Next Article in Journal
Clinical Applications of Autophagy Proteins in Cancer: From Potential Targets to Biomarkers
Next Article in Special Issue
miR-206-3p Inhibits 3T3-L1 Cell Adipogenesis via the c-Met/PI3K/Akt Pathway
Previous Article in Journal
Rapamycin Maintains the Chondrocytic Phenotype and Interferes with Inflammatory Cytokine Induced Processes
Previous Article in Special Issue
MicroRNA Signature of Human Microvascular Endothelium Infected with Rickettsia rickettsii
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

miRNAs in Normal and Malignant Hematopoiesis

Department of Hematology and Oncology, Tokai University School of Medicine, Hiratsuka 259-1193, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2017, 18(7), 1495; https://doi.org/10.3390/ijms18071495
Submission received: 1 May 2017 / Revised: 21 June 2017 / Accepted: 25 June 2017 / Published: 11 July 2017
(This article belongs to the Special Issue microRNA Regulation 2017)

Abstract

:
Lineage specification is primarily regulated at the transcriptional level and lineage-specific transcription factors determine cell fates. MicroRNAs (miRNAs) are 18–24 nucleotide-long non-coding RNAs that post-transcriptionally decrease the translation of target mRNAs and are essential for many cellular functions. miRNAs also regulate lineage specification during hematopoiesis. This review highlights the roles of miRNAs in B-cell development and malignancies, and discusses how miRNA expression profiles correlate with disease prognoses and phenotypes. We also discuss the potential for miRNAs as therapeutic targets and diagnostic tools for B-cell malignancies.

1. Introduction

Blood is composed of cells from multiple lineages, including myelocytes, lymphocytes, monocytes, platelets, and erythrocytes. Such cells play critical roles in multiple physiological functions, including generating immune responses, supplying oxygen to tissues, and maintaining hemostasis. Blood cells are continually generated from hematopoietic stem cells (HSCs) through a multi-step cell differentiation process called hematopoiesis. Each cell type has been characterized at specific developmental stages and the mechanisms of lineage specification have been extensively investigated during hematopoiesis. Notably, each step of hematopoiesis is regulated by a highly integrated network of transcription factors and microRNAs (miRNAs) [1,2,3].
In this review, we discuss the regulation of hematopoiesis by miRNAs, with a particular focus on B-cell differentiation. We also discuss the association between dysregulation of hematopoiesis and different blood cancers. The involvement of miRNAs in hematopoiesis might be used as a therapeutic alternative to treat blood cancers.

2. Hematopoiesis and B-Cell Development

Blood cells differentiate from HSCs through multiple stages, including multi-potent progenitors (MPPs), and restricted or committed progenitors, the fates of which are restricted at each step of hematopoiesis. A classical model of hematopoiesis describes the well-ordered step-wise lineage restrictions, including the differentiation of MPPs into lymphoid-restricted common lymphoid progenitors (CLPs) or myeloid and erythroid-restricted common myeloid and erythroid progenitors. In this context, only CLPs differentiate into B cells or T cells but not into myeloid or erythroid cell types, and vice versa.
Studies by Katsura’s group revealed a complex process in hematopoietic lineage restriction [4,5,6] in which progenitor cells differentiated into T myeloid cells or B myeloid cells at the single cell level, while no restricted T or B progenitor cells were apparent. Based on that data, a myeloid-based model for hematopoiesis was proposed [7]. Other groups also identified T-, B-, and myeloid-restricted progenitor cells, which is inconsistent with the classical model [8,9,10,11].
Previous studies revealed a complex network of transcription factors involved in the differentiation of lineages [1,2,3]. In B-cell development, PU.1, E2A, EBF1, and Pax5 are involved in lineage commitment [12,13,14,15]. Additionally, V (D) J recombination, which is a unique mechanism of genetic recombination that occurs in developing lymphocytes during the early stages of T- and B-cell maturation, is tightly regulated by transcription factors.
Mature B cells migrate to secondary lymph tissues, including the lymph nodes and spleen. Upon antigen stimulation, B cells relocate to the germinal center (GC), where they and helper T cells are activated. Activated mature B cells in the GC induce somatic hypermutation and class switch recombination, which involve genomic alterations. The activated B cells ultimately differentiate into plasma cells that produce antibodies and memory B cells.

3. miRNAs in B-Cell Differentiation

In addition to transcription factors, recent studies have identified miRNAs as regulators of gene expression. miRNAs are a class of 18–24 nucleotide (nt)-long noncoding RNAs that post-transcriptionally downregulate target genes. Primary miRNA transcripts are processed by the nuclear nuclease Drosha into ~60 nt precursor miRNA hairpins that are cleaved by the Dicer nuclease in the cytosol, generating mature miRNAs. Mature miRNAs are incorporated into the multiprotein RNA-induced silencing complex (RISC), which consists of Argonaut proteins (Ago). RISC post-transcriptionally represses target mRNAs by inducing their degradation or blocking translation [16,17,18]. Each miRNA has multiple target mRNAs, and in silico predictions suggest that more than one-third of all human genes are targets of miRNAs [19]. In animals, miRNAs regulate multiple developmental and physiological processes. For example, abnormal expression of specific miRNAs, including let-7 and lin-41, leads to developmental arrest in C. elegans [20].
miRNAs are also involved in the sequential differentiation of B cells, which was revealed in studies involving Ago2-deficient mice or mice lacking Dicer [21,22]. Ago2 is the primary component of the RISC protein complex, in which miRNAs bind target RNAs and prevent translation. Four Ago-family proteins exist in mammals; however, only Ago1 and Ago2 are expressed in hematopoietic cells (unpublished data). Therefore, loss of Ago2 largely inactivates the miRNA machinery during hematopoiesis.
Dicer performs the final step in miRNA processing, and thus, its loss prevents miRNA maturation. Notably, mice deficient in Ago2 and Dicer fail to convert pro-B cells into pre-B cells, which is critical for B-cell lineage commitment [23]. Multiple miRNAs, including miR-17–92, miR-34a, miR-125b, miR-150, miR-181a, and miR-212/132, regulate early B-cell development [24,25,26,27,28,29,30,31]. The miR-17–92 cluster and miR-181a positively regulate B-cell development, whereas the others negatively regulate the process. The miR-17–92 cluster is essential for the pro- to pre-B cell transition, since it prevents expression of the pro-apoptotic protein BIM [24]. Ectopic expression of miR-181a enhanced B-cell differentiation [30], while miR-34a negatively regulated B-cell development by downregulating FOXP1 [25].
Ectopic expression of miR-125b skewed hematopoiesis toward the myeloid lineage, whereas B-cell differentiation was decreased [27]. This finding was likely due to the repression of LIN28A by miR-125b. Ectopic expression of miR-150 impaired B-cell development [29]. B1-cell differentiation is also regulated by miR-150. The transcription factor c-Myb is a target of miR-150, and its repression may be responsible for those observations. The miR-212/132 cluster suppresses B-cell differentiation at the pre-B to pro-B transition stage by repressing SOX4 [31]. Thus, multiple miRNAs are involved in B-cell development.
During late B-cell maturation in follicles, the downregulation of miR-150 is required for GC selection and development of the adaptive humoral immune response [29]. Other miRNAs, including miR-155, miR-181b, miR-15a, miR-16, miR-15b, miR-34a, miR-9, miR-30, let-7a, miR-125b, miR-217, and miR-185, modulate the expression of genes involved in B-cell maturation [32]. The GC is a transient structure that forms within the peripheral lymphoid organs in response to B-lymphocyte stimulation. Zhang et al., categorized mature B cells into naïve, memory, GC, and plasma cells and found that miR-223 was downregulated specifically in GC-B cells. miR-223 targets LMO2, which is abundant in GC-B cells. Conversely, miR-30 and miR-9, which target the PRDM1 transcription factor, were upregulated in GC-B cells, in which PRDM1 was found to be critical for plasma cell differentiation [33]. Malpeli et al., reported that of 48 miRNAs differentially expressed in naïve, GC, and subepithelial B cells, eight (miR-323, miR-138, miR-9*, miR-211, miR-149, miR-373, miR-135a and miR-184) were specific to follicular cells [34]. miR-125b also repressed the expression of Prdm1 and Irf4, which encode transcription factors involved in plasma cell differentiation [26,27]. Additionally, miR-125b targeted Bright/ARID3a, which encodes a transcription factor involved in the expression of Ig heavy chains and B1-cell differentiation [28,35,36].

4. miRNAs in B Cell Malignancy

4.1. General View of miRNAs in Cancer

Multiple studies have correlated changes in miRNA expression profiles with human tumor phenotypes [35,36]. The first studies on the roles of miRNAs in cancer were of miR-15 in chronic lymphocytic leukemia (CLL). Croce et al., found that miR-15a and miR-16-1 are located at the 13q14 locus, which was deleted in the majority of CLL cases [37]. Since then, many miRNA have been revealed to have unique functions in cancer biology. For example, the expression of let-7 family members, which downregulate the expression of Ras and other proto-oncogenes, was reduced in lung cancer [38,39]. In several B cell malignancies, including acute lymphoblastic leukemia (ALL) and lymphoma, let-7 showed tumor suppressive functions [40,41]. Additionally, miR-15 family members, including miR-16, which was originally identified in CLL, a B cell malignancy, and miR-195, harbor tumor-suppressive functions by downregulating BCL2 in several cancers. miR-200 suppresses the epithelial-mesenchymal transition, which is involved in metastasis [42,43]. In gastric diffuse large B cell lymphoma (DLBCL), miR-200 is tumor suppressive [44].
In contrast, the expression of miR-21, miR-17–92, and miR-155 is typically increased in cancers [45], suggesting their roles as oncogenes [46,47]. miRNAs expression, similar to that of protein-coding genes, is regulated by multiple transcriptional networks as well as the epigenetic machinery. In addition, miRNAs can themselves repress key enzymes that drive epigenetic remodeling, generating regulatory circuits that have a significant effect in the transcriptional landscape of the cell. Recent evidence also suggests that miRNAs can directly modulate gene transcription in the nucleus through the recognition of specific target sites in promoter regions. These mechanisms are also involved in the expression of tumor suppressive and oncogenic miRNAs [48]. Collectively, miRNAs play a primary role in cancer biology. Here, we focused on the miR-21, miR-34a, miR-150, miR-155, and the miR-17–92 clusters, which are involved in B-cell malignancies (Figure 1).

4.2. miR-21

miR-21 is one of the most abundant miRNAs in solid tumors. miR-21 is expressed in non-Hodgkin lymphomas. In diffuse large B cell lymphoma (DLBCL), miR-21 was found expressed at a higher level in ABC-type than in GC-type DLBCL [49]. miR-21 expression is correlated with the prognosis of CLL [50]. Rossi et al., performed quantitative reverse-transcription polymerase chain reaction (qRT-PCR) profiling in 104 CLL patients with a well-defined chromosome 17p status and found that miR-21 was differentially expressed between CLLs with a 17p deletion and those with a normal 17p karyotype. Moreover, miR-21 levels were significantly higher in patients with a poor prognosis and were correlated with predictions of overall survival (OS). Accordingly, a 21FK score (miR-21 qRT-PCR, fluorescence in situ hybridization, karyotype) was developed to stratify patients according to OS. Importantly, studies found that patients with a low score had a significantly longer OS. Evaluations of the power of the 21FK score with common prognostic factors revealed that the score was most significant in both CLL cohorts. Thus, 21FK score is useful for distinguishing between good and poor prognoses in CLL patients [50].
miR-21 is induced by EBNA2, a protein encoded by the Epstein-Barr virus (EBV) genome that causes cell proliferation though phosphorylation of AKT [51]. One of the targets of miR-21 is PTEN, which encodes a tumor suppressor [52]. Exogenous expression of miR-21 induced pre-B-cell leukemia [53], while suppression of miR-21 caused apoptosis and tumor regression. miR-21 knockdown in DLBCL cells increased their sensitivity to cyclophosphamide, vincristine, Adriamycin, and prednisone (CHOP) chemotherapies [52]. Knockdown of NF-κB also induced the same effect in DLBCL cells. Genotoxic treatments upregulated oncogenic miR-21 expression by recruiting NF-κB to the miR-21 promoter, where it activated the signal transducer and activator of transcription 3 [54]. miR-21 regulates NF-κB through several pathways, including extracellular signal-regulated kinase (ERK) [55]. Those observations suggest that NF-κB and miR-21 formed a positive feedback loop.

4.3. miR-34a

miR-34a is a tumor suppressor whose expression is regulated by another tumor suppressor, p53. Downregulation of miR-34a is correlated with a poor prognosis in CLL, DLBCL, and mantle cell lymphoma (MCL). Exogenous expression of miR-34 downregulated MYB and E2F1 and caused cell-cycle arrest. miR-34a regulates AXL, which encodes a receptor, tyrosine kinase, overexpressed in CLL [56]. MYC negatively regulates miR-34a expression by binding to its promoter. Notably, miR-34a was downregulated in lymphomas with high levels of MYC, while it was also found to regulate many targets of MYC [57]. The details of this contradiction have not yet been elucidated. The downregulation of miR-34a also induced FOXP1 and BCL6, which resulted in proliferation of DLBCL [57].

4.4. miR-150

miR-150 is a tumor suppressor that is downregulated in DLBCL, MCL, aggressive CLL, and Burkitt lymphoma (BL) cells [58,59]. Rescue of miR-150 in BL cells suppresses cell proliferation [58]. The targets of miR-150 are MYB, GAB1, and FOXP1, which affect B-cell receptor (BCR) and AKT signaling [59]. GAB1 acts upstream of BCR signaling and juxtaposes phosphoinositide 3-kinase (PI3K) at the plasma membrane. FOXP1 regulates the expression of multiple genes upon activation of BCR signaling, and overexpression of FOXP1 was correlated with a poor prognosis in CLL, DLBCL, and follicular lymphomas (FLs) [59]. Overexpression of FOXP1 in B-cell lymphomas is caused by downregulation of miR-34a and miR-150.
The chemokine receptor CXCR4 is also a target of miR-150. Overexpression of CXCR4 is caused by downregulation of miR-150 and activates the invasion of tumor cells [60]. However, the mechanisms underlying miR-150 regulation are unknown.

4.5. miR-155

miR-155 overexpression occurred in DLBCL, CLL, FL, and MCL [61,62,63,64,65]. The ABC-type DLBCL had a poor prognosis and was associated with increased expression of miR-155. E(mu)-miR-155 transgenic mice developed malignant B-cell lymphomas [47], and miR-155 was involved in lymphomagenesis by repressing IL-6 signaling, which resulted in dysregulation of B-cell differentiation [66,67]. miR-155 also repressed Smad5, which suppressed TGF-beta signaling and increased cell proliferation [68,69,70,71]. Additionally, miR-155 facilitated cell migration by downregulating HGAL, which may contribute to malignant tumorigenesis [72]. Iqbal et al., identified predictive miRNA biomarker signatures in DLBCL, including miR-155, which was significantly associated with rituximab plus CHOP treatment failure [73].
miR-155 is regulated by NF-κB. LMP1, which is encoded by the EBV genome, activates NF-κB signaling and induces the expression of miR-155 in EBV-infected B-cell malignancies [74]. NF-κB-induced overexpression of miR-155 also plays a role in ABC-type DLBCL. Thus, since miR-155 is involved in B-cell lymphomagenesis, it may be a promising therapeutic target.

4.6. miR-17‒92 Cluster

The miR-17‒92 cluster is comprised of six miRNAs, including miR-17, miR-18a, miR-19a, miR-20a, miR-19b, and miR-92a. This cluster facilitates formation of B-cell lymphomas in E(mu)-c-MYC transgenic mice, and its overexpression is involved in the formation of malignant B-cell lymphomas. Exogenous expression of the cluster induced development of DLBCL and CLL, suggesting that erroneous expression of miR-17‒92 causes lymphomas [75].
Overexpression of miR-17‒92 caused B-cell lymphomas as a result of downregulation of PTEN and BIM. Enforced expression of the mir-17–92 cluster acted with c-MYC expression to accelerate tumor development in a mouse B-cell lymphoma model [46].
MYC upregulated miR-17‒92 and E2F1, whereas miR-17 and miR-20a suppressed apoptosis by regulating E2F1 [76]. Recent studies revealed that the miR-17‒92 cluster activated the PI3K-AKT-mTOR pathway. miR-19a and miR-19b have also been found to be responsible for lymphomagenesis.
Intriguingly, miR-92a suppressed the function of miR-19. miR-92 induced MYC expression and p53-dependent apoptosis by suppressing FBW7, a ubiquitination enzyme targeting MYC [77]. Conversely, miR-19 counteracted miR-92-dependent apoptosis by activating MDM2, which targets p53 [78]. Depletion of the miR-17‒92 cluster exhibits more potent tumorigenesis. Collectively, the tumorigenic potency of the miR-17‒92 cluster is regulated by miR-19 and miR-92a.

5. Application of miRNAs in Diagnosis and Therapy

miR-21 and miR-155 are promising diagnostic markers for DLBCL and CLL [52,61,73,79,80,81,82]. Using in situ hybridization for formalin-embedded sections, such diagnostic markers will become simple to access [83,84]. Recent trials have used liquid biopsies to diagnose cancer during the early stages by use of exosome, which is stable in fluids such as blood, urine, and cerebral spinal fluid.
Exosomes contain a myriad of miRNAs [85,86]. Several miRNAs are promising diagnostic markers for several lymphomas; however, the methods of collecting and analyzing exosomes are variable. Thus, the development of standardized procedures for collecting and analyzing exosomes will be important for future studies.
Therapeutic trials using miRNAs have been performed for hepatitis C, and the greatest challenge reported by these trials was related to drug delivery [87,88]. Drug delivery is more complicated in hematopoietic cells compared with other cell types, including hepatocytes. Importantly, Ivan et al., reported that modified nucleic acids can be used to overcome drug delivery challenges in lymphomas. That study revealed that suppression of miR-155 in a mouse lymphoma model caused remission within one week [89]. Thus, miRNA regulation is one approach for drug delivery.

6. Concluding Remarks

miRNAs were discovered in mammals in 2000 and have since been linked to many cancers. Studies showed that miRNAs were more precise indicators of cancer phenotypes and prognoses than mRNAs. Those studies also presented miRNAs as promising diagnostic targets. Multiple researchers explored the potential of miRNA-based therapeutics. In the antisense approach to inhibit specific miRNAs, the antisense of miR-122, miravirsen, showed good results for HCV hepatitis on a phase2a trial. Miravirsen is a locked nucleic acid-modified DNA phosphorothioate antisense oligonucleotide that sequesters mature miR-122 in a highly stable heteroduplex, thereby inhibiting its function [88]. In the replacement approach to enhance the function of specific miRNAs, the replacement of miR-34 for advanced cancer is now on phase one trial [90]. Accordingly, miRNAs were promising targets for the diagnosis and treatment of B-cell malignancies.

Acknowledgments

We thank our laboratory members for the helpful discussions. This work was supported by the Japan Society for the Promotion of Science and the Research Program on Hepatitis from the Japan Agency for Medical Research and Development.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mandel, E.M.; Grosschedl, R. Transcription control of early B cell differentiation. Curr. Opin. Immunol. 2010, 22, 161–167. [Google Scholar] [CrossRef] [PubMed]
  2. Miyazaki, K.; Miyazaki, M.; Murre, C. The establishment of B versus T cell identity. Trends Immunol. 2014, 35, 205–210. [Google Scholar] [CrossRef] [PubMed]
  3. Rothenberg, E.V. Transcriptional control of early T and B cell developmental choices. Annu. Rev. Immunol. 2014, 32, 283–321. [Google Scholar] [CrossRef] [PubMed]
  4. Kawamoto, H.; Ohmura, K.; Katsura, Y. Direct evidence for the commitment of hematopoietic stem cells to T, B and myeloid lineages in murine fetal liver. Int. Immunol. 1997, 9, 1011–1019. [Google Scholar] [CrossRef] [PubMed]
  5. Kawamoto, H.; Ohmura, K.; Fujimoto, S.; Katsura, Y. Emergence of T cell progenitors without B cell or myeloid differentiation potential at the earliest stage of hematopoiesis in the murine fetal liver. J. Immunol. 1999, 162, 2725–2731. [Google Scholar] [PubMed]
  6. Kawamoto, H.; Ikawa, T.; Ohmura, K.; Fujimoto, S.; Katsura, Y. T cell progenitors emerge earlier than B cell progenitors in the murine fetal liver. Immunity 2000, 12, 441–450. [Google Scholar] [CrossRef]
  7. Kawamoto, H.; Ikawa, T.; Masuda, K.; Wada, H.; Katsura, Y. A map for lineage restriction of progenitors during hematopoiesis: The essence of the myeloid-based model. Immunol. Rev. 2010, 238, 23–36. [Google Scholar] [CrossRef] [PubMed]
  8. Adolfsson, J.; Månsson, R.; Buza-Vidas, N.; Hultquist, A.; Liuba, K.; Jensen, C.T.; Bryder, D.; Yang, L.; Borge, O.J.; Thoren, L.A.; et al. Identification of Flt3+ lympho-myeloid stem cells lacking erythro-megakaryocytic potential a revised road map for adult blood lineage commitment. Cell 2005, 121, 295–306. [Google Scholar] [CrossRef] [PubMed]
  9. Forsberg, E.C.; Serwold, T.; Kogan, S.; Weissman, I.L.; Passegué, E. New evidence supporting megakaryocyte-erythrocyte potential of Flk2/Flt3+ multipotent hematopoietic progenitors. Cell 2006, 126, 415–426. [Google Scholar] [CrossRef] [PubMed]
  10. Chi, A.W.; Chavez, A.; Xu, L.; Weber, B.N.; Shestova, O.; Schaffer, A.; Wertheim, G.; Pear, W.S.; Izon, D.; Bhandoola, A. Identification of Flt3⁺CD150− myeloid progenitors in adult mouse bone marrow that harbor t lymphoid developmental potential. Blood 2011, 118, 2723–2732. [Google Scholar] [CrossRef] [PubMed]
  11. Luc, S.; Luis, T.C.; Boukarabila, H.; Macaulay, I.C.; Buza-Vidas, N.; Bouriez-Jones, T.; Lutteropp, M.; Woll, P.S.; Loughran, S.J.; Mead, A.J.; et al. The earliest thymic T cell progenitors sustain B cell and myeloid lineage potential. Nat. Immunol. 2012, 13, 412–419. [Google Scholar] [CrossRef] [PubMed]
  12. Ikawa, T.; Kawamoto, H.; Wright, L.Y.; Murre, C. Long-term cultured E2A-deficient hematopoietic progenitor cells are pluripotent. Immunity 2004, 20, 349–360. [Google Scholar] [CrossRef]
  13. Pongubala, J.M.; Northrup, D.L.; Lancki, D.W.; Medina, K.L.; Treiber, T.; Bertolino, E.; Thomas, M.; Grosschedl, R.; Allman, D.; Singh, H. Transcription factor EBF restricts alternative lineage options and promotes B cell fate commitment independently of Pax5. Nat. Immunol. 2008, 9, 203–215. [Google Scholar] [CrossRef] [PubMed]
  14. Györy, I.; Boller, S.; Nechanitzky, R.; Mandel, E.; Pott, S.; Liu, E.; Grosschedl, R. Transcription factor Ebf1 regulates differentiation stage-specific signaling, proliferation, and survival of B cells. Genes Dev. 2012, 26, 668–682. [Google Scholar] [CrossRef] [PubMed]
  15. Lin, Y.C.; Jhunjhunwala, S.; Benner, C.; Heinz, S.; Welinder, E.; Mansson, R.; Sigvardsson, M.; Hagman, J.; Espinoza, C.A.; Dutkowski, J.; et al. A global network of transcription factors, involving E2A, Ebf1 and FOXO1, that orchestrates B cell fate. Nat. Immunol. 2010, 11, 635–643. [Google Scholar] [CrossRef] [PubMed]
  16. Bagga, S.; Bracht, J.; Hunter, S.; Massirer, K.; Holtz, J.; Eachus, R.; Pasquinelli, A.E. Regulation by let-7 and lin-4 miRNAs results in target mrna degradation. Cell 2005, 122, 553–563. [Google Scholar] [CrossRef] [PubMed]
  17. Yekta, S.; Shih, I.; Bartel, D. MicroRNA-directed cleavage of hoxb8 mRNA. Science 2004, 304, 594–596. [Google Scholar] [CrossRef] [PubMed]
  18. Kotani, A.; Ha, D.; Hsieh, J.; Rao, P.K.; Schotte, D.; den Boer, M.L.; Armstrong, S.A.; Lodish, H.F. miR-128b is a potent glucocorticoid sensitizer in MLL-AF4 acute lymphocytic leukemia cells and exerts cooperative effects with miR-221. Blood 2009, 114, 4169–4178. [Google Scholar] [CrossRef] [PubMed]
  19. Lewis, B.P.; Burge, C.B.; Bartel, D.P. Conserved seed pairing, often flanked by adenosines, indicates that thousands of human genes are microRNA targets. Cell 2005, 120, 15–20. [Google Scholar] [CrossRef] [PubMed]
  20. Ambros, V. The functions of animal microRNAs. Nature 2004, 431, 350–355. [Google Scholar] [CrossRef] [PubMed]
  21. O’Carroll, D.; Mecklenbrauker, I.; Das, P.; Santana, A.; Koenig, U.; Enright, A.; Miska, E.; Tarakhovsky, A. A slicer-independent role for argonaute 2 in hematopoiesis and the microrna pathway. Genes Dev. 2007, 21, 1999–2004. [Google Scholar] [CrossRef] [PubMed]
  22. Koralov, S.; Muljo, S.; Galler, G.; Krek, A.; Chakraborty, T.; Kanellopoulou, C.; Jensen, K.; Cobb, B.; Merkenschlager, M.; Rajewsky, N.; et al. Dicer ablation affects antibody diversity and cell survival in the B lymphocyte lineage. Cell 2008, 132, 860–874. [Google Scholar] [CrossRef] [PubMed]
  23. Ademokun, A.; Turner, M. Regulation of B-cell differentiation by microRNAs and RNA-binding proteins. Biochem. Soc. Trans. 2008, 36, 1191–1193. [Google Scholar] [CrossRef] [PubMed]
  24. Ventura, A.; Young, A.G.; Winslow, M.M.; Lintault, L.; Meissner, A.; Erkeland, S.J.; Newman, J.; Bronson, R.T.; Crowley, D.; Stone, J.R.; et al. Targeted deletion reveals essential and overlapping functions of the miR-17 through 92 family of miRNA clusters. Cell 2008, 132, 875–886. [Google Scholar] [CrossRef] [PubMed]
  25. Rao, D.S.; O’Connell, R.M.; Chaudhuri, A.A.; Garcia-Flores, Y.; Geiger, T.L.; Baltimore, D. MicroRNA-34a perturbs b lymphocyte development by repressing the forkhead box transcription factor FOXP1. Immunity 2010, 33, 48–59. [Google Scholar] [CrossRef] [PubMed]
  26. Gururajan, M.; Haga, C.L.; Das, S.; Leu, C.M.; Hodson, D.; Josson, S.; Turner, M.; Cooper, M.D. MicroRNA 125b inhibition of B cell differentiation in germinal centers. Int. Immunol. 2010, 22, 583–592. [Google Scholar] [CrossRef] [PubMed]
  27. Chaudhuri, A.A.; So, A.Y.; Mehta, A.; Minisandram, A.; Sinha, N.; Jonsson, V.D.; Rao, D.S.; O’Connell, R.M.; Baltimore, D. Oncomir miR-125b regulates hematopoiesis by targeting the gene lin28a. Proc. Natl. Acad. Sci. USA 2012, 109, 4233–4238. [Google Scholar] [CrossRef] [PubMed]
  28. Puissegur, M.P.; Eichner, R.; Quelen, C.; Coyaud, E.; Mari, B.; Lebrigand, K.; Broccardo, C.; Nguyen-Khac, F.; Bousquet, M.; Brousset, P. B-cell regulator of immunoglobulin heavy-chain transcription (bright)/arid3a is a direct target of the oncomir microRNA-125b in progenitor B-cells. Leukemia 2012, 26, 2224–2232. [Google Scholar] [CrossRef] [PubMed]
  29. Xiao, C.; Calado, D.; Galler, G.; Thai, T.; Patterson, H.; Wang, J.; Rajewsky, N.; Bender, T.; Rajewsky, K. miR-150 controls B cell differentiation by targeting the transcription factor c-myb. Cell 2007, 131, 146–159. [Google Scholar] [CrossRef] [PubMed]
  30. Chen, C.Z.; Li, L.; Lodish, H.F.; Bartel, D.P. MicroRNAs modulate hematopoietic lineage differentiation. Science 2004, 303, 83–86. [Google Scholar] [CrossRef] [PubMed]
  31. Mehta, A.; Mann, M.; Zhao, J.L.; Marinov, G.K.; Majumdar, D.; Garcia-Flores, Y.; Du, X.; Erikci, E.; Chowdhury, K.; Baltimore, D. The microRNA-212/132 cluster regulates B cell development by targeting SOX4. J. Exp. Med. 2015, 212, 1679–1692. [Google Scholar] [CrossRef] [PubMed]
  32. Tan, L.P.; Wang, M.; Robertus, J.L.; Schakel, R.N.; Gibcus, J.H.; Diepstra, A.; Harms, G.; Peh, S.C.; Reijmers, R.M.; Pals, S.T.; et al. Mirna profiling of B-cell subsets: Specific miRNA profile for germinal center B cells with variation between centroblasts and centrocytes. Lab. Investig. 2009, 89, 708–716. [Google Scholar] [CrossRef] [PubMed]
  33. Zhang, J.; Jima, D.; Jacobs, C.; Fischer, R.; Gottwein, E.; Huang, G.; Lugar, P.; Lagoo, A.; Rizzieri, D.; Friedman, D.; et al. Patterns of microRNA expression characterize stages of human B-cell differentiation. Blood 2009, 113, 4586–4594. [Google Scholar] [CrossRef] [PubMed]
  34. Malpeli, G.; Barbi, S.; Zupo, S.; Tosadori, G.; Scardoni, G.; Bertolaso, A.; Sartoris, S.; Ugel, S.; Vicentini, C.; Fassan, M.; et al. Identification of microRNAs implicated in the late differentiation stages of normal B cells suggests a central role for miRNA targets zeb1 and tp53. Oncotarget 2017, 8, 11809–11826. [Google Scholar] [CrossRef] [PubMed]
  35. Lu, J.; Getz, G.; Miska, E.A.; Alvarez-Saavedra, E.; Lamb, J.; Peck, D.; Sweet-Cordero, A.; Ebert, B.L.; Mak, R.H.; Ferrando, A.A.; et al. MicroRNA expression profiles classify human cancers. Nature 2005, 435, 834–838. [Google Scholar] [CrossRef] [PubMed]
  36. Volinia, S.; Calin, G.A.; Liu, C.G.; Ambs, S.; Cimmino, A.; Petrocca, F.; Visone, R.; Iorio, M.; Roldo, C.; Ferracin, M.; et al. A microRNA expression signature of human solid tumors defines cancer gene targets. Proc. Natl. Acad. Sci. USA 2006, 103, 2257–2261. [Google Scholar] [CrossRef] [PubMed]
  37. Calin, G.A.; Dumitru, C.D.; Shimizu, M.; Bichi, R.; Zupo, S.; Noch, E.; Aldler, H.; Rattan, S.; Keating, M.; Rai, K.; et al. Frequent deletions and down-regulation of micro-RNA genes miR15 and miR16 at 13q14 in chronic lymphocytic leukemia. Proc. Natl. Acad. Sci. USA 2002, 99, 15524–15529. [Google Scholar] [CrossRef] [PubMed]
  38. Johnson, S.M.; Grosshans, H.; Shingara, J.; Byrom, M.; Jarvis, R.; Cheng, A.; Labourier, E.; Reinert, K.L.; Brown, D.; Slack, F.J. Ras is regulated by the let-7 microrna family. Cell 2005, 120, 635–647. [Google Scholar] [CrossRef] [PubMed]
  39. Takamizawa, J.; Konishi, H.; Yanagisawa, K.; Tomida, S.; Osada, H.; Endoh, H.; Harano, T.; Yatabe, Y.; Nagino, M.; Nimura, Y.; et al. Reduced expression of the let-7 micrornas in human lung cancers in association with shortened postoperative survival. Cancer Res. 2004, 64, 3753–3756. [Google Scholar] [CrossRef] [PubMed]
  40. Adams, C.M.; Hiebert, S.W.; Eischen, C.M. Myc induces miRNA-mediated apoptosis in response to HDAC inhibition in hematologic malignancies. Cancer Res. 2016, 76, 736–748. [Google Scholar] [CrossRef] [PubMed]
  41. Schotte, D.; De Menezes, R.X.; Akbari Moqadam, F.; Khankahdani, L.M.; Lange-Turenhout, E.; Chen, C.; Pieters, R.; Den Boer, M.L. Microrna characterize genetic diversity and drug resistance in pediatric acute lymphoblastic leukemia. Haematologica 2011, 96, 703–711. [Google Scholar] [CrossRef] [PubMed]
  42. Gregory, P.A.; Bert, A.G.; Paterson, E.L.; Barry, S.C.; Tsykin, A.; Farshid, G.; Vadas, M.A.; Khew-Goodall, Y.; Goodall, G.J. The miR-200 family and miR-205 regulate epithelial to mesenchymal transition by targeting zeb1 and sip1. Nat. Cell Biol. 2008, 10, 593–601. [Google Scholar] [CrossRef] [PubMed]
  43. Korpal, M.; Lee, E.S.; Hu, G.; Kang, Y. The miR-200 family inhibits epithelial-mesenchymal transition and cancer cell migration by direct targeting of E-cadherin transcriptional repressors zeb1 and zeb2. J. Biol. Chem. 2008, 283, 14910–14914. [Google Scholar] [CrossRef] [PubMed]
  44. Huang, W.T.; Kuo, S.H.; Cheng, A.L.; Lin, C.W. Inhibition of zeb1 by miR-200 characterizes helicobacter pylori-positive gastric diffuse large B-cell lymphoma with a less aggressive behavior. Mod. Pathol. 2014, 27, 1116–1125. [Google Scholar] [CrossRef] [PubMed]
  45. Metzler, M.; Wilda, M.; Busch, K.; Viehmann, S.; Borkhardt, A. High expression of precursor microRNA-155/bic RNA in children with burkitt lymphoma. Genes Chromosomes Cancer 2004, 39, 167–169. [Google Scholar] [CrossRef] [PubMed]
  46. He, L.; Thomson, J.; Hemann, M.; Hernando-Monge, E.; Mu, D.; Goodson, S.; Powers, S.; Cordon-Cardo, C.; Lowe, S.; Hannon, G.; et al. A microRNA polycistron as a potential human oncogene. Nature 2005, 435, 828–833. [Google Scholar] [CrossRef] [PubMed]
  47. Costinean, S.; Zanesi, N.; Pekarsky, Y.; Tili, E.; Volinia, S.; Heerema, N.; Croce, C.M. Pre-B cell proliferation and lymphoblastic leukemia/high-grade lymphoma in Eμ-miR155 transgenic mice. Proc. Natl. Acad. Sci. USA 2006, 103, 7024–7029. [Google Scholar] [CrossRef] [PubMed]
  48. Malumbres, M. miRNAs and cancer: An epigenetics view. Mol. Asp. Med. 2013, 34, 863–874. [Google Scholar] [CrossRef] [PubMed]
  49. Lawrie, C.; Soneji, S.; Marafioti, T.; Cooper, C.; Palazzo, S.; Paterson, J.; Cattan, H.; Enver, T.; Mager, R.; Boultwood, J.; 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, 121, 1156–1161. [Google Scholar] [CrossRef] [PubMed]
  50. Rossi, S.; Shimizu, M.; Barbarotto, E.; Nicoloso, M.S.; Dimitri, F.; Sampath, D.; Fabbri, M.; Lerner, S.; Barron, L.L.; Rassenti, L.Z.; et al. MicroRNA fingerprinting of CLL patients with chromosome 17p deletion identify a miR-21 score that stratifies early survival. Blood 2010, 116, 945–952. [Google Scholar] [CrossRef] [PubMed]
  51. Rosato, P.; Anastasiadou, E.; Garg, N.; Lenze, D.; Boccellato, F.; Vincenti, S.; Severa, M.; Coccia, E.M.; Bigi, R.; Cirone, M.; et al. Differential regulation of miR-21 and miR-146a by epstein-barr virus-encoded EBNA2. Leukemia 2012, 26, 2343–2352. [Google Scholar] [CrossRef] [PubMed]
  52. Bai, H.; Wei, J.; Deng, C.; Yang, X.; Wang, C.; Xu, R. MicroRNA-21 regulates the sensitivity of diffuse large B-cell lymphoma cells to the CHOP chemotherapy regimen. Int. J. Hematol. 2013, 97, 223–231. [Google Scholar] [CrossRef] [PubMed]
  53. Medina, P.P.; Nolde, M.; Slack, F.J. OncomiR addiction in an in vivo model of microRNA-21-induced pre-B-cell lymphoma. Nature 2010, 467, 86–90. [Google Scholar] [CrossRef] [PubMed]
  54. Niu, J.; Shi, Y.; Tan, G.; Yang, C.H.; Fan, M.; Pfeffer, L.M.; Wu, Z.H. DNA damage induces NF-κB-dependent microRNA-21 up-regulation and promotes breast cancer cell invasion. J. Biol. Chem. 2012, 287, 21783–21795. [Google Scholar] [CrossRef] [PubMed]
  55. Koti, M.; Gooding, R.J.; Nuin, P.; Haslehurst, A.; Crane, C.; Weberpals, J.; Childs, T.; Bryson, P.; Dharsee, M.; Evans, K.; et al. Identification of the IGF1/PI3K/NF κB/ERK gene signalling networks associated with chemotherapy resistance and treatment response in high-grade serous epithelial ovarian cancer. BMC Cancer 2013, 13, 549. [Google Scholar] [CrossRef] [PubMed]
  56. Boysen, J.; Sinha, S.; Price-Troska, T.; Warner, S.L.; Bearss, D.J.; Viswanatha, D.; Shanafelt, T.D.; Kay, N.E.; Ghosh, A.K. The tumor suppressor axis p53/miR-34a regulates axl expression in B-cell chronic lymphocytic leukemia: Implications for therapy in p53-defective CLL patients. Leukemia 2014, 28, 451–455. [Google Scholar] [CrossRef] [PubMed]
  57. Craig, V.J.; Cogliatti, S.B.; Imig, J.; Renner, C.; Neuenschwander, S.; Rehrauer, H.; Schlapbach, R.; Dirnhofer, S.; Tzankov, A.; Müller, A. Myc-mediated repression of microRNA-34a promotes high-grade transformation of B-cell lymphoma by dysregulation of FOXP1. Blood 2011, 117, 6227–6236. [Google Scholar] [CrossRef] [PubMed]
  58. Chen, S.; Wang, Z.; Dai, X.; Pan, J.; Ge, J.; Han, X.; Wu, Z.; Zhou, X.; Zhao, T. Re-expression of microRNA-150 induces EBV-positive burkitt lymphoma differentiation by modulating c-myb in vitro. Cancer Sci. 2013, 104, 826–834. [Google Scholar] [CrossRef] [PubMed]
  59. Mraz, M.; Chen, L.; Rassenti, L.Z.; Ghia, E.M.; Li, H.; Jepsen, K.; Smith, E.N.; Messer, K.; Frazer, K.A.; Kipps, T.J. miR-150 influences B-cell receptor signaling in chronic lymphocytic leukemia by regulating expression of GAB1 and FOXP1. Blood 2014, 124, 84–95. [Google Scholar] [CrossRef] [PubMed]
  60. Tano, N.; Kim, H.W.; Ashraf, M. MicroRNA-150 regulates mobilization and migration of bone marrow-derived mononuclear cells by targeting CXCR4. PLoS ONE 2011, 6, e23114. [Google Scholar] [CrossRef] [PubMed]
  61. Eis, P.; Tam, W.; Sun, L.; Chadburn, A.; Li, Z.; Gomez, M.; Lund, E.; Dahlberg, J. Accumulation of miR-155 and bic RNA in human B cell lymphomas. Proc. Natl. Acad. Sci. USA 2005, 102, 3627–3632. [Google Scholar] [CrossRef] [PubMed]
  62. Kluiver, J.; Poppema, S.; de Jong, D.; Blokzijl, T.; Harms, G.; Jacobs, S.; Kroesen, B.; van den Berg, A. Bic and miR-155 are highly expressed in hodgkin, primary mediastinal and diffuse large B cell lymphomas. J. Pathol. 2005, 207, 243–249. [Google Scholar] [CrossRef] [PubMed]
  63. Tam, W.; Dahlberg, J. miR-155/bic as an oncogenic microRNA. Genes Chromosomes Cancer 2006, 45, 211–212. [Google Scholar] [CrossRef] [PubMed]
  64. Ivan, M.; Harris, A.; Martelli, F.; Kulshreshtha, R. Hypoxia response and micrornas: No longer two separate worlds. J. Cell. Mol. Med. 2008, 12, 1426–1431. [Google Scholar] [CrossRef] [PubMed]
  65. Wang, M.; Tan, L.; Dijkstra, M.; van Lom, K.; Robertus, J.; Harms, G.; Blokzijl, T.; Kooistra, K.; van T’veer, M.; Rosati, S.; et al. miRNA analysis in B-cell chronic lymphocytic leukaemia: Proliferation centres characterized by low miR-150 and high bic/miR-155 expression. J. Pathol. 2008, 215, 13–20. [Google Scholar] [CrossRef] [PubMed]
  66. Huang, J.; Yang, Q.; He, L. Role of TLR4 and miR-155 in peripheral blood mononuclear cell-mediated inflammatory reaction in coronary slow flow and coronary arteriosclerosis patients. J. Clin. Lab. Anal. 2017. [Google Scholar] [CrossRef] [PubMed]
  67. Ma, F.; Liu, F.; Ding, L.; You, M.; Yue, H.; Zhou, Y.; Hou, Y. Anti-inflammatory effects of curcumin are associated with down regulating microRNA-155 in LPS-treated macrophages and mice. Pharm. Biol. 2017, 55, 1263–1273. [Google Scholar] [CrossRef] [PubMed]
  68. Louafi, F.; Martinez-Nunez, R.T.; Sanchez-Elsner, T. MicroRNA-155 targets SMAD2 and modulates the response of macrophages to transforming growth factor-β. J. Biol. Chem. 2010, 285, 41328–41336. [Google Scholar] [CrossRef] [PubMed]
  69. Ji, H.; Li, Y.; Jiang, F.; Wang, X.; Zhang, J.; Shen, J.; Yang, X. Inhibition of transforming growth factor β/SMAD signal by miR-155 is involved in arsenic trioxide-induced anti-angiogenesis in prostate cancer. Cancer Sci. 2014, 105, 1541–1549. [Google Scholar] [CrossRef] [PubMed]
  70. Cazac, B.B.; Roes, J. TGF-β receptor controls B cell responsiveness and induction of IgA in vivo. Immunity 2000, 13, 443–451. [Google Scholar] [CrossRef]
  71. Zhang, D.; Cui, Y.; Li, B.; Luo, X.; Tang, Y. miR-155 regulates high glucose-induced cardiac fibrosis via the TGF-β signaling pathway. Mol. Biosyst. 2016, 13, 215–224. [Google Scholar] [CrossRef] [PubMed]
  72. Dagan, L.N.; Jiang, X.; Bhatt, S.; Cubedo, E.; Rajewsky, K.; Lossos, I.S. miR-155 regulates HGAL expression and increases lymphoma cell motility. Blood 2012, 119, 513–520. [Google Scholar] [CrossRef] [PubMed]
  73. Iqbal, J.; Shen, Y.; Huang, X.; Liu, Y.; Wake, L.; Liu, C.; Deffenbacher, K.; Lachel, C.M.; Wang, C.; Rohr, J.; et al. Global microRNA expression profiling uncovers molecular markers for classification and prognosis in aggressive B-cell lymphoma. Blood 2015, 125, 1137–1145. [Google Scholar] [CrossRef] [PubMed]
  74. Wang, Y.; Sun, L.E. Knockdown of LMP1-induced miR-155 sensitizes nasopharyngeal carcinoma cells to radiotherapy in vitro. Oncol. Lett. 2016, 11, 3451–3456. [Google Scholar] [CrossRef] [PubMed]
  75. Xiao, C.; Srinivasan, L.; Calado, D.; Patterson, H.; Zhang, B.; Wang, J.; Henderson, J.; Kutok, J.; Rajewsky, K. Lymphoproliferative disease and autoimmunity in mice with increased miR-17–92 expression in lymphocytes. Nat. Immunol. 2008, 9, 405–414. [Google Scholar] [CrossRef] [PubMed]
  76. Woods, K.; Thomson, J.; Hammond, S. Direct regulation of an oncogenic micro-rna cluster by E2F transcription factors. J. Biol. Chem. 2007, 282, 2130–2134. [Google Scholar] [CrossRef] [PubMed]
  77. Olive, V.; Sabio, E.; Bennett, M.J.; De Jong, C.S.; Biton, A.; McGann, J.C.; Greaney, S.K.; Sodir, N.M.; Zhou, A.Y.; Balakrishnan, A.; et al. A component of the miR-17–92 polycistronic oncomir promotes oncogene-dependent apoptosis. Elife 2013, 2, e00822. [Google Scholar] [CrossRef] [PubMed]
  78. Li, X.; Xie, W.; Xie, C.; Huang, C.; Zhu, J.; Liang, Z.; Deng, F.; Zhu, M.; Zhu, W.; Wu, R.; et al. Curcumin modulates miR-19/PTEN/AKT/p53 axis to suppress bisphenol a-induced MCF-7 breast cancer cell proliferation. Phytother. Res. 2014, 28, 1553–1560. [Google Scholar] [CrossRef] [PubMed]
  79. Rai, D.; Karanti, S.; Jung, I.; Dahia, P.; Aguiar, R. Coordinated expression of microRNA-155 and predicted target genes in diffuse large B-cell lymphoma. Cancer Genet. Cytogenet. 2008, 181, 8–15. [Google Scholar] [CrossRef] [PubMed]
  80. Gu, L.; Song, G.; Chen, L.; Nie, Z.; He, B.; Pan, Y.; Xu, Y.; Li, R.; Gao, T.; Cho, W.C.; et al. Inhibition of miR-21 induces biological and behavioral alterations in diffuse large B-cell lymphoma. Acta. Haematol. 2013, 130, 87–94. [Google Scholar] [CrossRef] [PubMed]
  81. Go, H.; Jang, J.Y.; Kim, P.J.; Kim, Y.G.; Nam, S.J.; Paik, J.H.; Kim, T.M.; Heo, D.S.; Kim, C.W.; Jeon, Y.K. MicroRNA-21 plays an oncogenic role by targeting FOXO1 and activating the PI3K/AKT pathway in diffuse large B-cell lymphoma. Oncotarget 2015, 6, 15035–15049. [Google Scholar] [CrossRef] [PubMed]
  82. Li, J.; Fu, R.; Yang, L.; Tu, W. miR-21 expression predicts prognosis in diffuse large B-cell lymphoma. Int. J. Clin. Exp. Pathol. 2015, 8, 15019–15024. [Google Scholar] [PubMed]
  83. Caivano, A.; Laurenzana, I.; De Luca, L.; La Rocca, F.; Simeon, V.; Trino, S.; D’Auria, F.; Traficante, A.; Maietti, M.; Izzo, T.; et al. High serum levels of extracellular vesicles expressing malignancy-related markers are released in patients with various types of hematological neoplastic disorders. Tumour Biol. 2015, 36, 9739–9752. [Google Scholar] [CrossRef] [PubMed]
  84. Hoshina, S.; Sekizuka, T.; Kataoka, M.; Hasegawa, H.; Hamada, H.; Kuroda, M.; Katano, H. Profile of exosomal and intracellular microrna in γ-herpesvirus-infected lymphoma cell lines. PLoS ONE 2016, 11, e0162574. [Google Scholar] [CrossRef] [PubMed]
  85. Valadi, H.; Ekström, K.; Bossios, A.; Sjöstrand, M.; Lee, J.J.; Lötvall, J.O. Exosome-mediated transfer of mRNAs and microRNAs is a novel mechanism of genetic exchange between cells. Nat. Cell Biol. 2007, 9, 654–659. [Google Scholar] [CrossRef] [PubMed]
  86. Lotvall, J.; Valadi, H. Cell to cell signalling via exosomes through esRNA. Cell Adhes. Migr. 2007, 1, 156–158. [Google Scholar] [CrossRef]
  87. Lanford, R.E.; Hildebrandt-Eriksen, E.S.; Petri, A.; Persson, R.; Lindow, M.; Munk, M.E.; Kauppinen, S.; Ørum, H. Therapeutic silencing of microRNA-122 in primates with chronic hepatitis C virus infection. Science 2010, 327, 198–201. [Google Scholar] [CrossRef] [PubMed]
  88. Janssen, H.L.; Reesink, H.W.; Lawitz, E.J.; Zeuzem, S.; Rodriguez-Torres, M.; Patel, K.; van der Meer, A.J.; Patick, A.K.; Chen, A.; Zhou, Y.; et al. Treatment of HCV infection by targeting microRNA. N. Engl. J. Med. 2013, 368, 1685–1694. [Google Scholar] [CrossRef] [PubMed]
  89. Cheng, C.J.; Bahal, R.; Babar, I.A.; Pincus, Z.; Barrera, F.; Liu, C.; Svoronos, A.; Braddock, D.T.; Glazer, P.M.; Engelman, D.M.; et al. MicroRNA silencing for cancer therapy targeted to the tumour microenvironment. Nature 2015, 518, 107–110. [Google Scholar] [CrossRef] [PubMed]
  90. Bader, A.G. miR-34—A microRNA replacement therapy is headed to the clinic. Front. Genet. 2012, 3, 120. [Google Scholar] [CrossRef] [PubMed]
Figure 1. B-cell development and miRNAs. B-cell development is regulated by multiple transcription factors and miRNAs. The expression of miRNAs differs during each stage of B-cell development, and the dysregulation of specific miRNAs impairs B-cell development, causing tumor formation.
Figure 1. B-cell development and miRNAs. B-cell development is regulated by multiple transcription factors and miRNAs. The expression of miRNAs differs during each stage of B-cell development, and the dysregulation of specific miRNAs impairs B-cell development, causing tumor formation.
Ijms 18 01495 g001

Share and Cite

MDPI and ACS Style

Kotaki, R.; Koyama-Nasu, R.; Yamakawa, N.; Kotani, A. miRNAs in Normal and Malignant Hematopoiesis. Int. J. Mol. Sci. 2017, 18, 1495. https://doi.org/10.3390/ijms18071495

AMA Style

Kotaki R, Koyama-Nasu R, Yamakawa N, Kotani A. miRNAs in Normal and Malignant Hematopoiesis. International Journal of Molecular Sciences. 2017; 18(7):1495. https://doi.org/10.3390/ijms18071495

Chicago/Turabian Style

Kotaki, Ryutaro, Ryo Koyama-Nasu, Natsuko Yamakawa, and Ai Kotani. 2017. "miRNAs in Normal and Malignant Hematopoiesis" International Journal of Molecular Sciences 18, no. 7: 1495. https://doi.org/10.3390/ijms18071495

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

Article Metrics

Back to TopTop