当前小RNA研究策略

Current strategies for microRNA research
2012-09-19 10:12点击:1402次发表评论
作者:Shuji Takada
期刊: MOD RHEUMATOL2012年9月5期22卷

Abstract  
The short ribonucleic acid molecules (20–24 nucleotides) known as microRNAs (miRNAs) are non-coding RNAs that disrupt translation or degrade target mRNAs posttranscriptionally in a sequence-specific manner, and the miRNA may not be completely complementary to its targets. This class of RNAs is thought to be functionally important because many individual miRNAs are evolutionally conserved across widely diverse phyla. Further, miRNAs are associated with diverse biological phenomena, such as cell growth, apoptosis, development, differentiation, cancer, and arthritis. MicroRNA research is one of a number of rapidly evolving fields of basic and biomedical science, with many new techniques being developed. However, miRNA experiments require modifications of preexisting molecular techniques or specialized methods, given the difficulties stemming from their small size. In this review, we summarize current in-silico and biochemical strategies for miRNA research to outline the current status of the research techniques now being employed to find a new direction in the field.

Keywords  MicroRNA – Method – Profiling – In silico


Introduction

The class of small RNA species [20–24 nucleotides (nt)] known as microRNAs (miRNAs) was discovered relatively recently. These molecules degrade or disrupt the translation of target mRNAs by binding to complementary sites in the mRNA 3′ untranslated region (UTR) [15]. The functional importance of miRNAs is suggested by evolutionary conservation across widely diverse phyla [69]. Further, miRNAs are associated with diverse biological phenomena, such as cell growth, apoptosis, development, differentiation, and tumorigenesis. For example, Lu et al. [10] have reported that miRNA expression profiles reflect the developmental lineage and differentiation state of tumors. Poorly differentiated tumors were more accurately classified by their miRNA profiles than by their mRNA profiles, suggesting that miRNA profiling may be useful in cancer diagnostics. Further, some miRNAs are associated with oncogenesis. For example, the miR-17-92 miRNA cluster is often overexpressed in B-cell lymphomas [11] and lung cancer [12]. This cluster has been identified as a potential human oncogene and named oncomiR-1 [11]. In addition, miRNAs are involved in other diseases, including nervous system disorders [1314], diabetes [1517], infectious diseases [1820], cardiovascular diseases [2123], asthma [24], autoimmune diseases [2527], and arthritis [25262829].

Here, we review miRNA biology and the methodological background of miRNA analyses.


Biogenesis, mechanisms, and functions of miRNAs

The multistep process of miRNA biogenesis begins with transcription, which is carried out primarily by RNA polymerase II to produce long transcripts [primary-miRNA (pri-miRNA)] from miRNA-encoding genes [30]. There are at least two types of miRNA promoters: some miRNAs have their own promoters, but others are located within an mRNA or intron and are transcribed as part of the host gene. Pri-miRNAs are processed in the nucleus by the RNase III enzyme Drosha, resulting in precursor molecules called precursor-miRNAs (pre-miRNAs) that are approximately 70 nt long with stem-loop structures containing 2-nt 3′ overhangs [31]. Pre-miRNAs translocate through nuclear pores to the cytoplasm via Exportin-5 [3236] and are further processed to mature miRNAs by the RNase III enzyme Dicer [3738]. Biogenesis pathways other than this canonical miRNA pathway have also been reported. Some pre-miRNAs are generated by the RNA splicing machinery and lariat-debranching enzyme (the mirtron pathway), which yield pre-miRNA-like hairpins [39]. The mirtron pathway was first discovered in Drosophila [39] and is also found in mammals [40]. The Dicer-independent miRNA biogenesis pathway is another noncanonical pathway, in which pre-miRNA is incorporated into Argonaute (Ago) complexes and cleaved by the slicer activity of Ago to generate an intermediate 3′ end, which is then further trimmed [41].

RNA-induced silencing complexes (RISCs) incorporate miRNAs, which then bind to the 3′ UTR of target mRNAs. Ago1, Ago2, Ago3, and Ago4, members of the Ago protein family involved in RISC function, play central roles in small RNA-mediated gene regulation [42]. Although recognition of mRNA targets is sequence-specific, the miRNA may not be completely complementary to its target. When the miRNA and its target sequence are 100% complementary, the mRNA target is digested by Ago2, which is the only Ago protein with slicer activity. When the miRNA and its target sequence are not 100% complementary, mRNA function is suppressed by translation inhibition or RNA instability. The seed sequence (2–8 nt from the miRNA 5′ end) is important for recognition between the miRNA and its target [43]. The seed sequence and corresponding region of its target are 100% complementary.


The role of tissue-specific miRNAs in development and pathogenesis

Many miRNAs show tissue-specific expression patterns [4446] and play a crucial role in tissue development or homeostasis [4446]. For example, targeted disruption of miR1–2, which is expressed in cardiac myocytes, causes a severe phenotype in heart development [4748]. Mice deficient forbic/miR-155 display increased lung airway remodeling [4950]. Several miRNAs have been associated with diseases, such as cancer and viral infections [51]. One of the human diseases in which miRNAs are involved is the neuropsychiatric disorder Tourette’s syndrome (TS) [52]. In some TS patients, the 3′ UTR of SLITRK1 containing a miR-189-binding site is mutated. This dysregulation of SLITRK1 by miR-189 is implicated in the onset of TS.

miRNA expression levels are also altered in primary human tumors [1053]. Differential expression patterns of miRNA profiles in various types of tumors suggest these miRNAs can be used as diagnostic markers or/and therapeutic targets [1044]. The miR-15a and miR-16-1 genes are located on the chromosome region 13q14, which is deleted in most cases of chronic lymphocytic leukemia [54]. These miRNAs target B cell lymphoma 2 (Bcl2), an antiapoptotic gene, suggesting that loss of miR-15a and miR-16-1 in B cells may lead to the inhibition of apoptosis, giving rise to malignancies [55]. miRNAs with oncogenic potential are expressed from the miR17-92 locus 13q31, which is amplified in some tumors, such as B-cell lymphoma [1156]. These miRNAs support a shift from apoptosis toward proliferation by the downregulation of E2F1. miR-372 and miR-373 are expressed specifically in testicular germ cell tumors [57]. In part, their function may be mediated by targeting the tumor suppressor LATS2 [57].


miRNAs in the immune system

There is accumulating evidence that miRNAs are important in the immune system. Using genetic deletion and transgenic approaches, miR-155 has been shown to specifically regulate T-helper cell differentiation to produce an optimal T-cell-dependent antibody response. Also, the overexpression of miR-150 in hematopoietic stem cells, followed by bone marrow transplantation, altered B-cell maturation [58].

Expression profiling of 200 miRNAs in human monocytes has revealed that several miRNAs (miR-146a/bmiR-132, and miR-155) are endotoxin-responsive genes that can also be induced by cytokines [59]. By means of promoter analysis, miR-146a was found to be a nuclear factor (NF)-κB-dependent gene [59]. Importantly, miR-146a and miR-146b were predicted to base-pair with sequences in the 3′ UTRs of the TRAF-6 and IRAK-1 genes, and these UTRs inhibited the expression of a linked reporter gene [59]. These observations suggest that miR-146 downregulates Toll-like receptor and cytokine signaling.

In addition, it has recently been reported that miR-146 plays an important role in the development and activation of T-reg, a critical regulator of autoimmune disease pathogenesis [60]. We reported that miR-146 was significantly increased in rheumatoid arthritis synovium compared with osteoarthritis synovium and normal synovium [2661]. Consistent with the potential role of miR-146 functions in macrophages to attenuate inflammatory signals, the introduction of miR-146 in vitro inhibited tumor necrosis factor (TNF)-α-dependent inflammatory responses (our preliminary data). However, it is still unknown whether the introduction of miR-146 in arthritis can diminish rheumatic disease progression and be used as a therapeutic tool.


Therapeutic potential of targeting miRNAs

Accumulating evidence for the role of miRNAs in cancer, metabolic diseases, and viral infections suggests that they could be a new class of drug targets. Antisense oligonucleotide approaches for inhibiting miRNA function and small interfering (si) RNA-like technologies for the replacement of miRNAs are currently being explored as tools for uncovering miRNA biology and as potential therapeutic agents. Therapeutic trials using silencing miRNAs have recently been conducted [62].

Decreased expression of both miR-133 and miR-1, which belong to the same transcriptional unit, has been observed in mouse models and in human cardiac hypertrophy. In vivo inhibition of miR-133 by a single infusion of an antagomir caused marked and sustained cardiac hypertrophy, suggesting its potential therapeutic application in heart disease [6263].

Recently, we have provided evidence that a cartilage-specific miRNA, “miR-140”, plays a critical role in cartilage homeostasis, and we showed that mice overexpressing miR-140 in chondrocytes were resistant to osteoarthritis changes [64]. Further research on miRNAs, such as the methods outlined below, should provide us with a strategy to utilize miRNAs as therapeutic tools in rheumatoid arthritis and osteoarthritis.


Molecular biology tools for miRNA detection
Northern blot analysis

With the following modifications, Northern blot analysis can be used to detect miRNAs: a polyacrylamide gel (approximately 15%) is used instead of an agarose gel to separate the small miRNAs; blotting is performed electrically; and 32P-labeled locked nucleic acid (LNA) probes [6566] are often used to increase sensitivity. Northern blot analysis is not very sensitive for detecting miRNAs, but provides the advantage of detecting both pre-miRNA and miRNA in a single experiment so that an abundance ratio can be determined.

Quantitative reverse transcription-polymerase chain reaction analysis
Reverse transcription polymerase chain reaction (RT-PCR) can also be used to detect mature miRNA molecules. Because miRNAs do not contain poly(A) tails at the 3′ end, alternative methods have been developed to synthesize complementary DNA (cDNA). The stem-loop RT-PCR miRNA assay is the most popular technique [67] (Fig. 1). This method uses primers that have 5′ end stem-loop structures that hybridize to the 3′ end of miRNAs. After the cDNA is synthesized, quantitative PCR is carried out using a TaqMan Life Technologies (Grand Island, NY) probe and primers designed from the miRNA and the stem-loop structure of the reverse transcription primer. This method is fast, sensitive, and easy, but can be used only to detect mature miRNA, not pre-miRNA.
MediaObjects/10165_2011_583_Fig1_HTML.gif
Fig. 1 Schematic diagram illustrating stem-loop reverse transcription polymerase chain reaction (RT-PCR). The stem-loop primer hybridizes to the 3′-terminal region of the microRNA (miRNA) to initiate the reverse transcription reaction. After the cDNA is synthesized, quantitative PCR is carried out with a TaqMan probe. Amplification of the target sequence separates the fluorophore from a quencher fluorophore, generating a fluorescent signal. F fluorescence, Q quencher

Expression profiling

There are three types of miRNA expression profiling assays, which are based on hybridization, RT-PCR, and sequencing.

Microarrays are hybridization-based assays that are widely used for miRNA profiling. Microarrays for miRNAs are supplied by several companies, including Exiqon (Vedbaek, Denmark) (miRCURY LNA miRNA array) and Agilent (Santa Clare, CA) (miRNA microarray). Although this is a simple and efficient method to obtain miRNA expression profiles, only known miRNAs can be detected.

Stem-loop RT-PCR miRNA assays are available in low-density formats array from Life Technologies Corporation. With this array, 384 miRNA RT-PCR assays can be carried out on a single plate. This method is fast, sensitive, and easy, but requires several dedicated instruments.

miRNA expression profiles can be obtained by miRNA cloning followed by a large number of sequencing reactions [66870]. This method is based on the observation that high-copy-number miRNAs are sequenced at a higher frequency than low-copy-number miRNAs. Therefore, the number of sequence reads of each miRNA can be converted into an miRNA expression profile. Capillary sequencers can be used for this technique, but next-generation sequencers have increasingly been used for this purpose. One problem is that a ubiquitously expressed miRNA has not been identified for use as an internal control, in the same way that glyceraldehyde 3-phosphate dehydrogenase and β-actin are used as mRNA internal controls. Expression profiles obtained by sequencing are generally presented as ratios (i.e., each miRNA sequence read number is divided by all miRNA sequence read numbers) (Fig. 2).
MediaObjects/10165_2011_583_Fig2_HTML.gif
Fig. 2 Estimation of miRNA expression from sequence data. In this example of a sequence-based miRNA profiling method, 10 sequence reads are obtained. The sequence reads match miRNA A (6/10), miRNA B (3/10), and miRNA C (1/10), indicating that miRNA A, miRNA B, and miRNA C comprise 60, 30, and 10% of all miRNAs, respectively

In situ hybridization

Whole-mount and section in situ hybridization are also used for miRNA detection. The procedures are similar to mRNA in situ hybridization techniques. Digoxygenin (DIG) is incorporated into the nucleic acid probe, but labeling methods for miRNA differ from those used for mRNA. To detect mRNA, DIG is incorporated into the probe as it is synthesized using RNA polymerase (internal labeling), but to detect miRNA, DIG is added to the 3′ end of an LNA-containing probe using terminal deoxynucleotidyl transferase. Although in situ hybridization is a powerful method for the identification of cellular or tissue miRNA expression, the sensitivity of this method is relatively low. To detect pri-miRNAs or pre-miRNAs, the conventional in situ hybridization method designed for mRNA can be used without modification. A detailed protocol has been published by Kloosterman et al. [71].

Sensor
To overcome the low sensitivity of in situ hybridization for miRNAs, Mansfield et al. [72] developed the miRNA sensor method, which relies on the observation that RISCs cleave mRNAs with a 3′ UTR that is 100% complementary to an miRNA in the RISC protein assembly. This method relies on a transgenic mouse that expresses a reporter (i.e., green florescence protein or LacZ regulated by a constitutive promoter such as cytomegalovirus (CMV) or CMV early enhancer/chicken beta-actin (CAG) promoter) with a 3′ UTR that is 100% complementary to the miRNA of interest (Fig. 3). The reporter is ubiquitously expressed throughout the tissues of the transgenic mouse, but is cleaved in tissues expressing the miRNA of interest. Thus, cells that do not show reporter activity can be identified as those in which the miRNA is expressed. Sensor is a sensitive method for miRNA detection but it is time-consuming because transgenic mice must be generated.
MediaObjects/10165_2011_583_Fig3_HTML.gif
Fig. 3 The sensor method. The figure depicts a transgenic mouse that expresses a LacZ reporter containing sequences complementary to the miRNA of interest. LacZ activity can be detected in cells that do not express the miRNA of interest (bottom left), but the LacZ reporter construct is cleaved in cells that express the miRNA (bottom right). The miRNA-expressing cells can be visualized as white cells by X-gal staining


Identification of miRNA targets
In-silico methods

Because mRNA targets are specifically recognized by miRNAs with both perfect and imperfect complementary matching, it is difficult to identify these targets based on sequence information alone. Web-based programs that predict targets of miRNAs include miRanda (http://www.microrna.org/microrna/home.do), PicTar [73] (http://pictar.mdc-berlin.de/cgi-bin/PicTar_vertebrate.cgi), TargetScan [74] (http://www.targetscan.org/), MicroCosm (http://www.ebi.ac.uk/enright-srv/microcosm/cgi-bin/targets/v5/search.pl) [7576], and PITA (http://genie.weizmann.ac.il/pubs/mir07/mir07_prediction.html) [77]. These online tools use methods based on combinations of target features, such as seed sequence matching and the evolutionary conservation of 3′ UTR sequences of candidate genes. This in-silico approach to miRNA target identification is easy and quick; however, it is necessary to verify candidate targets by biochemical methods and ascertain whether the miRNA and candidate targets are co-expressed in the cell type of interest.

Microarray

One function of miRNAs is reducing the stability of target mRNAs. Levels of a target mRNA can be reduced by overexpression of the cognate miRNA. Thus, an miRNA target may be detected by comparing mRNA expression profiles in cell lines overexpressing the cognate miRNA with cell lines that do not express the miRNA. However, it is still necessary to determine whether the targets are authentic using biochemical methods.

Luciferase assay

The luciferase assay is the biochemical method most widely used to identify miRNA targets. This method uses a luciferase reporter to detect inhibition of translation and mRNA degradation, but it cannot distinguish between these two processes. In-silico analysis and/or microarray assays are typically performed first to identify candidate targets of the miRNA. A complete or partial 3′ UTR sequence, or the miRNA target site is cloned into the 3′ UTR of the luciferase gene contained in a plasmid (test plasmid). A control plasmid is then constructed by generating one or two point mutations in the seed sequence of the candidate target sequence. The test plasmid or control plasmid is then transfected into a cell line with or without the miRNA overexpression vector, and luciferase assays are carried out in parallel. When the candidate is an authentic target, luciferase activity is lower in miRNA-overexpressingsss cells containing the control plasmid compared with those containing the test plasmid. An miRNA antagonist (see “Loss-of-function assays”) can be also used, but in this case, luciferase activity is higher in cells containing the control plasmid.

Argonaute HITS-CLIP
High-throughput sequencing of RNAs isolated by cross-linking and immunoprecipitation (HITS-CLIP) allows genome-wide mapping of protein interaction sites on RNA in vivo [78]. With this method, RNA and bound protein are crosslinked by ultraviolet light, the total cellular RNA undergoes partial RNase digestion, and the RNA–protein complexes are then purified by immunoprecipitation. After removing the proteins, protein interaction sites on the RNAs are sequenced using next-generation sequencers. This method is modified for miRNAs by using anti-Argonaute antibodies for the immunoprecipitation step (Argonaute HITS-CLIP) (Fig. 4) [79]. The immunoprecipitated complexes obtained by this method contain miRNAs and fragments of the target sites derived from the mRNAs. The miRNA and target fragments are isolated and sequenced with the Genome Analyzer (Illumina, San Diego, CA). Comparison of both datasets (miRNA and target sequences) enables the identification of miRNA targets.
MediaObjects/10165_2011_583_Fig4_HTML.gif
Fig. 4 Schematic representation of Argonaute-high-throughput sequencing of RNAs isolated by cross-linking and immunoprecipitation (HITS-CLIP). RNAs and proteins are cross-linked by UV irradiation followed by RNase A treatment and immunoprecipitation. Dephosphorylation at the 5′ ends is followed by 3′ adaptor ligation and kination at the 5′ ends. RNA–protein complexes are then separated by sodium dodecylsulfate-polyacrylamide gel electrophoresis (SDS-PAGE) and the isolated mRNAs and miRNAs are amplified by RT-PCR after 5′ adaptor ligation. Amplicons are sequenced by next-generation sequencers, and computer analysis identifies miRNA targets

Pulsed (p) SILAC
Stable isotope labeling using amino acids in cell culture (SILAC) enables the relative quantitation of proteins in two samples [8081]. Using this method, proteins are metabolically labeled in cells cultured in growth medium that contains radiolabeled amino acids. Labeled and unlabeled proteins are mixed, and the relative abundance of a protein can be determined by mass spectrometry. The SILAC method was used by Selbach et al. [82] and Baek et al. [83] to identify miRNA targets. The technique has been modified such that the growth medium containing the heavy isotope is added to cell culture for a very short time to identify only newly synthesized proteins; this method is called pulsed SILAC (pSILAC). Selbach et al. [82] added 13C615N4 l-arginine and 13C615N2 l-lysine (“heavy” SILAC medium) to cells as a control and 13C6-l-arginine and 40 mg/l D4-l-lysine (“medium-heavy” SILAC medium) to cells overexpressing the miRNA of interest. They obtained three signals by pSILAC from each protein species: heavy, medium-heavy, and light (unlabeled protein). The miRNA targets can be identified by the ratio of heavy and medium-heavy isotope signal intensities. The signal intensity of the medium-heavy isotopic form of the protein is decreased when the mRNA is an miRNA target, otherwise heavy and medium-heavy isotopes produce similar signal intensities.

Gain-of-function and loss-of-function experiments
Gain-of-function assays

Gain-of-function experiments are performed by transfecting a plasmid containing a constitutive promoter (e.g., CMV or CAG) to overexpress a pri-miRNA or a pre-miRNA sequence. Viral vectors can also be used, or the pre-miRNA itself can be transfected. Life Technologies Corporation offers the Pre-mi miRNA Precursor Molecule, an miRNA mimic that is chemically synthesized as a modified double-stranded oligonucleotide.

Loss-of-function assays

Loss-of-function experiments against miRNAs have also been reported. Krützfeldt et al. [84] reported the long-term specific suppression of miR-16, miR-122, miR-192, and miR-194 in liver, lung, kidney, heart, intestine, fat, skin, bone marrow, muscle, ovaries, and adrenal glands of mice by the intravenous administration of a corresponding 21- to 23-nt antagomir (antisense 2′-O-methyl oligoribonucleotide). The antagomirs also work in transient transfection of cultured cells [158588]. Elmén et al. [89] reported that liver-expressed miR-122 was antagonized in the African green monkey by a phosphate-buffered saline-formulated LNA-modified oligonucleotide administered by intravenous injection. The LNA-modified oligonucleotide is also effective in transient transfection assays in cultured cells. These miRNA inhibitors show potential for clinical applications.

Several companies offer reagents for miRNA antagonism, including Exiqon A/S (miRCURY LNA microRNA Inhibitors) and Life Technologies Corporation (Anti-miRNA miRNA Inhibitors).

Most miRNAs can be downregulated by Dicer gene knockout in embryonic stem cells. However, because mouse embryos lacking Dicer die at 7.5 days post coitum, a conditional knockout mouse can be used for tissue-specific or cell type-specific Dicer knockout experiments [90].


Conclusion

The introduction of stem-loop RT-PCR and next-generation sequencers has increased the sensitivity and dynamic range of miRNA detection and profiling. The sensor method is also sensitive, but it requires expensive and time-consuming transgenic mouse production. An in-situ method for detecting miRNA that combines easy handling with high sensitivity is needed.

One of the most difficult issues in miRNA research is the identification of miRNA targets. In-silico methods are rapid and simple, but require validation by biochemical experiments. Thus, it may be necessary to develop an easy screening assay based on biochemical methods for the identification of miRNA targets.

Results of miRNA studies show promise for clinical applications, such as miRNA detection for diagnostics and gain-of-function/loss-of-function therapy. Therefore, the continuing development and improvement of miRNA techniques is crucial.

Acknowledgments  
This work was supported in part by Grants-in-Aid for Scientific Research (C) (23570265) from the Ministry of Education, Culture, Sports, Science, and Technology of Japan (to S.T.) and by CREST (Core Research for Evolutional Science and Technology) from the Japan Science and Technology Corporation (JST) (to H.A.).
Conflict of interest  
None.


References

1. Filipowicz W. RNAi: the nuts and bolts of the RISC machine. Cell. 2005;122:17–20.
PubMed CrossRef ChemPort
 
2. Guo H, Ingolia NT, Weissman JS, Bartel DP. Mammalian microRNAs predominantly act to decrease target mRNA levels. Nature. 2010;466:835–40.
PubMed CrossRef ChemPort
 
3. Hammond SM. Dicing and slicing: the core machinery of the RNA interference pathway. FEBS Lett. 2005;579:5822–9.
PubMed CrossRef ChemPort
 
4. Hannon GJ. RNA interference. Nature. 2002;418:244–51.
PubMed CrossRef ChemPort
 
5. Mattick JS, Makunin IV. Small regulatory RNAs in mammals. Hum Mol Genet. 2005;14:R121–32.
PubMed CrossRef ChemPort
 
6. Lagos-Quintana M, Rauhut R, Lendeckel W, Tuschl T. Identification of novel genes coding for small expressed RNAs. Science. 2001;294:853–8.
PubMed CrossRef ChemPort
 
7. Lau NC, Lim LP, Weinstein EG, Bartel DP. An abundant class of tiny RNAs with probable regulatory roles in Caenorhabditis elegans. Science. 2001;294:858–62.
PubMed CrossRef ChemPort
 
8. Lee RC, Ambros V. An extensive class of small RNAs in Caenorhabditis elegans. Science. 2001;294:862–4.
PubMed CrossRef ChemPort
 
9. Pasquinelli AE, Reinhart BJ, Slack F, Martindale MQ, Kuroda MI, Maller B, Hayward DC, Ball EE, Degnan B, Müller P, Spring J, Srinivasan A, Fishman M, Finnerty J, Corbo J, Levine M, Leahy P, Davidson E, Ruvkun G (2000) Conservation of the sequence and temporal expression of let-7 heterochronic regulatory RNA. Nature 408:86–89.
 
10. Lu J, Getz G, Miska EA, Alvarez-Saavedra E, Lamb J, Peck D, Sweet-Cordero A, Ebert BL, Mak RH, Ferrando AA, Downing JR, Jacks T, Horvitz HR, Golub TR. MicroRNA expression profiles classify human cancers. Nature. 2005;435:834–8.
PubMed CrossRef ChemPort
 
11. He L, Thomson JM, Hemann MT, Hernando-Monge E, Mu D, Goodson S, Powers S, Cordon-Cardo C, Lowe SW, Hannon GJ, Hammond SM. A microRNA polycistron as a potential human oncogene. Nature. 2005;435:828–33.
PubMed CrossRef ChemPort
 
12. Hayashita Y, Osada H, Tatematsu Y, Yamada H, Yanagisawa K, Tomida S, Yatabe Y, Kawahara K, Sekido Y, Takahashi T. A polycistronic microRNA cluster, miR-17-92, is overexpressed in human lung cancers and enhances cell proliferation. Cancer Res. 2005;65:9628–32.
PubMed CrossRef ChemPort
 
13. Hébert SS, Horré K, Nicolaï L, Papadopoulou AS, Mandemakers W, Silahtaroglu AN, Kauppinen S, Delacourte A, De Strooper B. Loss of microRNA cluster miR-29a/b-1 in sporadic Alzheimer’s disease correlates with increased BACE1/beta-secretase expression. Proc Natl Acad Sci USA. 2008;105:6415–20.
PubMed CrossRef
 
14. Saba R, Goodman CD, Huzarewich RL, Robertson C, Booth SA (2008) A miRNA signature of prion induced neurodegeneration. PLoS ONE 3e:3652.
 
15. Poy MN, Eliasson L, Krutzfeldt J, Kuwajima S, Ma X, Macdonald PE, Pfeffer S, Tuschl T, Rajewsky N, Rorsman P, Stoffel M. A pancreatic islet-specific microRNA regulates insulin secretion. Nature. 2004;432:226–30.
PubMed CrossRef ChemPort
 
16. Tang X, Tang G, Ozcan S. Role of microRNAs in diabetes. Biochim Biophys Acta. 2008;1779:697–701.
PubMed CrossRef ChemPort
 
17. Lovis P, Roggli E, Laybutt DR, Gattesco S, Yang JY, Widmann C, Abderrahmani A, Regazzi R. Alterations in microRNA expression contribute to fatty acid-induced pancreatic beta-cell dysfunction. Diabetes. 2008;57:2728–36.
PubMed CrossRef ChemPort
 
18. Pfeffer S, Zavolan M, Grässer FA, Chien M, Russo JJ, Ju J, John B, Enright AJ, Marks D, Sander C, Tuschl T. Identification of virus-encoded microRNAs. Science. 2004;304:734–6.
PubMed CrossRef ChemPort
 
19. Pfeffer S, Sewer A, Lagos-Quintana M, Sheridan R, Sander C, Grässer FA, van Dyk LF, Ho CK, Shuman S, Chien M, Russo JJ, Ju J, Randall G, Lindenbach BD, Rice CM, Simon V, Ho DD, Zavolan M, Tuschl T. Identification of microRNAs of the herpesvirus family. Nat Methods. 2005;2:269–76.
PubMed CrossRef ChemPort
 
20. Ghosh Z, Mallick B, Chakrabarti J. Cellular versus viral microRNAs in host–virus interaction. Nucleic Acids Res. 2009;37:1035–48.
PubMed CrossRef ChemPort
 
21. Sayed D, Hong C, Chen IY, Lypowy J, Abdellatif M. MicroRNAs play an essential role in the development of cardiac hypertrophy. Circ Res. 2007;100:416–24.
PubMed CrossRef ChemPort
 
22. Thum T, Gross C, Fiedler J, Fischer T, Kissler S, Bussen M, Galuppo P, Just S, Rottbauer W, Frantz S, Castoldi M, Soutschek J, Koteliansky V, Rosenwald A, Basson MA, Licht JD, Pena JT, Rouhanifard SH, Muckenthaler MU, Tuschl T, Martin GR, Bauersachs J, Engelhardt S. MicroRNA-21 contributes to myocardial disease by stimulating MAP kinase signalling in fibroblasts. Nature. 2008;456:980–4.
PubMed CrossRef ChemPort
 
23. Morton SU, Scherz PJ, Cordes KR, Ivey KN, Stainier DY, Srivastava D. microRNA-138 modulates cardiac patterning during embryonic development. Proc Natl Acad Sci USA. 2008;105:17830–5.
PubMed CrossRef ChemPort
 
24. Tan Z, Randall G, Fan J, Camoretti-Mercado B, Brockman-Schneider R, Pan L, Solway J, Gern JE, Lemanske RF, Nicolae D, Ober C. Allele-specific targeting of microRNAs to HLA-G and risk of asthma. Am J Hum Genet. 2007;81:829–34.
PubMed CrossRef ChemPort
 
25. Stanczyk J, Pedrioli DM, Brentano F, Sanchez-Pernaute O, Kolling C, Gay RE, Detmar M, Gay S, Kyburz D. Altered expression of microRNA in synovial fibroblasts and synovial tissue in rheumatoid arthritis. Arthritis Rheum. 2008;58:1001–9.
PubMed CrossRef
 
26. Nakasa T, Miyaki S, Okubo A, Hashimoto M, Nishida K, Ochi M, Asahara H. Expression of microRNA-146 in rheumatoid arthritis synovial tissue. Arthritis Rheum. 2008;58:1284–92.
PubMed CrossRef ChemPort
 
27. Dai Y, Huang YS, Tang M, Lv TY, Hu CX, Tan YH, Xu ZM, Yin YB. Microarray analysis of microRNA expression in peripheral blood cells of systemic lupus erythematosus patients. Lupus. 2007;16:939–46.
PubMed CrossRef ChemPort
 
28. Miyaki S, Nakasa T, Otsuki S, Grogan SP, Higashiyama R, Inoue A, et al. MicroRNA-140 is expressed in differentiated human articular chondrocytes and modulates interleukin-1 responses. Arthritis Rheum. 2009;60:2723–30.
PubMed CrossRef ChemPort
 
29. Stanczyk J, Ospelt C, Karouzakis E, Filer A, Raza K, Kolling C, et al. Altered expression of microRNA-203 in rheumatoid arthritis synovial fibroblasts and its role in fibroblast activation. Arthritis Rheum. 2011;63:373–81.
PubMed CrossRef
 
30. Lee Y, Kim M, Han J, Yeom KH, Lee S, Baek SH, Kim VN. MicroRNA genes are transcribed by RNA polymerase II. EMBO J. 2004;23:4051–60.
PubMed CrossRef ChemPort
 
31. Lee Y, Ahn C, Han J, Choi H, Kim J, Yim J, Lee J, Provost P, Rådmark O, Kim S, Kim VN. The nuclear RNase III Drosha initiates microRNA processing. Nature. 2003;425:415–9.
PubMed CrossRef ChemPort
 
32. Bohnsack MT, Czaplinski K, Gorlich D. Exportin 5 is a RanGTP-dependent dsRNA-binding protein that mediates nuclear export of pre-miRNAs. RNA. 2004;10:185–91.
PubMed CrossRef ChemPort
 
33. Brownawell AM, Macara IG. Exportin-5, a novel karyopherin, mediates nuclear export of double-stranded RNA binding proteins. J Cell Biol. 2002;156:53–64.
PubMed CrossRef ChemPort
 
34. Lund E, Güttinger S, Calado A, Dahlberg JE, Kutay U. Nuclear export of microRNA precursors. Science. 2004;303:95–8.
PubMed CrossRef ChemPort
 
35. Shibata S, Sasaki M, Miki T, Shimamoto A, Furuichi Y, Katahira J, Yoneda Y. Exportin-5 orthologues are functionally divergent among species. Nucleic Acids Res. 2006;34:4711–21.
PubMed CrossRef ChemPort
 
36. Yi R, Qin Y, Macara IG, Cullen BR. Exportin-5 mediates the nuclear export of pre-microRNAs and short hairpin RNAs. Genes Dev. 2003;17:3011–6.
PubMed CrossRef ChemPort
 
37. Hutvagner G, Simard MJ, Mello CC, Zamore PD. Sequence-specific inhibition of small RNA function. PLoS Biol. 2004;2:E98.
PubMed CrossRef
 
38. Ketting RF, Fischer SE, Bernstein E, Sijen T, Hannon GJ, Plasterk RH. Dicer functions in RNA interference and in synthesis of small RNA involved in developmental timing in C. elegans. Genes Dev. 2001;15:2654–9.
PubMed CrossRef ChemPort
 
39. Okamura K, Hagen JW, Duan H, Tyler DM, Lai EC. The mirtron pathway generates microRNA-class regulatory RNAs in Drosophila. Cell. 2007;130:89–100.
PubMed CrossRef ChemPort
 
40. Berezikov E, Chung WJ, Willis J, Cuppen E, Lai EC. Mammalian mirtron genes. Mol Cell. 2007;28:328–36.
PubMed CrossRef ChemPort
 
41. Cheloufi S, Dos Santos CO, Chong MM, Hannon GJ. A dicer-independent miRNA biogenesis pathway that requires Ago catalysis. Nature. 2010;465:584–9.
PubMed CrossRef ChemPort
 
42. Parker JS, Barford D. Argonaute: a scaffold for the function of short regulatory RNAs. Trends Biochem Sci. 2006;31:622–30.
PubMed CrossRef ChemPort
 
43. Lim LP, Lau NC, Garrett-Engele P, Grimson A, Schelter JM, Castle J, Bartel DP, Linsley PS, Johnson JM. Microarray analysis shows that some microRNAs downregulate large numbers of target mRNAs. Nature. 2005;433:769–73.
PubMed CrossRef ChemPort
 
44. Calin GA, Croce CM. MicroRNA signatures in human cancers. Nat Rev Cancer. 2006;6:857–66.
PubMed CrossRef ChemPort
 
45. Bartel DP. MicroRNAs: genomics, biogenesis, mechanism, and function. Cell. 2004;116:281–97.
PubMed CrossRef ChemPort
 
46. Farh KK, et al. The widespread impact of mammalian microRNAs on mRNA repression and evolution. Science. 2005;310:1817–21.
PubMed CrossRef ChemPort
 
47. Zhao Y, et al. Dysregulation of cardiogenesis, cardiac conduction, and cell cycle in mice lacking miRNA-1-2. Cell. 2007;129:303–17.
PubMed CrossRef ChemPort
 
48. van Rooij E, et al. Control of stress-dependent cardiac growth and gene expression by a microRNA. Science. 2007;316:575–9.
PubMed CrossRef
 
49. Thai TH, et al. Regulation of the germinal center response by microRNA-155. Science. 2007;316:604–8.
PubMed CrossRef ChemPort
 
50. Rodriguez A, et al. Requirement of bic/microRNA-155 for normal immune function. Science. 2007;316:608–11.
PubMed CrossRef ChemPort
 
51. Lecellier CH, et al. A cellular microRNA mediates antiviral defense in human cells. Science. 2005;308:557–60.
PubMed CrossRef ChemPort
 
52. Abelson JF, et al. Sequence variants in SLITRK1 are associated with Tourette’s syndrome. Science. 2005;310:317–20.
PubMed CrossRef ChemPort
 
53. Calin GA, et al. MicroRNA profiling reveals distinct signatures in B cell chronic lymphocytic leukemias. Proc Natl Acad Sci USA. 2004;101:11755–60.
PubMed CrossRef ChemPort
 
54. Calin GA, 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;99:15524–9.
PubMed CrossRef ChemPort
 
55. Cimmino A, et al. miR-15 and miR-16 induce apoptosis by targeting BCL2. Proc Natl Acad Sci USA. 2005;102:13944–9.
PubMed CrossRef ChemPort
 
56. Dews M, et al. Augmentation of tumor angiogenesis by a Myc-activated microRNA cluster. Nat Genet. 2006;38:1060–5.
PubMed CrossRef ChemPort
 
57. Voorhoeve PM, et al. A genetic screen implicates miRNA-372 and miRNA-373 as oncogenes in testicular germ cell tumors. Cell. 2006;124:1169–81.
PubMed CrossRef ChemPort
 
58. Zhou B, et al. miR-150, a microRNA expressed in mature B and T cells, blocks early B cell development when expressed prematurely. Proc Natl Acad Sci USA. 2007;104:7080–5.
PubMed CrossRef ChemPort
 
59. Taganov KD, et al. NF-kappaB-dependent induction of microRNA miR-146, an inhibitor targeted to signaling proteins of innate immune responses. Proc Natl Acad Sci USA. 2006;103:12481–6.
PubMed CrossRef ChemPort
 
60. Lu LF, et al. Function of miR-146a in controlling Treg cell-mediated regulation of Th1 responses. Cell. 2010;142:914–29.
PubMed CrossRef ChemPort
 
61. Yamasaki K, et al. Expression of microRNA-146a in osteoarthritis cartilage. Arthritis Rheum. 2009;60:1035–41.
PubMed CrossRef ChemPort
 
62. Yang B, et al. The muscle-specific microRNA miR-1 regulates cardiac arrhythmogenic potential by targeting GJA1 and KCNJ2. Nat Med. 2007;13:486–91.
PubMed CrossRef ChemPort
 
63. Care A, et al. MicroRNA-133 controls cardiac hypertrophy. Nat Med. 2007;13:613–8.
PubMed CrossRef ChemPort
 
64. Miyaki S, Sato T, Inoue A, Otsuki S, Ito Y, Yokoyama S, et al. MicroRNA-140 plays dual roles in both cartilage development and homeostasis. Genes Dev. 2010;24:1173–85.
PubMed CrossRef ChemPort
 
65. Várallyay E, Burgyán J, Havelda Z. MicroRNA detection by northern blotting using locked nucleic acid probes. Nat Protoc. 2008;3:190–6.
PubMed CrossRef
 
66. Válóczi A, Hornyik C, Varga N, Burgyán J, Kauppinen S, Havelda Z. Sensitive and specific detection of microRNAs by northern blot analysis using LNA-modified oligonucleotide probes. Nucleic Acids Res. 2004;32:e175.
PubMed CrossRef
 
67. Varkonyi-Gasic E, Hellens RP. Quantitative stem-loop RT-PCR for detection of microRNAs. Methods Mol Biol. 2011;744:145–57.
PubMed SpringerLink ChemPort
 
68. Takada S, Berezikov E, Yamashita Y, Lagos-Quintana M, Kloosterman WP, Enomoto M, Hatanaka H, Fujiwara S, Watanabe H, Soda M, Choi YL, Plasterk RH, Cuppen E, Mano H. Mouse microRNA profiles determined with a new and sensitive cloning method. Nucleic Acids Res. 2006;34:e115.
PubMed CrossRef
 
69. Takada S, Mano H. Profiling of microRNA expression by mRAP. Nat Protoc. 2007;2:3136–45.
PubMed CrossRef ChemPort
 
70. Mano H, Takada S. mRAP, a sensitive method for determination of microRNA expression profiles. Methods. 2007;3:118–22.
CrossRef
 
71. Kloosterman WP, Wienholds E, de Bruijn E, Kauppinen S, Plasterk RH. In situ detection of miRNAs in animal embryos using LNA-modified oligonucleotide probes. Nat Methods. 2006;3:27–9.
PubMed CrossRef ChemPort
 
72. Mansfield JH, Harfe BD, Nissen R, Obenauer J, Srineel J, Chaudhuri A, Farzan-Kashani R, Zuker M, Pasquinelli AE, Ruvkun G, Sharp PA, Tabin CJ, McManus MT. MicroRNA-responsive ‘sensor’ transgenes uncover Hox-like and other developmentally regulated patterns of vertebrate microRNA expression. Nat Genet. 2004;36:1079–83.
PubMed CrossRef ChemPort
 
73. Krek A, Grün D, Poy MN, Wolf R, Rosenberg L, Epstein EJ, MacMenamin P, da Piedade I, Gunsalus KC, Stoffel M, Rajewsky N. Combinatorial microRNA target predictions. Nat Genet. 2005;37:495–500.
PubMed CrossRef ChemPort
 
74. Lewis BP, Shih IH, Jones-Rhoades MW, Bartel DP, Burge CB. Prediction of mammalian microRNA targets. Cell. 2003;115:787–98.
PubMed CrossRef ChemPort
 
75. Enright AJ, John B, Gaul U, Tuschl T, Sander C, Marks DS. MicroRNA targets in Drosophila. Genome Biol. 2003;5:R1.
PubMed CrossRef
 
76. Griffiths-Jones S, Saini HK, van Dongen S, Enright AJ. miRBase: tools for microRNA genomics. Nucleic Acids Res. 2008;36(Database issue):D154–8.
PubMed ChemPort
 
77. Kertesz M, Iovino N, Unnerstall U, Gaul U, Segal E. The role of site accessibility in microRNA target recognition. Nat Genet. 2007;39:1278–84.
PubMed CrossRef ChemPort
 
78. Licatalosi DD, Mele A, Fak JJ, Ule J, Kayikci M, Chi SW, Clark TA, Schweitzer AC, Blume JE, Wang X, Darnell JC, Darnell RB. HITS-CLIP yields genome-wide insights into brain alternative RNA processing. Nature. 2008;456:464–9.
PubMed CrossRef ChemPort
 
79. Chi SW, Zang JB, Mele A, Darnell RB. Argonaute HITS-CLIP decodes microRNA–mRNA interaction maps. Nature. 2009;460:479–86.
PubMed ChemPort
 
80. Mann M. Functional and quantitative proteomics using SILAC. Nat Rev Mol Cell Biol. 2006;7:952–8.
PubMed CrossRef ChemPort
 
81. Ong SE, Blagoev B, Kratchmarova I, Kristensen DB, Steen H, Pandey A, Mann M. Stable isotope labeling by amino acids in cell culture, SILAC, as a simple and accurate approach to expression proteomics. Mol Cell Proteomics. 2002;1:376–86.
PubMed CrossRef ChemPort
 
82. Selbach M, Schwanhäusser B, Thierfelder N, Fang Z, Khanin R, Rajewsky N. Widespread changes in protein synthesis induced by microRNAs. Nature. 2008;455:58–63.
PubMed CrossRef ChemPort
 
83. Baek D, Villéén J, Shin C, Camargo FD, Gygi SP, Bartel DP. The impact of microRNAs on protein output. Nature. 2008;455:64–71.
PubMed CrossRef ChemPort
 
84. Krützfeldt J, Rajewsky N, Braich R, Rajeev KG, Tuschl T, Manoharan M, Stoffel M. Silencing of microRNAs in vivo with ‘antagomirs’. Nature. 2005;438:685–9.
PubMed CrossRef
 
85. Chen CZ, Lodish HF. MicroRNAs as regulators of mammalian hematopoiesis. Semin Immunol. 2005;17:155–65.
PubMed CrossRef ChemPort
 
86. Hutvágner G, McLachlan J, Pasquinelli AE, Bálint E, Tuschl T, Zamore PD. A cellular function for the RNA-interference enzyme Dicer in the maturation of the let-7 small temporal RNA. Science. 2001;293:834–8.
PubMed CrossRef
 
87. Meister G, Landthaler M, Dorsett Y, Tuschl T. Sequence-specific inhibition of microRNA- and siRNA-induced RNA silencing. RNA. 2004;10:544–50.
PubMed CrossRef ChemPort
 
88. Zhao Y, Samal E, Srivastava D. Serum response factor regulates a muscle-specific microRNA that targets Hand2 during cardiogenesis. Nature. 2005;436:214–20.
PubMed CrossRef ChemPort
 
89. Elmén J, Lindow M, Schütz S, Lawrence M, Petri A, Obad S, Lindholm M, Hedtjärn M, Hansen HF, Berger U, Gullans S, Kearney P, Sarnow P, Straarup EM, Kauppinen S. LNA-mediated microRNA silencing in non-human primates. Nature. 2008;452:896–9.
PubMed CrossRef
 
90. Harfe BD, McManus MT, Mansfield JH, Hornstein E, Tabin CJ. The RNaseIII enzyme Dicer is required for morphogenesis but not patterning of the vertebrate limb. Proc Natl Acad Sci USA. 2005;102:10898–903.
PubMed CrossRef ChemPort

学科代码:风湿病学   关键词:MicroRNA Method Profiling In silico
来源: http://www.elseviermed.cn/journal/detail/mod_rheum
顶一下(0
您可能感兴趣的文章
    发表评论网友评论(0)
      发表评论
      登录后方可发表评论,点击此处登录