; TRBP, TAR RNA-binding protein; RISC, RNA-induced silencing complex; Ago, Argonaute 2; and ORF, open reading frame. (B) The seed region of the miRNA (bp 2-8) binds to complementary sites on the 3′-UTR of target mRNAs where it predominantly causes a decrease in mRNA levels but may also inhibit translation.

Figure 1

miRNA biogenesis and function. (A) The canonical miRNA biosynthesis pathway. Drosha indicates an RNase type III endonuclease; DGCR8, DiGeorge syndrome critical region 8; Dicer, another RNase type III endonuclease; TRBP, TAR RNA-binding protein; RISC, RNA-induced silencing complex; Ago, Argonaute 2; and ORF, open reading frame. (B) The seed region of the miRNA (bp 2-8) binds to complementary sites on the 3′-UTR of target mRNAs where it predominantly causes a decrease in mRNA levels but may also inhibit translation.

Close modal

A fundamental aspect of miRNA function relates to mRNA targeting: most miRNAs are predicted to target multiple mRNAs, and most mRNAs have predicted targets for many miRNAs. Different algorithms are publicly available that predict miRNA binding sites, including Miranda, TargetScan, PicTar, and miRBase. The miRNA associates with mRNA by complementary Watson-Crick base pairing (Figure 1B), with sequences usually, but not always,18  in the 3′-untranslated region (UTR) of mRNA. The seed region of nucleotides 2 to 8 at the 5′ end of the miRNA has perfect complementarity, and this sequence defines families of miRNAs. The miRNA sequence 3′ to the seed sequence has variable degrees of complementarity with the mRNA. The traditional view of miRNA-mediated translation repression posits that lower complementarity induces translation inhibition. But recent work by Guo et al indicates that miRNA knock-down of protein expression is primarily via mRNA degradation, with translation inhibition representing only a minor mechanism.19  miRNAs have been aptly referred to as “rheostats” because their regulatory impact is generally to fine-tune but not abolish protein expression.20,21  The significance of proper miRNA synthesis and function is underscored by diseases caused by genetic defects at virtually all steps in miRNA biogenesis and targeting, including miRNA gene deletions/duplications, single nucleotide polymorphisms in the target mRNA 3′-UTR, miRNA “decoys,” and genetic variants in the proteins mediating miRNA biogenesis.22-24  Importantly, small differences (as little as a 20% change) in miRNA levels have been shown to cause autoimmune disease and predispose to malignancy.25 

Traditionally, mechanistic studies on hematopoietic cell self-renewal and differentiation have focused primarily on transcription factors.26  Over the past 6 to 7 years, numerous laboratories have established the essential role of Dicer127  and miRNAs in hematopoiesis,28-30  including embryonic stem cell differentiation,18  erythropoiesis,31-36  granulocytopoiesis/monocytopoiesis,37-40  and lymphopoiesis.28,41-44  Since the first report by Garzon et al in 2006,45  nearly 20 studies have addressed various aspects of miRNAs in megakaryocytopoiesis using megakaryocytes generated from in vitro cultured CD34+ hematopoietic stem cells (HSCs) or transformed cell lines with megakaryocytic properties (Table 2). Both unbiased miRNA profiling and candidate miRNA studies have established associations between specific miRNAs and the developmental stage of megakaryocyte progenitors. Functionality has been tested by assessing the effects of candidate miRNAs on in vitro and in vivo proliferation and differentiation after overexpression or knock-down. Mechanistic assessment of these miRNAs has included target protein knock-down and reporter gene assays using constructs with the putative target 3′-UTR. Several of the best-studied miRNAs will be discussed for their role in megakaryocytopoiesis.

Table 2

miRNAs in megakaryocyte literature

StudyFindings
Georgantas et al29  Overexpression of miR-155 in K562 cells caused a block in megakaryocytic differentiation. 
Garzon et al45  miRNA profile of in vitro differentiated CD34+ bone marrow cells. Identified miR-130a targets MAFB and miR-10a targets HOXA1. 
O'Connell et al40  Overexpression of miR-155 in transplanted bone marrow led to a decrease in erythrocytes, megakaryocytes (MKs), and lymphocytes. 
Guimaraes-Sternberg et al63  Calcium release suppresses pre-miR-181a in Meg-01 cells. A miR-181a analog blocked Ca+-induced differentiation and induced apoptosis. 
Labbaye et al50  The PLZF transcription factor represses levels of miR-146a, which in turn targets CXCR4 expression. Forced expression or repression alters MK development. 
Lu et al47  Expression of miR-150 drives MK-erythroid precurors toward the megakaryocytic fate in murine bone marrow transplant experiments. 
Barroga et al49  Thrombopoietin induces expression of miR-150 in UT-7/TPO cells, which in turn, targets expression of the transcription factor c-Myb. 
Romania et al46  miR-155 levels decrease during megakaryopoiesis in cultured human cord blood. Enforced expression of miR-155 impairs MK proliferation and development. 
Hussein et al74  miRNA profile of laser-dissected MKs from primary myelofibrosis essential thrombocythemia patients. 
Navarro et al59  miR-34a is rapidly increased during TPA-induced differentiation of K562 cells into MKs. Overexpression of miR-34a enhances MK differentiation in HSCs and regulates c-Myb expression. 
Ben-Ami et al60  TPA induces Runx1 binding to the miR-27a regulatory region and causes an increase in miR-27a in K562 cells. miR-27a, in turn, targets and suppresses Runx1 levels. 
Giradot et al62  miR-28 targets the thrombopoietin receptor, MPL. Enforced expression prevents MK differentiation in CD34+ cells and is overexpressed in a fraction of platelets in patients with myeloproliferative diseases. 
Opalinska et al52  miR-146a increases during MK development. 
Starczynowski et al53  Enforced expression of miR-146a in transduced bone marrow cells has no effect on platelet number. 
Klusmann et al 61  miR-125b-2 is overexpressed in Down syndrome, acute megakaryoblastic leukemia. miR-125b-2 overexpression induces proliferation and differentiation of MK and MK/erythroid precursors. 
Ichimura et al58  Phorbol myristate acetate treatment in K562 cells increases miR-34a levels. miR-34a, in turn, targets mitogen-activated protein kinase kinase 1 (MEK1) and represses proliferation. 
Starczynowski et al51  miR-145 and miR-146a are mediators of the 5q-syndrome phenotype. 
StudyFindings
Georgantas et al29  Overexpression of miR-155 in K562 cells caused a block in megakaryocytic differentiation. 
Garzon et al45  miRNA profile of in vitro differentiated CD34+ bone marrow cells. Identified miR-130a targets MAFB and miR-10a targets HOXA1. 
O'Connell et al40  Overexpression of miR-155 in transplanted bone marrow led to a decrease in erythrocytes, megakaryocytes (MKs), and lymphocytes. 
Guimaraes-Sternberg et al63  Calcium release suppresses pre-miR-181a in Meg-01 cells. A miR-181a analog blocked Ca+-induced differentiation and induced apoptosis. 
Labbaye et al50  The PLZF transcription factor represses levels of miR-146a, which in turn targets CXCR4 expression. Forced expression or repression alters MK development. 
Lu et al47  Expression of miR-150 drives MK-erythroid precurors toward the megakaryocytic fate in murine bone marrow transplant experiments. 
Barroga et al49  Thrombopoietin induces expression of miR-150 in UT-7/TPO cells, which in turn, targets expression of the transcription factor c-Myb. 
Romania et al46  miR-155 levels decrease during megakaryopoiesis in cultured human cord blood. Enforced expression of miR-155 impairs MK proliferation and development. 
Hussein et al74  miRNA profile of laser-dissected MKs from primary myelofibrosis essential thrombocythemia patients. 
Navarro et al59  miR-34a is rapidly increased during TPA-induced differentiation of K562 cells into MKs. Overexpression of miR-34a enhances MK differentiation in HSCs and regulates c-Myb expression. 
Ben-Ami et al60  TPA induces Runx1 binding to the miR-27a regulatory region and causes an increase in miR-27a in K562 cells. miR-27a, in turn, targets and suppresses Runx1 levels. 
Giradot et al62  miR-28 targets the thrombopoietin receptor, MPL. Enforced expression prevents MK differentiation in CD34+ cells and is overexpressed in a fraction of platelets in patients with myeloproliferative diseases. 
Opalinska et al52  miR-146a increases during MK development. 
Starczynowski et al53  Enforced expression of miR-146a in transduced bone marrow cells has no effect on platelet number. 
Klusmann et al 61  miR-125b-2 is overexpressed in Down syndrome, acute megakaryoblastic leukemia. miR-125b-2 overexpression induces proliferation and differentiation of MK and MK/erythroid precursors. 
Ichimura et al58  Phorbol myristate acetate treatment in K562 cells increases miR-34a levels. miR-34a, in turn, targets mitogen-activated protein kinase kinase 1 (MEK1) and represses proliferation. 
Starczynowski et al51  miR-145 and miR-146a are mediators of the 5q-syndrome phenotype. 

Georgantas et al performed both mRNA and miRNA expression profiling on CD34+ HSCs from healthy subjects29  and used a bioinformatic approach to predict candidate miRNAs that target mRNAs encoding transcription factors associated with hematopoietic differentiation. miR-155 was predicted to repress expression of 9 different CD34+ mRNAs that regulate myelopoiesis. miR-155 expression was dramatically reduced when CD34+ HSCs were differentiated along the megakaryocyte lineage.29,46  Forced overexpression of miR-155 inhibited K562 (a chronic myelogenous leukemia cell line) differentiation and reduced CD34+ HSC-derived myeloid and erythroid colony formation in vitro.29  Transplantation of HSCs overexpressing miR-155 into irradiated mice caused reduced numbers of megakaryocytes in recipient bone marrow.40  Although the effect of miR-155 on platelet count has not been studied, these studies provide strong support that miR-155 inhibits megakaryocytopoiesis.

By profiling primary cells from human umbilical cord blood, Lu et al discovered miR-150 levels increased as megakaryocyte-erythrocyte progenitors (MEPs) differentiated toward the megakaryocyte lineage, but not the erythroid lineage.47  Overexpression of miR-150 enhanced both in vitro and in vivo megakaryocyte differentiation at the expense of erythroid differentiation, suggesting a critical switching function at the level of the MEP. miR-150 also knocked down expression of MYB via its 3′-UTR, consistent with data showing that low c-Myb levels promote megakaryocytopoiesis.48  These findings, coupled with work from the Kaushansky laboratory showing thrombopoietin up-regulates miR-150,49  underscore a critical role for miR-150 in promoting megakaryocytopoiesis.

miR-146a levels have been reported to dramatically change on megakaryocytic differentiation of HSCs, but there is conflicting evidence as to the direction of change.50-53  Opalinska et al52  reported that, in murine and human hematopoietic stem cells induced to differentiate into megakaryocytes, the expression level of miR-146a increased. However, forced expression of miR-146a had no effect on megakaryocyte colony forming units (CFU-MK), marker expression, or platelet activation.52  In contrast, Labbaye et al50  found that miR-146a decreased in human cord blood stem cells induced to differentiate into megakaryocytes and that forced expression caused a reduction in the number of polyploid cells. Conversely, inhibition of miR-146a by an antagomir caused an increase in the number of polyploid cells.50  Two recent reports by Starczynowski et al53  did not clarify the matter. They report that the level of miR-146a is lower in megakaryocyte/erythroid precursors relative to hematopoietic stem cells, similar to Labbaye et al,50  but that forced expression had no effect on platelet number, similar to Opalinska et al52  However, when down-regulated by decoy targets or anti-miRNA locked nucleic acids, CFU-MK and platelet number increase and the ploidy of the megakaryocytes decreases.51  These apparently conflicting findings may stem from species differences or differing experimental culture conditions (because miR-146a levels vary according to lineage). It is also possible that different experimental conditions result in functionally different levels of miR-146a. This is because Ago proteins preferentially associate with transcripts that contain targets for the highly expressed miRNAs, and mRNAs with a greater number of target sites in their 3′-UTR are subject to stricter control.54,55  Therefore, differing levels of miRNA expression can result in different mRNA levels and experimental results. Lastly, because single nucleotide polymorphisms in the 3′-UTR binding site as well as the surrounding sequence8  can significantly alter miRNA effects,23  it may be necessary to consider the precise mRNA target sequences in these different studies. Recent work by our group77  and by Landry et al56  profiling platelets both found miR-146a to be in the top quartile of expressed miRNAs (Figure 2). An interesting possibility is that the effect of miR-146a is not cell autologous. Opalinska et al found than the expression levels of tumor necrosis factor-α, interferon-β, and interleukin-1β were all repressed in macrophages overexpressing miR-146a.52  In addition, Starczynowski et al53  reported that miR-146a directly targets TRAF6, and down-regulating it results in increased TRAF6 and interleukin-6 levels. Interleukin-6 has been reported to have an effect on megakaryocyte development, and it is possible that differences in the reports are because of complications resulting from changes in cytokine signaling pathways.57 

Figure 2

Variation in levels of platelet miRNAs. Levels of individual human platelet miRNAs are shown, arbitrarily ordered from lowest to highest expression (unpublished data). As examples, the expression levels of miRNA-150, miRNA-155, miRNA-126, and miRNA-146a are highlighted in green. A total of 750 human miRNAs were profiled from leukocyte-depleted platelet RNAs from 19 human volunteers using the miRCURY LNA Array, Version 11.0 (Exiqon). Expression levels ranged over 4 orders on magnitude. The horizontal dashed line indicates background level.

Figure 2

Variation in levels of platelet miRNAs. Levels of individual human platelet miRNAs are shown, arbitrarily ordered from lowest to highest expression (unpublished data). As examples, the expression levels of miRNA-150, miRNA-155, miRNA-126, and miRNA-146a are highlighted in green. A total of 750 human miRNAs were profiled from leukocyte-depleted platelet RNAs from 19 human volunteers using the miRCURY LNA Array, Version 11.0 (Exiqon). Expression levels ranged over 4 orders on magnitude. The horizontal dashed line indicates background level.

Close modal

Increased levels of miR-34a have consistently been observed in K562 cells stimulated to differentiate with phorbol myristate acetate.58,59  When overexpressed in K562 cells, miR-34a inhibited K562 cell proliferation and promoted differentiation. Importantly, miR-34a increased megakaryocyte colony formation from CD34+ HSCs.59  Thus, miR-34a appears to enhance megakaryocytopoiesis.

Several other miRNAs may regulate megakaryocytopoiesis, although these miRNAs have been less well studied. Several investigators have observed that more miRNAs were down-regulated than up-regulated during megakaryocytopoiesis, but biphasic patterns of expression have also been observed.31,45,52  Several studies have linked reciprocal associations between miRNA levels and target mRNA levels, a particularly interesting finding when the mRNA encodes a transcription factor known to be involved in hematopoiesis. Examples include miR-130a and MAFB, miR-10a and HOXA1, and miR-27a and Runx1.45,60  Forced expression of the oncomiR miR-125b-2 increased proliferation and self-renewal of MEP and megakaryocyte progenitors,61  and miR-28 appears to exert a negative effect on megakaryocyte differentiation.62  Lastly, both decreasing miR-181 and increasing miR-27a levels have been associated with megakaryocytic differentiation.60,63  In the case of megakaryocytic cell lines, such associations should be viewed cautiously because the correlation between miRNA expression in primary megakaryocytes and cell lines may not always be strong.64 Figure 3 summarizes the studies of miRNAs in megakaryocytopoiesis.

Figure 3

miRNAs and megakaryocytopoiesis. This diagram summarizes the miRNAs known to participate in megakaryocytopoiesis. Megakaryocytes (MK) originate from self-renewing HSC, which differentiate progressively into common myeloid precursors (CMP), MEPs, and megakaryocyte precursors (MP). The height of the triangles indicates the level of the miRNA, with some miRNAs decreasing during megakaryocytopoiesis, and others increasing. There is conflicting evidence regarding the role of miR-146a. Levels of miR-125b-2 do not change during progression from MEP to MP.

Figure 3

miRNAs and megakaryocytopoiesis. This diagram summarizes the miRNAs known to participate in megakaryocytopoiesis. Megakaryocytes (MK) originate from self-renewing HSC, which differentiate progressively into common myeloid precursors (CMP), MEPs, and megakaryocyte precursors (MP). The height of the triangles indicates the level of the miRNA, with some miRNAs decreasing during megakaryocytopoiesis, and others increasing. There is conflicting evidence regarding the role of miR-146a. Levels of miR-125b-2 do not change during progression from MEP to MP.

Close modal

Platelets contain miRNAs,31,56,65-68  and we and others have used genome-wide profiling to demonstrate normal human platelets express high levels of miRNA (Figure 2).56  Microarray profile levels are readily validated using a novel stem-loop reverse-transcribed polymerase chain reaction that allows quantification of specific miRNAs.12  Emerging evidence over the past year suggests that platelet miRNAs are biologically and clinically relevant as (1) potential regulators of platelet protein translation and expression, (2) markers of mature megakaryocyte miRNA levels, (3) biomarkers for hematologic disease and platelet reactivity, and (4) a tool for understanding basic mechanisms of megakaryocyte/platelet gene expression.

It is well known that platelets have mRNA and mRNA splicing machinery, and translate mRNA into proteins relevant to hemostasis and inflammation.69-71  Notably, platelets stored in blood banks increase synthesis of integrin-β3.72  Work from the Provost laboratory has demonstrated that human platelets also contain miRNA processing machinery, including Dicer, TAR RNA-binding protein 2, and Ago2, and that platelets are able to process pre-miRNA into mature miRNA.56  Our in silico analysis indicates that each platelet miRNA targets an average of 307 distinct mRNAs (range, 12-1417; unpublished data), consistent with prior predictions.73,77  Thus, platelet miRNAs have ample opportunity to regulate platelet function, although direct evidence has yet to be reported.

As described in “miRNAs and megakaryocytopoiesis,” functional miRNA levels change during megakaryocyte differentiation of cultured CD34+ HSCs. It is not difficult to imagine similar changes occurring in vivo, but this has not been formally studied. In vitro culture conditions lack numerous in vivo factors that could alter miRNA levels, including other marrow niche cells, innervation, plasma components, spatial constraints, and environmental factors. Hussein et al attempted to directly assess in vivo megakaryocyte miRNA levels using laser microdissection to isolate mature megakaryocytes from patient bone marrow biopsies.74  The extent to which platelet and mature megakaryocyte miRNA profiles correlate is unknown. As an initial attempt to address this issue, we compared platelet miRNA profiles from 19 healthy subjects with the megakaryocyte miRNA profiles from the 2 patients reported by Hussein et al.74  As shown in Figure 4A, there was a significant correlation between the platelet and megakaryocyte miRNAs. We suspect that the true correlation is even stronger because different profiling platforms and RNA preparations were used and only 2 megakaryocyte samples were assessed. Furthermore, the megakaryocyte samples were from patients, whereas the platelets were from healthy donors. As expected, when we compared our dataset with the platelet miRNAs reported by Landry et al,56  we found a greater degree of correlation (Figure 4B).

Figure 4

Correlation between human platelet and megakaryocyte miRNAs. miRNA expression levels were determined in platelets from 19 healthy donors by microarray (our unpublished data) and in patient megakaryocytes by quantitative reverse-transcribed polymerase chain reaction.74  The miRNAs that were queried in both studies were rank ordered and plotted. Statistical significance was calculated by Pearson rank correlation. Dotted line represents 1:1 ratio of rank. (A) Platelet miRNAs correlated with megakaryocyte miRNAs. (B) Platelet miRNAs from the author's laboratory correlated with platelet miRNAs from the Provost laboratory.56 

Figure 4

Correlation between human platelet and megakaryocyte miRNAs. miRNA expression levels were determined in platelets from 19 healthy donors by microarray (our unpublished data) and in patient megakaryocytes by quantitative reverse-transcribed polymerase chain reaction.74  The miRNAs that were queried in both studies were rank ordered and plotted. Statistical significance was calculated by Pearson rank correlation. Dotted line represents 1:1 ratio of rank. (A) Platelet miRNAs correlated with megakaryocyte miRNAs. (B) Platelet miRNAs from the author's laboratory correlated with platelet miRNAs from the Provost laboratory.56 

Close modal

miRNAs are very stable and, compared with mRNAs, have superior performance characteristics as biomarkers for disease activity.47,75,76  The first human platelet miRNA profiling study was performed in 2008 by Bruchova et al as part of a study testing for differentially expressed miRNAs in patients with polycythemia vera.65  These investigators found that miR-26b was significantly higher in polycythemia vera platelets than in platelets from healthy donors. Landry et al described how miR-223 could target the mRNA of the adenosine 5′-diphosphate receptor, P2Y12, and that P2Y12 mRNA was found in Ago2 immunoprecipitates in both megakaryocytes and platelets.56  Our group has investigated associations between miRNAs and platelet reactivity, both to better understand megakaryocyte/platelet gene expression and to identify potential biomarkers for thrombotic risk. Because a single nucleotide polymorphism in the VAMP8 3′-UTR has been associated with coronary artery disease and because VAMP8 mRNA is differentially expressed between platelets of differing reactivity,67  we considered whether a miRNA might affect VAMP8 expression. Indeed, miR-96 was shown to knock down VAMP8 mRNA and protein; and in a small number of subjects, miR-96 was differentially expressed between platelets of differing reactivity in a manner that was consistent with its effect on VAMP8 expression.67  Recent data from our laboratory demonstrate the use of differentially expressed mRNA-miRNA pairs for identifying functional miRNAs, as assessed by the ability of the miRNA to target and knock down the mRNA was confirmed in cell culture.77 

In conclusion, miRNAs have an established role in hematopoiesis and megakaryocytopoiesis, and platelet miRNAs have potential as tools for understanding basic mechanisms of megakaryocyte/platelet gene expression. Genome-wide association studies have identified orphan loci not in or near protein-coding genes, and at least one such variant proved to be caused by a mutation in a miRNA gene.78,79  It is probable that additional megakaryocyte/platelet disease-producing genetic variants in miRNA biogenesis will be identified. Very exciting opportunities exist for future translational research involving miRNAs in megakaryocytopoiesis and platelet biology. Environmental stresses and aging can affect miRNA levels,80,81  and it will be important to test whether these factors influence miRNA-mediated megakaryocyte/platelet physiology. Gene therapy approaches are using tissue-specific or developmental stage-specific miRNA expression to avoid off-target effects.82  For example, Gentner et al have incorporated miRNA target sequences into an expression vector that results in suppressed ectopic gene expression in HSCs but enhanced expression in mature hematopoietic cells.83  miRNAs can be released via shed microvesicles,84  raising the possibility of platelet-mediated delivery of miRNAs to targeted vascular sites. Platelets have been engineered to ectopically deliver recombinant factor VIII and reduce bleeding85 ; perhaps P-selectin bearing platelet microvesicles could target miRNAs to PSGL-1 expressing leukocytes at sites of inflammation. Other potential directions include manipulating platelet miRNAs to modify platelet life span or the platelet storage defect in banked platelets.

Contribution: L.C.E. and P.F.B. analyzed and interpreted data and wrote the manuscript.

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

Correspondence: Paul F. Bray, Thomas Jefferson University, Cardeza Foundation for Hematologic Research and the Department of Medicine, Jefferson Medical College, 1015 Walnut St, Curtis Bldg, Rm 324, Philadelphia, PA 19107; e-mail: paul.bray@jefferson.edu.

1
Gregory
 
TR
Synergy between sequence and size in large-scale genomics.
Nat Rev Genet
2005
, vol. 
6
 
9
(pg. 
699
-
708
)
2
Shabalina
 
SA
Spiridonov
 
NA
The mammalian transcriptome and the function of non-coding DNA sequences.
Genome Biol
2004
, vol. 
5
 
4
pg. 
105
 
3
Aravin
 
AA
Hannon
 
GJ
Brennecke
 
J
The Piwi-piRNA pathway provides an adaptive defense in the transposon arms race.
Science
2007
, vol. 
318
 
5851
(pg. 
761
-
764
)
4
Mattick
 
JS
Makunin
 
IV
Non-coding RNA.
Hum Mol Genet
2006
, vol. 
15
 (pg. 
R17
-
R29
Spec 1
5
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
, vol. 
75
 
5
(pg. 
855
-
862
)
6
Lee
 
RC
Feinbaum
 
RL
Ambros
 
V
The C. elegans heterochronic gene lin-4 encodes small RNAs with antisense complementarity to lin-14.
Cell
1993
, vol. 
75
 
5
(pg. 
843
-
854
)
7
Bartel
 
DP
MicroRNAs: genomics, biogenesis, mechanism, and function.
Cell
2004
, vol. 
116
 
2
(pg. 
281
-
297
)
8
Bartel
 
DP
MicroRNAs: target recognition and regulatory functions.
Cell
2009
, vol. 
136
 
2
(pg. 
215
-
233
)
9
Griffiths-Jones
 
S
Saini
 
HK
van Dongen
 
S
Enright
 
AJ
miRBase: tools for microRNA genomics.
Nucleic Acids Res
2008
, vol. 
36
 (pg. 
D154
-
D158
(Database issue)
10
Friedman
 
RC
Farh
 
KK
Burge
 
CB
Bartel
 
DP
Most mammalian mRNAs are conserved targets of microRNAs.
Genome Res
2009
, vol. 
19
 
1
(pg. 
92
-
105
)
11
Wienholds
 
E
Kloosterman
 
WP
Miska
 
E
, et al. 
MicroRNA expression in zebrafish embryonic development.
Science
2005
, vol. 
309
 
5732
(pg. 
310
-
311
)
12
Chen
 
C
Ridzon
 
DA
Broomer
 
AJ
, et al. 
Real-time quantification of microRNAs by stem-loop RT-PCR.
Nucleic Acids Res
2005
, vol. 
33
 
20
pg. 
e179
 
13
Borchert
 
GM
Lanier
 
W
Davidson
 
BL
RNA polymerase III transcribes human microRNAs.
Nat Struct Mol Biol
2006
, vol. 
13
 
12
(pg. 
1097
-
1101
)
14
Cai
 
X
Hagedorn
 
CH
Cullen
 
BR
Human microRNAs are processed from capped, polyadenylated transcripts that can also function as mRNAs.
RNA
2004
, vol. 
10
 
12
(pg. 
1957
-
1966
)
15
Lee
 
Y
Kim
 
M
Han
 
J
, et al. 
MicroRNA genes are transcribed by RNA polymerase II.
EMBO J
2004
, vol. 
23
 
20
(pg. 
4051
-
4060
)
16
Cheloufi
 
S
Dos Santos
 
CO
Chong
 
MM
Hannon
 
GJ
A dicer-independent miRNA biogenesis pathway that requires Ago catalysis.
Nature
2010
, vol. 
465
 
7298
(pg. 
584
-
589
)
17
Cifuentes
 
D
Xue
 
H
Taylor
 
DW
, et al. 
A novel miRNA processing pathway independent of Dicer requires Argonaute2 catalytic activity.
Science
2010
, vol. 
328
 
5986
(pg. 
1694
-
1698
)
18
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
, vol. 
455
 
7216
(pg. 
1124
-
1128
)
19
Guo
 
H
Ingolia
 
NT
Weissman
 
JS
Bartel
 
DP
Mammalian microRNAs predominantly act to decrease target mRNA levels.
Nature
2010
, vol. 
466
 
7308
(pg. 
835
-
840
)
20
Selbach
 
M
Schwanhausser
 
B
Thierfelder
 
N
Fang
 
Z
Khanin
 
R
Rajewsky
 
N
Widespread changes in protein synthesis induced by microRNAs.
Nature
2008
, vol. 
455
 
7209
(pg. 
58
-
63
)
21
Baek
 
D
Villen
 
J
Shin
 
C
Camargo
 
FD
Gygi
 
SP
Bartel
 
DP
The impact of microRNAs on protein output.
Nature
2008
, vol. 
455
 
7209
(pg. 
64
-
71
)
22
Bandiera
 
S
Hatem
 
E
Lyonnet
 
S
Henrion-Caude
 
A
microRNAs in diseases: from candidate to modifier genes.
Clin Genet
2010
, vol. 
77
 
4
(pg. 
306
-
313
)
23
Sethupathy
 
P
Collins
 
FS
MicroRNA target site polymorphisms and human disease.
Trends Genet
2008
, vol. 
24
 
10
(pg. 
489
-
497
)
24
Poliseno
 
L
Salmena
 
L
Zhang
 
J
Carver
 
B
Haveman
 
WJ
Pandolfi
 
PP
A coding-independent function of gene and pseudogene mRNAs regulates tumour biology.
Nature
2010
, vol. 
465
 
7301
(pg. 
1033
-
1038
)
25
Alimonti
 
A
Carracedo
 
A
Clohessy
 
JG
, et al. 
Subtle variations in Pten dose determine cancer susceptibility.
Nat Genet
2010
, vol. 
42
 
5
(pg. 
454
-
458
)
26
Orkin
 
SH
Zon
 
LI
SnapShot: hematopoiesis.
Cell
2008
, vol. 
132
 
4
pg. 
712
 
27
Raaijmakers
 
MHGP
Mukherjee
 
S
Guo
 
S
, et al. 
Bone progenitor dysfunction induces myelodysplasia and secondary leukaemia.
Nature
2010
, vol. 
464
 
7290
(pg. 
852
-
857
)
28
Chen
 
C-Z
Li
 
L
Lodish
 
HF
Bartel
 
DP
MicroRNAs modulate hematopoietic lineage differentiation.
Science
2004
, vol. 
303
 
5654
(pg. 
83
-
86
)
29
Georgantas
 
RW
Hildreth
 
R
Morisot
 
S
, et al. 
CD34+ hematopoietic stem-progenitor cell microRNA expression and function: a circuit diagram of differentiation control.
Proc Natl Acad Sci U S A
2007
, vol. 
20;104
 
8
(pg. 
2750
-
2755
)
30
Monticelli
 
S
Ansel
 
KM
Xiao
 
C
, et al. 
MicroRNA profiling of the murine hematopoietic system.
Genome Biol
2005
, vol. 
6
 
8
pg. 
R71
 
31
Bruchova
 
H
Yoon
 
D
Agarwal
 
AM
Mendell
 
J
Prchal
 
JT
Regulated expression of microRNAs in normal and polycythemia vera erythropoiesis.
Exp Hematol
2007
, vol. 
35
 
11
(pg. 
1657
-
1667
)
32
Choong
 
ML
Yang
 
HH
McNiece
 
I
MicroRNA expression profiling during human cord blood-derived CD34 cell erythropoiesis.
Exp Hematol
2007
, vol. 
35
 
4
(pg. 
551
-
564
)
33
Dore
 
LC
Amigo
 
JD
Dos Santos
 
CO
, et al. 
A GATA-1-regulated microRNA locus essential for erythropoiesis.
Proc Natl Acad Sci U S A
2008
, vol. 
105
 
9
(pg. 
3333
-
3338
)
34
Felli
 
N
Fontana
 
L
Pelosi
 
E
, et al. 
MicroRNAs 221 and 222 inhibit normal erythropoiesis and erythroleukemic cell growth via kit receptor down-modulation.
Proc Natl Acad Sci U S A
2005
, vol. 
102
 
50
(pg. 
18081
-
18086
)
35
Pase
 
L
Layton
 
JE
Kloosterman
 
WP
Carradice
 
D
Waterhouse
 
PM
Lieschke
 
GJ
miR-451 regulates zebrafish erythroid maturation in vivo via its target gata2.
Blood
2009
, vol. 
113
 
8
(pg. 
1794
-
1804
)
36
Zhao
 
H
Kalota
 
A
Jin
 
S
Gewirtz
 
AM
The c-myb proto-oncogene and microRNA-15a comprise an active autoregulatory feedback loop in human hematopoietic cells.
Blood
2009
, vol. 
113
 
3
(pg. 
505
-
516
)
37
Fazi
 
F
Rosa
 
A
Fatica
 
A
, et al. 
A minicircuitry comprised of microRNA-223 and transcription factors NFI-A and C/EBPalpha regulates human granulopoiesis.
Cell
2005
, vol. 
123
 
5
(pg. 
819
-
831
)
38
Fontana
 
L
Pelosi
 
E
Greco
 
P
, et al. 
MicroRNAs 17-5p–20a-106a control monocytopoiesis through AML1 targeting and M-CSF receptor upregulation.
Nat Cell Biol
2007
, vol. 
9
 
7
(pg. 
775
-
787
)
39
Johnnidis
 
JB
Harris
 
MH
Wheeler
 
RT
, et al. 
Regulation of progenitor cell proliferation and granulocyte function by microRNA-223.
Nature
2008
, vol. 
451
 
7182
(pg. 
1125
-
1129
)
40
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
, vol. 
205
 
3
(pg. 
585
-
594
)
41
Neilson
 
JR
Zheng
 
GXY
Burge
 
CB
Sharp
 
PA
Dynamic regulation of miRNA expression in ordered stages of cellular development.
Genes Dev
2007
, vol. 
21
 
5
(pg. 
578
-
589
)
42
Ventura
 
A
Young
 
AG
Winslow
 
MM
, et al. 
Targeted deletion reveals essential and overlapping functions of the miR-17 through 92 family of miRNA clusters.
Cell
2008
, vol. 
132
 
5
(pg. 
875
-
886
)
43
Xiao
 
C
Calado
 
DP
Galler
 
G
, et al. 
miR-150 controls B cell differentiation by targeting the transcription factor c-Myb.
Cell
2007
, vol. 
131
 
1
(pg. 
146
-
159
)
44
Zhou
 
B
Wang
 
S
Mayr
 
C
Bartel
 
DP
Lodish
 
HF
miR-150, a microRNA expressed in mature B and T cells, blocks early B cell development when expressed prematurely.
Proc Natl Acad Sci U S A
2007
, vol. 
104
 
17
(pg. 
7080
-
7085
)
45
Garzon
 
R
Pichiorri
 
F
Palumbo
 
T
, et al. 
MicroRNA fingerprints during human megakaryocytopoiesis.
Proc Natl Acad Sci U S A
2006
, vol. 
103
 
13
(pg. 
5078
-
5083
)
46
Romania
 
P
Lulli
 
V
Pelosi
 
E
Biffoni
 
M
Peschle
 
C
Marziali
 
G
MicroRNA 155 modulates megakaryopoiesis at progenitor and precursor level by targeting Ets-1 and Meis1 transcription factors.
Br J Haematol
2008
, vol. 
143
 
4
(pg. 
570
-
580
)
47
Lu
 
J
Guo
 
S
Ebert
 
BL
, et al. 
MicroRNA-mediated control of cell fate in megakaryocyte-erythrocyte progenitors.
Dev Cell
2008
, vol. 
14
 
6
(pg. 
843
-
853
)
48
Emambokus
 
N
Vegiopoulos
 
A
Harman
 
B
Jenkinson
 
E
Anderson
 
G
Frampton
 
J
Progression through key stages of haemopoiesis is dependent on distinct threshold levels of c-Myb.
EMBO J
2003
, vol. 
22
 
17
(pg. 
4478
-
4488
)
49
Barroga
 
CF
Pham
 
H
Kaushansky
 
K
Thrombopoietin regulates c-Myb expression by modulating micro RNA 150 expression.
Exp Hematol
2008
, vol. 
36
 
12
(pg. 
1585
-
1592
)
50
Labbaye
 
C
Spinello
 
I
Quaranta
 
MT
, et al. 
A three-step pathway comprising PLZF/miR-146a/CXCR4 controls megakaryopoiesis.
Nat Cell Biol
2008
, vol. 
10
 
7
(pg. 
788
-
801
)
51
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
, vol. 
16
 
1
(pg. 
49
-
58
)
52
Opalinska
 
JB
Bersenev
 
A
Zhang
 
Z
, et al. 
MicroRNA expression in maturing murine megakaryocytes.
Blood
2010
, vol. 
116
 
23
(pg. 
e128
-
e138
)
53
Starczynowski
 
DT
Kuchenbauer
 
F
Wegrzyn
 
J
, et al. 
MicroRNA-146a disrupts hematopoietic differentiation and survival.
Exp Hematol
2011
, vol. 
39
 
2
(pg. 
167
-
178
)
54
Landthaler
 
M
Gaidatzis
 
D
Rothballer
 
A
, et al. 
Molecular characterization of human Argonaute-containing ribonucleoprotein complexes and their bound target mRNAs.
RNA
2008
, vol. 
14
 
12
(pg. 
2580
-
2596
)
55
Schmitter
 
D
Filkowski
 
J
Sewer
 
A
, et al. 
Effects of Dicer and Argonaute down-regulation on mRNA levels in human HEK293 cells.
Nucleic Acids Res
2006
, vol. 
34
 
17
(pg. 
4801
-
4815
)
56
Landry
 
P
Plante
 
I
Ouellet
 
DL
Perron
 
MP
Rousseau
 
G
Provost
 
P
Existence of a microRNA pathway in anucleate platelets.
Nat Struct Mol Biol
2009
, vol. 
16
 
9
(pg. 
961
-
966
)
57
Ishibashi
 
T
Kimura
 
H
Uchida
 
T
Kariyone
 
S
Friese
 
P
Burstein
 
SA
Human interleukin 6 is a direct promoter of maturation of megakaryocytes in vitro.
Proc Natl Acad Sci U S A
1989
, vol. 
86
 
15
(pg. 
5953
-
5957
)
58
Ichimura
 
A
Ruike
 
Y
Terasawa
 
K
Shimizu
 
K
Tsujimoto
 
G
MicroRNA-34a inhibits cell proliferation by repressing mitogen-activated protein kinase kinase 1 during megakaryocytic differentiation of K562 cells.
Mol Pharmacol
2010
, vol. 
77
 
6
(pg. 
1016
-
1024
)
59
Navarro
 
F
Gutman
 
D
Meire
 
E
, et al. 
miR-34a contributes to megakaryocytic differentiation of K562 cells independently of p53.
Blood
2009
, vol. 
114
 
10
(pg. 
2181
-
2192
)
60
Ben-Ami
 
O
Pencovich
 
N
Lotem
 
J
Levanon
 
D
Groner
 
Y
A regulatory interplay between miR-27a and Runx1 during megakaryopoiesis.
Proc Natl Acad Sci U S A
2009
, vol. 
106
 
1
(pg. 
238
-
243
)
61
Klusmann
 
JH
Li
 
Z
Bohmer
 
K
, et al. 
miR-125b-2 is a potential oncomiR on human chromosome 21 in megakaryoblastic leukemia.
Genes Dev
2010
, vol. 
24
 
5
(pg. 
478
-
490
)
62
Girardot
 
M
Pecquet
 
C
Boukour
 
S
, et al. 
miR-28 is a thrombopoietin receptor targeting microRNA detected in a fraction of myeloproliferative neoplasm patient platelets.
Blood
2010
, vol. 
116
 
3
(pg. 
437
-
445
)
63
Guimaraes-Sternberg
 
C
Meerson
 
A
Shaked
 
I
Soreq
 
H
MicroRNA modulation of megakaryoblast fate involves cholinergic signaling.
Leuk Res
2006
, vol. 
30
 
5
(pg. 
583
-
595
)
64
Ramkissoon
 
SH
Mainwaring
 
LA
Ogasawara
 
Y
, et al. 
Hematopoietic-specific microRNA expression in human cells.
Leukemia Res
2006
, vol. 
30
 
5
(pg. 
643
-
647
)
65
Bruchova
 
H
Merkerova
 
M
Prchal
 
JT
Aberrant expression of microRNA in polycythemia vera.
Haematologica
2008
, vol. 
93
 
7
(pg. 
1009
-
1016
)
66
Kannan
 
M
Mohan
 
KV
Kulkarni
 
S
Atreya
 
C
Membrane array-based differential profiling of platelets during storage for 52 miRNAs associated with apoptosis.
Transfusion
2009
, vol. 
49
 
7
(pg. 
1443
-
1450
)
67
Kondkar
 
AA
Bray
 
MS
Leal
 
SM
, et al. 
VAMP8/endobrevin is overexpressed in hyperreactive human platelets: suggested role for platelet microRNA.
J Thromb Haemost
2010
, vol. 
8
 
2
(pg. 
369
-
378
)
68
Merkerova
 
M
Belickova
 
M
Bruchova
 
H
Differential expression of microRNAs in hematopoietic cell lineages.
Eur J Haematol
2008
, vol. 
81
 
4
(pg. 
304
-
310
)
69
Warshaw
 
AL
Laster
 
L
Shulman
 
NR
The stimulation by thrombin of glucose oxidation in human platelets.
J Clin Invest
1966
, vol. 
45
 
12
(pg. 
1923
-
1934
)
70
Weyrich
 
AS
Schwertz
 
H
Kraiss
 
LW
Zimmerman
 
GA
Protein synthesis by platelets: historical and new perspectives.
J Thromb Haemost
2009
, vol. 
7
 
2
(pg. 
241
-
246
)
71
Denis
 
MM
Tolley
 
ND
Bunting
 
M
, et al. 
Escaping the nuclear confines: signal-dependent pre-mRNA splicing in anucleate platelets.
Cell
2005
, vol. 
122
 
3
(pg. 
379
-
391
)
72
Thon
 
JN
Devine
 
DV
Translation of glycoprotein IIIa in stored blood platelets.
Transfusion
2007
, vol. 
47
 
12
(pg. 
2260
-
2270
)
73
Miranda
 
KC
Huynh
 
T
Tay
 
Y
, et al. 
A pattern-based method for the identification of MicroRNA binding sites and their corresponding heteroduplexes.
Cell
2006
, vol. 
126
 
6
(pg. 
1203
-
1217
)
74
Hussein
 
K
Theophile
 
K
Dralle
 
W
Wiese
 
B
Kreipe
 
H
Bock
 
O
MicroRNA expression profiling of megakaryocytes in primary myelofibrosis and essential thrombocythemia.
Platelets
2009
, vol. 
20
 
6
(pg. 
391
-
400
)
75
Kai
 
ZS
Pasquinelli
 
AE
MicroRNA assassins: factors that regulate the disappearance of miRNAs.
Nat Struct Mol Biol
2010
, vol. 
17
 
1
(pg. 
5
-
10
)
76
Scholer
 
N
Langer
 
C
Dohner
 
H
Buske
 
C
Kuchenbauer
 
F
Serum microRNAs as a novel class of biomarkers: a comprehensive review of the literature.
Exp Hematol
2010
, vol. 
38
 
12
(pg. 
1126
-
1130
)
77
Nagalla
 
S
Shaw
 
C
Kong
 
X
, et al. 
Platelet microRNA-mRNA coexpression profiles correlate with platelet reactivity [published online ahead of print March 17, 2011].
Blood
 
78
Mencia
 
A
Modamio-Hoybjor
 
S
Redshaw
 
N
, et al. 
Mutations in the seed region of human miR-96 are responsible for nonsyndromic progressive hearing loss.
Nat Genet
2009
, vol. 
41
 
5
(pg. 
609
-
613
)
79
Modamio-Hoybjor
 
S
Moreno-Pelayo
 
MA
Mencia
 
A
, et al. 
A novel locus for autosomal dominant nonsyndromic hearing loss, DFNA50, maps to chromosome 7q32 between the DFNB17 and DFNB13 deafness loci.
J Med Genet
2004
, vol. 
41
 
2
pg. 
e14
 
80
Bhattacharyya
 
SN
Habermacher
 
R
Martine
 
U
Closs
 
EI
Filipowicz
 
W
Relief of microRNA-mediated translational repression in human cells subjected to stress.
Cell
2006
, vol. 
125
 
6
(pg. 
1111
-
1124
)
81
Ibanez-Ventoso
 
C
Yang
 
M
Guo
 
S
Robins
 
H
Padgett
 
RW
Driscoll
 
M
Modulated microRNA expression during adult lifespan in Caenorhabditis elegans.
Aging Cell
2006
, vol. 
5
 
3
(pg. 
235
-
246
)
82
Brown
 
BD
Naldini
 
L
Exploiting and antagonizing microRNA regulation for therapeutic and experimental applications.
Nat Rev Genet
2009
, vol. 
10
 
8
(pg. 
578
-
585
)
83
Gentner
 
B
Visigalli
 
I
Hiramatsu
 
H
, et al. 
Identification of hematopoietic stem cell-specific miRNAs enables gene therapy of globoid cell leukodystrophy.
Sci Transl Med
2010
, vol. 
2
 
58
(pg. 
58ra
-
84
)
84
Yuan
 
A
Farber
 
EL
Rapoport
 
AL
, et al. 
Transfer of microRNAs by embryonic stem cell microvesicles.
PLoS ONE
2009
, vol. 
4
 
3
pg. 
e4722
 
85
Greene
 
TK
Wang
 
C
Hirsch
 
JD
, et al. 
In vivo efficacy of platelet-delivered, high specific activity factor VIII variants.
Blood
2010
, vol. 
116
 
26
(pg. 
6114
-
6122
)
86
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
, vol. 
113
 
25
(pg. 
6411
-
6418
)
87
Garzon
 
R
Heaphy
 
CE
Havelange
 
V
, et al. 
MicroRNA 29b functions in acute myeloid leukemia.
Blood
2009
, vol. 
114
 
26
(pg. 
5331
-
5341
)
88
Debernardi
 
S
Skoulakis
 
S
Molloy
 
G
Chaplin
 
T
Dixon-McIver
 
A
Young
 
BD
MicroRNA miR-181a correlates with morphological sub-class of acute myeloid leukaemia and the expression of its target genes in global genome-wide analysis.
Leukemia
2007
, vol. 
21
 
5
(pg. 
912
-
916
)
89
Garzon
 
R
Garofalo
 
M
Martelli
 
MP
, et al. 
Distinctive microRNA signature of acute myeloid leukemia bearing cytoplasmic mutated nucleophosmin.
Proc Natl Acad Sci U S A
2008
, vol. 
105
 
10
(pg. 
3945
-
3950
)
90
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
, vol. 
140
 
5
(pg. 
652
-
665
)
91
Marton
 
S
Garcia
 
MR
Robello
 
C
, et al. 
Small RNAs analysis in CLL reveals a deregulation of miRNA expression and novel miRNA candidates of putative relevance in CLL pathogenesis.
Leukemia
2008
, vol. 
22
 
2
(pg. 
330
-
338
)
92
Visone
 
R
Rassenti
 
LZ
Veronese
 
A
, et al. 
Karyotype-specific microRNA signature in chronic lymphocytic leukemia.
Blood
2009
, vol. 
114
 
18
(pg. 
3872
-
3879
)
93
Xu
 
W
Li
 
JY
MicroRNA gene expression in malignant lymphoproliferative disorders.
Chin Med J (Engl)
2007
, vol. 
120
 
11
(pg. 
996
-
999
)
94
Fulci
 
V
Chiaretti
 
S
Goldoni
 
M
, et al. 
Quantitative technologies establish a novel microRNA profile of chronic lymphocytic leukemia.
Blood
2007
, vol. 
109
 
11
(pg. 
4944
-
4951
)
95
Bottoni
 
A
Piccin
 
D
Tagliati
 
F
Luchin
 
A
Zatelli
 
MC
degli Uberti
 
EC
miR-15a and miR-16-1 down-regulation in pituitary adenomas.
J Cell Physiol
2005
, vol. 
204
 
1
(pg. 
280
-
285
)
96
Calin
 
GA
Sevignani
 
C
Dumitru
 
CD
, et al. 
Human microRNA genes are frequently located at fragile sites and genomic regions involved in cancers.
Proc Natl Acad Sci U S A
2004
, vol. 
101
 
9
(pg. 
2999
-
3004
)
97
Lawrie
 
CH
Soneji
 
S
Marafioti
 
T
, et al. 
MicroRNA expression distinguishes between germinal center B cell-like and activated B cell-like subtypes of diffuse large B cell lymphoma.
Int J Cancer
2007
, vol. 
121
 
5
(pg. 
1156
-
1161
)
98
Eis
 
PS
Tam
 
W
Sun
 
L
, et al. 
Accumulation of miR-155 and BIC RNA in human B cell lymphomas.
Proc Natl Acad Sci U S A
2005
, vol. 
102
 
10
(pg. 
3627
-
3632
)
99
He
 
L
Thomson
 
JM
Hemann
 
MT
, et al. 
A microRNA polycistron as a potential human oncogene.
Nature
2005
, vol. 
435
 
7043
(pg. 
828
-
833
)

National Institutes of Health

Sign in via your Institution