Change: miRNAs Control Aging

created on Nov. 17, 2012, 10:23 p.m. by Hevok & updated on Dec. 2, 2012, 2:52 a.m. by Hevok

======================== miRNAs Controlling Aging ======================== :Abstract: Gene silencing via RNA interference is mediated by microRNAs which are major controller of developmental processes and its plasticity conferred by diet. The function of specific miRNAs is going beyond the developmental period and extends in the control of lifespan via regulating the onset and the rate of aging.

.. contents:: Contents

.. header:: miRNAs Control Aging_

.. _miRNAs Control Aging: http://denigma.de/articles/miRNAs_Control_Aging

.. footer:: ###Page###

Introduction

Non-coding RNAs (ncRNA) do not encode a protein gene product but functional RNA. ncRNA can be very long or extremely short. Among the former which are less then 200 nucleotides long are microRNA, short interfering RNA (siRNA), Piwi-interacting RNAs (piRNAs) and small nucleolar RNAs (snoRNAs).

MicroRNAs (miRNAs) are endogenous non-coding RNAs that regulate gene expression through translation modulation, RNA degradation or induction of heterochromatin. miRNA can mediate either repression [19239886] or induction [18498749; 18048652] of translation depending on the context of the miRNA binding site. Silencing occurs via near-perfect or partially complementary to target mRNA. The miRNA seed sequence comprise 2-7 bases and is important for recognition.

miRNA regulate developmental timing, neuronal asymmetry, germline cell division, reprogramming to induced pluripotent stem cells, p53-induced cell senescence and cancer progression. miRNAs act in regulatory pathways that extend and reduce lifespan. Heterochronic development circuit miRNAs lin-4 and lin-14 for example regulate longevity post-developmentally. miRNAs that change with age in C. elegans, mice and humans were identified by microarrays and RNA sequencing. miRNA expression in general declines with aging but in some tissues it appears to be the opposite. Some groups of the small RNAs exhibit particular large expression changes and some are consistently up or downregulated with age. For instance, let-7 is associated with larval development and cancer had the greatest decrease with age. Many of the age-regulated miRNAs are expressed primarily in the intestine, neurons, and somatic gonad, all tissues associated with regulation of aging.

Non-coding RNAs Modulate Senescence, Ageing and Longevity

Small non-coding RNAs (ncRNAs) compromise classes such as endogenous small inhibitory RNAs (siRNAs), PIWI-interacting RNAs (piRNAs), QDE1-interacting RNAs (qiRNAs) and microRNAs (miRNAs). Some ncRNAs have not yet been identified in mammals and are dependent on RNA-dependent RNA polymerases. The mammalian enzyme consists of hTERT together with RMPR, an alternative RNA component. miRNAs are regulated during cellular senescence. miRNAs contribute to tissue regeneration by regulation of stem cell function. At least one miRNA modules C. elegans lifespan. miRNAs act as inhibitors of proteins mediating the insulin/IGF1 and TOR signaling [20080172].

miRNA Regulate Developmental Processes

miRNA regulate developmental decisions in C. elegans and are well conserved in plants and animals. miRNAs play an important role in regulating embryonic development also in vertebrates [He et al., 2011]. Specific non-coding RNA (ncRNA) circuitry mediates brain development, plasticity, stress response as well as aging [Qureshi and Mehler, 2011]. Developmental progression is coordinated epigenetically and affects the proliferation status of cells. For instance, mir-34b/c are able to modulate CpG methylation (differentiation) and interact with p53 signaling (cell proliferation control). microRNAs are mediators of life history changes which determine whether an animal ages or stays in a non-aging state [21343388].

Disrupting General RNAi Decreases Lifespan

Adult-specific knockdown of the C. elegans argonaute-like gene 1 alg-1 results in shortened lifespan. Similar observations were made in fruit fly. In Drosophila the endo-siRNA pathway regulates metabolism, defence against stress and aging. Dicer-2 (Dcr-2) loss changes the expression of mostly metabolic genes implicated in stress resistance and aging. Dcr-2 mutants had reduced lifespan and were hypersensitive to oxidative, endoplasmic reticulum, starvation and cold stresses as well as abnormal lipid and carbohydrate metabolism [Lim et al., 2011].

Manipulating Specific miRNAs Impacts on Lifespan

In C. elegans miRNA lin-4 and its target gene lin-14, which encodes a transcription factor, control larval developmental timing. Overexpression of lin-4 miRNA or suppression of lin-14 extends lifespan [Boehm and Slack, 2005]. In line with this, lin-4 loss-of-function mutation or lin-14 gain-of-function mutant (lacks lin-4 binding site) leads to decreased lifespan. Knockdown of lin-14 only during adulthood is sufficient to extend lifespan and suppresses the short lifespan phenotype of lin-4 mutants. lin-4 mutants display an increased rate of aging, while lin-14 mutants exhibit retarded aging. Lifespan extension conferred by lin-14 reduction is mediated by DAF-16 and HSF-1. lin-4 is itself regulated by DAF-16 in L1 arrest.

Specific miRNA in C. elegans can promote or antagonize aging. While mutating miR-71, miR-238 and miR-246 decreases lifespan, the overexpression of miR-71 or miR-246 increase lifespan. Whereas miR-239 mutants have a shortened lifespan, overexpression of miR-239 blesses the animal with extended longevity. Lifespan extension by mir-239 loss-of-function requires daf-16 [de Lencastre et al., 2010].

mir-58(n4640), miR-61 mir-250(nDf59), mir-64-66 mir229(nDf63), miR-71(mir-71(n4115), mir-80 mir-227(nDf53; mir81-82(nDf54) and mir-238 significantly shorten the lifespan if knocked down [Boulias et al. 2012] (at 20 degree Celsius; maybe also miR-49).

pash-1(mj100) loss-of-function mutation deactivates miRNA biogenesis and decreases the lifespan [23097426].

In Drosophila mutating miR-14 (which is a cell death suppressor) reduces lifespan [Xu et al., 2003].

Differential Expression of miRNA during Aging

Specific miRNAs are altered in expression during aging [de Lencastre et al., 2010; Ibanez-Ventoso and Driscoll, 2009; Kato et al., 2011].

In C. elegans more than 34 miRNAs exhibit changes in expression during adulthood (31 > two-fold) [Ibanez-Ventoso et al., 2006]. In whole body of worms there is an overall decline in miRNA expression [de Lencastre et al., 2010; Ibanez-Ventoso and Driscoll, 2009; Ibanez-Ventoso et al., 2006]. Expression patterns of developmental timing regulators lin-4 and let-7 miRNAs as well as conserved muscle miRNA miR-1 are regulated also during the adulthood [Ibanez-Ventoso et al., 2006]. In whole body of worms there is an overall decline in miRNA expression [de Lencastre et al., 2010; Ibanez-Ventoso and Driscoll, 2009].

Similar in mammals the majority of miRNA decrease in abundance with aging in brain and peripheral blood monocular cells [Noren Hooten et al., 2010; Persengiev et al., 2011]. However, in specific somatic tissues such as liver there were more upregulated than downregulated miRNAs during aging [Li et al., 2011a; Maes et al., 2008].

In livers of aged mice (4 vs. 33 month) four miRNA (miR-93, miR-669c, miR-214 and miR-709) were especially upregulated and correlated with the protein expression of their target genes which are associated to mitochondrial function, oxidative stress and proliferation [Maes et al., 2008]. miR-669c and miR-709 levels increase during mid-age (18 to 33 month) in murine liver. miR-93 and miR-214 are higher in extreme old (33 vs. 4 or 10 month) mice [Maes et al., 2008]. miR-93, miR-214 and miR669c all target detoxification and regeneration genes which decline in activity with age [Maes et al., 2008]. miR93, miR-214 and miR-709 increase with age and target UQCRC1 (a component of the cytochrome c complex) which also declines in older liver. miR-34a and miR-93 expression increase with age and target MGST1 and SIRT1 as well as SP1 and NRF2 (both are transcription factors targeting MGST1 and SIRT1), which are all becoming downregulated during aging [Li et al., 2011b].

About 70 miRNAs are upregulated during mid-adulthood (18 month) in murine brain. Some of them are exactly the same which are also upregulated in liver [Li et al., 2011a]. Specifically upregulated in brain were miR-22, miR-101a, miR-720 and miR-721 [Li et al., 2011a]. 27 of the 70 miRNAs have predicted target gene encoding components of the mitochondrial transport chain and F1F0-ATPase, which decrease which aging and therefore explaining the decreased respiration rates occurring during aging [Li et al., 2011a].

Aging in murine muscle leads to differentially expression of at least 67 miRNA including miR-221 which regulates myogenic precursor [Hamrick et al., 2010]. miR-7, miR-468, miR-542 and miR-698 were most substantially increased, while miR-124a, miR-181a, miR-221, miR-382, miR-434 and miR-455 levels were most strongly decreased. Treatment with the hunger hormone leptin reversed a considerable number of these changes. Also in humans at least 18 miRNAs are differentially expressed during muscle aging (31 vs. 73 years). let-7b and let-7e are increased in expression in skeletal muscle during human aging. These let-7 family members repress cell cycle regulators CDK6, CDK25A and CDC34 as well as PAX7, crucial for satellite cell turnover [Drummond et al., 2011].

pri-miRNA-1-1, -1-2, 133a-1 and -133a-2 expression were elevated in skeletal muscle of older humans [Drummond et al., 2008].

Numerous miRNAs are commonly downregulated and a few are commonly upregulated during aging in the cortex and cerebellum of humans, chimpanzees and macaques. For instance, among those is miR-144 that targets ataxin-1 which is also increase in all three species during aging [Persengiev et al., 2011].

In human peripheral blood mononuclear at least 9 miRNA (miR-103, miR-107, miR-128, miR-130a, miR-155, miR-24, miR-221, miR-496 and miR-1538) were lower in older individuals. Their predicted target genes such as PI3K, c-Kit and H2AX are elevated with advanced age. Decreasing expression of miR-221 was sufficient to cause increase in its target gene PI3K which in nematodes if knockout extends lifespan by ten-fold.

Peripheral blood of adult woman (22 - 25 vs. 36 - 39 years) exhibits significantly distinct patterns of miRNAs, but mRNA gene expression. However there exists a weak global correlation between these two types of expression levels. Most differentially expressed miRNAs include miR-155, -18a, -142, -340, -363, -195, and -24 and there target genes are enriched for estrogen regulation [Sredni et al., 2011].

Deep sequencing identified miRNAs that change with age (young; day 0 of adulthood vs. middle-aged; day 10) in wild-type and long-lived daf-2 mutants. 11 were new miRNAs, several of which share homology with miRNAs in higher eukaryotes.

Although in general it appears the miRNA expression gradually declines with aging, some miRNA increase with age in expression levels, such as miR-144 in chimpanzees and humans as well as let-7 and miR-34a in mouse [Inukai et al. 2012; Persengiev et al. 2011; Li et al. 2011].

miR-145 and miR-375 decline in mouse brain during aging and have target genes involved in insulin signaling [Noren Hooten et al. 2012; Inukai et al. 2012].

Declining miRNAs in human blood are associated with the loss of repression of cancer-associated genes as they target p53 signaling, citrat cycle and other cancer-related processes [Elsharawy et al. 2012].

Age-related miRNA dysregulation occurs in groups

It seems like that miRNAs differentially expressed with age clustered, which may be duo to shared promoter sequences or genomic location [Elsharawy et al. 2012]. miRNA appear to act in concert and are therefor regulated in packs [Lanceta et al. 2010].

During aging miRNA expression dysregulation generally occurs in groups, indicating that they are functionally coordinated as a ‘pack’ by common transcriptional regulators. Many miRNAs are coordinated in their expression, due to either close proximity of their genomic location or sharing the same transcriptional regulation [Lanceta et al., 2010].

miRNA control mid-life decline

Decline in cellular health status occurs at mid-life. This decline may involve a universal or system-specific programmatic shift of signaling control [Wang, 2007].

Correlation Between Differential Expression and Lifespan Impact

Some of the miRNAs upregulated with age regulate longevity. Three miRNAs with large expression increases with age affected lifespan. miR-239 that increased longevity (i.e. positive gerontogene) and stress resistance, while its overexpression shortens lifespan (negative gerontogene). miR-71 and miR-246 mutants had decreased lifespan, while overexpression increased lifespan (positive and negative ageing-suppressors). miR-239 and miR-71 likely function in IIS (insulin/insulin-like growth factor signaling), because miR-239-mediated effects on longevity are daf-16-dependent and loss of daf-16 does not further shorten miR-71’s lifespan. pdk-1 (IIS) and cdc-25.1 (cell cycle checkpoint) had predicted miRNA-binding sites in their 3’UTR and their expression is altered in the miR-71 mutant, thus function as a link between IIS and DNA-damage checkpoint longevity pathways [de Lencastre et al., 2010]. The miRNA mutants did not appear to have obvious changes in developmental rates, reproductive timing, or progeny production, while longevity regulators like daf-2, eat-2 and mitochondrial mutants had obvious developmental phenotypes. miR-71 is upregulated during starvation, induced by diapause and during post-dauer recovery as well as during starvation-induced early dormancy [Karp et al., 2011]. Both miR-71 and miR-238 are upregulated during starvation-induced early dormancy. miR-71 is actually required for dauer recovery by suppressing certain heterochronic genes [Zhang et al., 2011].

miRNA-Mediated Stem Cell Aging

Polycomb group (PcG) proteins like BMI1 control stem cell aging through regulation of aging-related genes such as the p16INKa/p19ARF locus, by triggering an increase in repressive histone marks (e.g. H3K27me3) [Bracken et al., 2007]. PcG proteins BMI1 and EZH2 are regulated by DNA-methyltransferases in a mechanism dependent on specific miRNA expression [So et al., 2011].

Senescence Entry and Maintenance via miRNAs

Senescence is an irreversible growth arrest program of cells, which can be induced by telomere attrition, oxidative stress, oncogene expression, or DNA damage signaling.

The establishment of senescence is associated with specific gene expression changes in p16, p53, and p21 [18711403], chromatin silencing of E2F target genes [12809602] and protein phosphorylation by DNA damage checkpoint kinases [14608368].

Senescence-associated miRNAs regulate genes related cell cycle, cytoskeletal remodelling and proliferation signaling. Different senescence-induction mechanisms regulate a common set of processes [19419692].

mir-34 family, particularly miR-34a, as downstream effectors of p53 involved in cell cycle [17656095], leads to cell cycle arrest, increased expression of Beta-galactosidase [17554337] and downregulation of E3F family target genes [17875987]. MDM2 inhibiting drug Nutlin-3, leads to p53 activation, induced up-regulation of primarily miR-34a and later miR-3b and miR-34c [18451145].

Overexpression of hsa-miR-20a in mouse embryonic fibroblasts induces senescence by lowering LRF 9a transcriptional repressor of the MDM2 inhibitor p19ARF [15662416; 9529248] protein levels and in turn increasing p19ARF levels [18596985].

hsa-miR-371, hsa-miR-369-5p, hsa-miR-29c, hsa-miR-499 and hsa-let-7f are significantly upregulated in senescent human mesenchymal stem cells (hMSCs) when compared to early passage hMSC [18493317]. has-miR-217 is also significantly upregulated but overall had very low expression levels.

miR-372 and miR-373 expression is able to bypass RAS-induced senescence in presence of wild-type p53 [16564011].

Upregulation of miR-143 is linked to senescence-dependent growth arrest in human fibroblasts [Bonifacio and Jarstfer, 2010].

Expression of senescence-related miRNAs, including let-7 family, miR-23a, miR-26a and miR-30a might be regulated by the activity of HDACs (Histone deacetylase complexes) [Lee et al., 2011].

miR-17, miR19b and miR-20a are downregulated during aging and/or senescence. Their decreased expression correlated with the increased transcript levels of their target genes, such as CDKN1A (p21) [Hackl et al., 2010].

miR-29 and miR-30 miRNA families are upregulated during induced and replicative senescence. Their upregulation requires the Rb pathway. B-Myb oncogene is repressed by miR-29 and miR-30. Interference with miR-29 and miR-30 expression inhibits senescence [Martinez et al., 2011].

miR-203 regulates p63 as well as Cav-1 and is involved in differentiation of skin stem cells [Lena et al. 2008; Yi et al. 2008]. Cav-1 is involved in cholesterol homeostasis, vesicular transport and signal transduction regulation [Williams et al. 2006]. miR-203 target sites in both Cav-1 and p63 3’UTRs exhibit high conservation across a number of species. Both p63 and Cav-1 are repressed by DR via miR-203.

A cohort of miRNA including miR-138, -181a, -181b and -130b increased in expression with serial passages, miR-138m -181a an -181b, but not miR-130b overexpression was sufficient to induce senescence. Sirt1 is a direct target of miR-138m -181a and -181b, whereas ΔNp63 expression was inhibited by miR-130b. ΔNp63 inhibits miR-138, -181a, -181b and -130b expression by binding directly to p63-responsive elements located in close proximity. Changes in miRNA expression by modulating the levels of regulatory proteins like p63 and Sirt1 strongly contributes to induction of senescence [Cervo et al., 2012].

Senescent cells secrete various kinds of inflammatory cytokines such as IL-6 and IL-8, as part of the senescence associated secretory phenotype which induces a bystander effect and leads to malignancy. miR-146a/b is significantly elevated during senescence as a compensatory response to restrain inflammation via the suppression of IL-6 and IL-8 secretion and downregulation of IRAK1 (component of IL-1 receptor signaling). IL-1α neutralizing antibodies abolish both miR-146a/b expression as well as IL-6 secretion [Bhaumik et al., 2009].

miR-17-92 cluster mediates ups an downs in ageing and cancer

miR-19 reduces PTEN levels and thus activates AKT/MTOR pathway. miR-19 and other members of the miR-17-92 cluster are commonly downregulated in several human replicative and organismal ageing models [Grillari et al., 2010].

Premature Aging is Due to Deregulation of miRNAs

LMNA mutation results in Hutchinson-Gilford Progeria Syndrome (HGPS). Zmpste24-deficient mice which age prematurely (a murine model of HGPS) have deregulated miR-29 expression [Ugalde et al., 2011]. miR-29 is also upregulated in somatic tissues from normal old mice. Its increased expression is strongly associated with DNA damage and p53-pathway [Ugalde et al., 2011]. These mice have also higher levels of miR-1 (targets IGF1) in liver, kidney as well as muscle.

Dietary Restriction Reverses miRNA Changes

Diet restriction (DR) extends the lifespan in all biomedical tested model organisms. Strikingly under DR there is instead of an age-dependent decrease in miRNA expression an age-dependent increase in brain miR-34a, miR30e and miR181-a-1, which target Bcl2, CREB1 and HTT [Khanna et al., 2011]. DR and exercise induce neurotropic factors (including BDNF), cytoprotective, chaperons and anti-apoptotic proteins. Neuroprotective effectors of DR and exercise are mediated, in part, by gene transcription changes [21236352]. The beneficial effects of DR on the aging brain may be due to down-regulation of miRNAs that target mRNAs encoding cell survival proteins [21415464]. Levels of the three miRNAs (miR-34a, miR30e and miR181-a-1) are lower in brain tissue samples form old mice (24 - 28 months) maintained on DR (40% beginning at 4 months). All three miRNAs have at least one target site for Bcl2, anti-apoptotic protein, which increase with DR. Overexpression of miR-30a, miR-34a and miR-181a result in decreased BCL2 levels and increased apoptosis via increases in pro-apoptosis genes such as Bac and cleavage of Caspases. miRNA expression change occurred in both cortical and hippocampal tissues, indicating a more global repression of these miRNAs in the central nervous system (CNS) due to DR. Additional shared targets were cAMP response element binding protein 1 (CREB1), important activator of several immediate response genes that are critical to synaptic plasticity [Deisseroth et al. 2003] and Huntington (HTT). miR-30a has an identical seed region to miR-30e acts to functionally repress BDNF expression in cortex [Mellios et al. 2008]. BDNF upregulation by DR mediates, in part, the increased neurogenesis by DR and is important in learning and memory [Lee et al. 2002; Mattson et al. 2004].

DR induces expression of miR-29c, miR-203, miR-150 and miR-30 in breast tissue of mice. miR-29 and miR-30 are involved in senescence [Martinez et al. 2011].

mir-34 Modulates Lifespan by Repressing Autophagy

miR-34a target genes are highly enriched for growth factor signaling and cell cycle progression which form a dense network that orchestrate the proliferative response to external growth stimuli.

mir-34 mutation extends lifespan by enhancing autophagic flux in C. elegans and mir-34 represses autophagy by inhibiting Atg9 in mammalian cells. mir-34 loss of function mutation in C. elegans delays age-related physiological decline, extends lifespan, and increases resistance to heat and oxidative stress. RNAi against autophagy-related genes, atg4, bec-1 or atg9 reversed lifespan extending effect of mir-34 mutants. miR-34a inhibits Atg9A expression post-transcriptional [Yang et al., 2011].

miR-34 regulates age-associated events and long-term brain integrity in Drosophila. miR-34 expression begins in the adulthood, is enriched in brain and increase with age. mir-34 loss triggers a gene expression profile of accelerated brain aging, late-onset brain degeneration and catastrophic decline in survival, while mir-34 upregulation extends median lifespan and mitigated neurodegeneration induced by polyglutamine. mir-34 silences Eip74EF (ETS domain transcription factor involved in steroid hormone pathways). miRNAs silence antagonistic pleiotropic developmental genes. Most miRNAs maintain a steady state level or decrease with gene, but not mir-34 which increases. It is barely detectable during development, but becomes high in adulthood and only its isoform c becomes upregulated with age. miR-34 mutants have no obvious developmental defects. miR-34 binding sites within the 3’UTR of the Eip74EF gene are conserved in orthologous Eip74EF genes from different Drosophila species. Eip74EF is a component of the steroid hormone signalling pathways which was studied primarily during development but was also implicated in lifespan regulation [Simon et al. 2003]. Flies lacking miR-34 had markedly increased E74A protein levels. E74A protein is highly expressed in young flies, but underwent a marked decrease within a 24h time window, which is mutually exclusive to the expression of miR-34. In flies lacking miR-34, the downregulation of E74A protein during the critical period was dampened. Deregulated expression of E74A has a negative impact on normal aging. One function of miR-34 is therefore to silence E74A in the adult to prevent the adult-stage deleterious activity of E74A on brain integrity and viability.

mir-34 expression is elevated with age in C. elegans too [de Lecanstre et al. 2010; Ibanez-Ventoso et al 2006]. Mammalian mir-34 orthologous are highly expressed in adult brain [Bak et al. 2008], increase with age and misregulated in degenerative diseases in humans [Zovoilis et al. 2011; Minones-Moyano et al 2011; Khanna et al. 2011; Gaughwin et al. 2011]. miR-34 function is neutral or adverse in C. elegans [de Lencastre et al. 2010, Yang et al. 2011] and can be either protective or contributory to age-associated events in vertebrates [Zovoilis et al. 2011; Minones-Moyano et al 2011; Khanna et al. 2011; Gaughwin et al. 2011].

Long-lived Mutants Retard Aging via miRNAs

Ames dwarf mice, which are long lived due to deficiency in pituitary hormones, have increased expression of ten miRNAs, including miR-27a. miR-27a post-transcriptionally suppresses key proteins of intermediate metabolism (glutathione metabolism, the urea cycle and polyamine biosynthesis), notable the biosynthetic pathway involving ornithine decarboxylase and spermidine synthase. The repressive action of miRNA-27a on ornithine decarboxylase in dwarf mouse liver was found as early as 2 months of age [Bates et al., 2010].

Various miRNAs are upregulated in the hippocampus of both long-lived Ames dwarf mice and growth hormone-receptor-knockout mice [Liang et al., 2011]. miR-470, miR-669b and miR-681 suppress expression of IGF1R and AKT (as well as its phosphorylation) which results in decreased FOXO3 phosphorylation (hence activation).

miRNAs as Biomarkers of Aging and Predictors of Lifespan

The expression patterns of just few miRNAs in C. elegans are predictive for the longevity of an individual and capable to predict up to 47% lifespan differences. mir-71 and mir-246 correlate with lifespan, while mir-239 anti-correlates and act upstream of insulin/IGF-1 like signalling (IIS) and other longevity pathways [Pincus et al., 2011].

In mice blood circulating miR-34a levels start to raise as early as 4 months, just before tissue deterioration begins and correlates with downregulation of its target gene SIRT1 in the blood as well as brain [Li et al., 2011c].

In mice and humans, miRNA levels measured in peripheral blood mononuclear cells correlate with aging [Noren Hooten et al., 2010] and even a single miRNA can function as marker for brain aging, neurodegeneration and memory impairment [Zovoilis et al., 2011].

16 miRNAs are up-regulate and 64 down-regulated in long-lived individuals [Elsharawy et al., 2012].

Boehm M, Slack F. A developmental timing microRNA and its target regulate life span in C. elegans. Science. 310(5756), 1954–1957 (2005).

Reprogramming

A specific group of miRNAs can reprogram cells very rapidly and efficiently. Suppression of Hdac2 is necessary part of this miRNA-mediated reprogramming process [Anokye-Danso et al., 2011].

.. http://www.cell.com/cell-stem-cell/abstract/S1934-5909(11)00111-1 Rethinking Reprogramming: A New Way To Make Stem Cellshttp://www.youtube.com/watch?v=OgyXpU2OTGg

Can expression of miRNAs highly overexpressed in young cells revert old cells back to a young state?

Conclusions

miRNAs might control the senescence of particular ‘rate-limiting’ cells or tissues, or maybe important in the coordination of the aging rates of different tissues. miRNAs are more responsive than transcription factors to stressful conditions. miRNAs appear to be important in biological decisions from the earliest to the latest stages of C. elegans’ life as well as those of other animals including our own species.

References

Anokye-Danso, F., Trivedi, C.M., Juhr, D., Gupta, M., Cui, Z., Tian, Y., Zhang, Y., Yang, W., Gruber, P.J., Epstein, J.A., et al. (2011). Highly Efficient miRNA-Mediated Reprogramming of Mouse and Human Somatic Cells to Pluripotency. Cell Stem Cell 8, 376-388.

Bates, D.J., Li, N., Liang, R., Sarojini, H., An, J., Masternak, M.M., Bartke, A., and Wang, E. (2010). MicroRNA regulation in Ames dwarf mouse liver may contribute to delayed aging. Aging cell 9, 1-18.

Berdasco, M., and Esteller, M. (2012). Hot topics in epigenetic mechanisms of aging: 2011. Aging cell.

Boehm, M., and Slack, F. (2005). A developmental timing microRNA and its target regulate life span in C. elegans. Science 310, 1954-1957.

Cervo, P.R., Lena, A.M., Nicoloso, M., Rossi, S., Mancini, M., Zhou, H., Saintigny, G., Dellambra, E., Odorisio, T., Mahe, C., et al. (2012). p63-microRNA feedback in keratinocyte senescence. Proceedings of the National Academy of Sciences of the United States of America.

de Lencastre, A., Pincus, Z., Zhou, K., Kato, M., Lee, S.S., and Slack, F.J. (2010). MicroRNAs both promote and antagonize longevity in C. elegans. Current biology : CB 20, 2159-2168.

Drummond, M.J., McCarthy, J.J., Fry, C.S., Esser, K.A., and Rasmussen, B.B. (2008). Aging differentially affects human skeletal muscle microRNA expression at rest and after an anabolic stimulus of resistance exercise and essential amino acids. American journal of physiology Endocrinology and metabolism 295, E1333-1340.

Drummond, M.J., McCarthy, J.J., Sinha, M., Spratt, H.M., Volpi, E., Esser, K.A., and Rasmussen, B.B. (2011). Aging and microRNA expression in human skeletal muscle: a microarray and bioinformatics analysis. Physiological genomics 43, 595-603.

Elsharawy, A., Keller, A., Flachsbart, F., Wendschlag, A., Jacobs, G., Kefer, N., Brefort, T., Leidinger, P., Backes, C., Meese, E., et al. (2012). Genome-wide miRNA signatures of human longevity. Aging cell.

Hackl, M., Brunner, S., Fortschegger, K., Schreiner, C., Micutkova, L., Muck, C., Laschober, G.T., Lepperdinger, G., Sampson, N., Berger, P., et al. (2010). miR-17, miR-19b, miR-20a, and miR-106a are down-regulated in human aging. Aging cell 9, 291-296.

Hamrick, M.W., Herberg, S., Arounleut, P., He, H.Z., Shiver, A., Qi, R.Q., Zhou, L., Isales, C.M., and Mi, Q.S. (2010). The adipokine leptin increases skeletal muscle mass and significantly alters skeletal muscle miRNA expression profile in aged mice. Biochemical and biophysical research communications 400, 379-383.

He, X., Yan, Y.L., DeLaurier, A., and Postlethwait, J.H. (2011). Observation of miRNA gene expression in zebrafish embryos by in situ hybridization to microRNA primary transcripts. Zebrafish 8, 1-8.

Hutchison, E., and Mattson, M.P. (2011). Eating less suppresses microRNA assassins in the brain. Aging 3, 179-180.

Ibanez-Ventoso, C., and Driscoll, M. (2009). MicroRNAs in C. elegans Aging: Molecular Insurance for Robustness? Current genomics 10, 144-153.

Ibanez-Ventoso, C., Yang, M., Guo, S., Robins, H., Padgett, R.W., and Driscoll, M. (2006). Modulated microRNA expression during adult lifespan in Caenorhabditis elegans. Aging cell 5, 235-246.

Karp, X., Hammell, M., Ow, M.C., and Ambros, V. (2011). Effect of life history on microRNA expression during C. elegans development. RNA 17, 639-651.

Kato, M., Chen, X., Inukai, S., Zhao, H., and Slack, F.J. (2011). Age-associated changes in expression of small, noncoding RNAs, including microRNAs, in C. elegans. RNA 17, 1804-1820.

Khanna, A., Muthusamy, S., Liang, R., Sarojini, H., and Wang, E. (2011). Gain of survival signaling by down-regulation of three key miRNAs in brain of calorie-restricted mice. Aging 3, 223-236.

Li, N., Bates, D.J., An, J., Terry, D.A., and Wang, E. (2011a). Up-regulation of key microRNAs, and inverse down-regulation of their predicted oxidative phosphorylation target genes, during aging in mouse brain. Neurobiology of aging 32, 944-955.

Li, N., Muthusamy, S., Liang, R., Sarojini, H., and Wang, E. (2011b). Increased expression of miR-34a and miR-93 in rat liver during aging, and their impact on the expression of Mgst1 and Sirt1. Mechanisms of ageing and development.

Li, X., Khanna, A., Li, N., and Wang, E. (2011c). Circulatory miR34a as an RNAbased, noninvasive biomarker for brain aging. Aging 3, 985-1002.

Liang, R., Khanna, A., Muthusamy, S., Li, N., Sarojini, H., Kopchick, J.J., Masternak, M.M., Bartke, A., and Wang, E. (2011). Post-transcriptional regulation of IGF1R by key microRNAs in long-lived mutant mice. Aging cell 10, 1080-1088.

Lim, D.H., Oh, C.T., Lee, L., Hong, J.S., Noh, S.H., Hwang, S., Kim, S., Han, S.J., and Lee, Y.S. (2011). The endogenous siRNA pathway in Drosophila impacts stress resistance and lifespan by regulating metabolic homeostasis. FEBS letters 585, 3079-3085.

Liu, N., Landreh, M., Cao, K., Abe, M., Hendriks, G.J., Kennerdell, J.R., Zhu, Y., Wang, L.S., and Bonini, N.M. (2012). The microRNA miR-34 modulates ageing and neurodegeneration in Drosophila. Nature 482, 519-523.

Maes, O.C., An, J., Sarojini, H., and Wang, E. (2008). Murine microRNAs implicated in liver functions and aging process. Mechanisms of ageing and development 129, 534-541.

Martinez, I., Cazalla, D., Almstead, L.L., Steitz, J.A., and DiMaio, D. (2011). miR-29 and miR-30 regulate B-Myb expression during cellular senescence. Proceedings of the National Academy of Sciences of the United States of America 108, 522-527.

Noren Hooten, N., Abdelmohsen, K., Gorospe, M., Ejiogu, N., Zonderman, A.B., and Evans, M.K. (2010). microRNA expression patterns reveal differential expression of target genes with age. PloS one 5, e10724.

Orom, U., Lim, M.K., Savage, J.E., Jin, L., Saleh, A.D., Lisanti, M.P., and Simone, N.L. (2012). microRNA-203 regulates caveolin-1 in breast tissue during caloric restriction. Cell Cycle 11.

Persengiev, S., Kondova, I., Otting, N., Koeppen, A.H., and Bontrop, R.E. (2011). Genome-wide analysis of miRNA expression reveals a potential role for miR-144 in brain aging and spinocerebellar ataxia pathogenesis. Neurobiology of aging 32, 2316 e2317-2327.

Pincus, Z., Smith-Vikos, T., and Slack, F.J. (2011). MicroRNA predictors of longevity in Caenorhabditis elegans. PLoS genetics 7, e1002306.

Qureshi, I.A., and Mehler, M.F. (2011). Non-coding RNA networks underlying cognitive disorders across the lifespan. Trends in molecular medicine.

Sredni, S.T., Gadd, S., Jafari, N., and Huang, C.C. (2011). A Parallel Study of mRNA and microRNA Profiling of Peripheral Blood in Young Adult Women. Frontiers in genetics 2, 49.

Ugalde, A.P., Ramsay, A.J., de la Rosa, J., Varela, I., Marino, G., Cadinanos, J., Lu, J., Freije, J.M., and Lopez-Otin, C. (2011). Aging and chronic DNA damage response activate a regulatory pathway involving miR-29 and p53. The EMBO journal 30, 2219-2232.

Xu, P., Vernooy, S.Y., Guo, M., and Hay, B.A. (2003). The Drosophila microRNA Mir-14 suppresses cell death and is required for normal fat metabolism. Current biology : CB 13, 790-795.

Yang, J., Chen, D., He, Y., Melendez, A., Feng, Z., Hong, Q., Bai, X., Li, Q., Cai, G., Wang, J., et al. (2011). MiR-34 modulates Caenorhabditis elegans lifespan via repressing the autophagy gene atg9. Age (Dordr).

Zhang, X., Zabinsky, R., Teng, Y., Cui, M., and Han, M. (2011). microRNAs play critical roles in the survival and recovery of Caenorhabditis elegans from starvation-induced L1 diapause. Proceedings of the National Academy of Sciences of the United States of America 108, 17997-18002.

Zovoilis, A., Agbemenyah, H.Y., Agis-Balboa, R.C., Stilling, R.M., Edbauer, D., Rao, P., Farinelli, L., Delalle, I., Schmitt, A., Falkai, P., et al. (2011). microRNA-34c is a novel target to treat dementias. The EMBO journal 30, 4299-4308.

Bonifacio LN, Jarstfer MB (2010) MiRNA profile associated with replicative senescence, extended cell culture, and ectopic telomerase expression in human foreskin fibroblasts. PLoS One. 5(9), e12519.

Lena AM, Shalom-Feuerstein R, Rivetti di Val Cervo P, Aberdam D, Knight RA, Melino G, et al. miR-203 represses ‘stemness’ by repressing DeltaNp63. Cell Death Differ 2008; 15:1187-95; PMID:18483491; http://dx.doi.org/10.1038/cdd.2008.69.

Yi R, Poy MN, Stoffel M, Fuchs E. A skin microRNA promotes differentiation by repressing ‘stemness’. Nature 2008; 452:225-9; PMID:18311128; http://dx.doi.org/10.1038/nature06642.

Williams TM, Sotgia F, Lee H, Hassan G, Di Vizio D, Bonuccelli G, et al. Stromal and epithelial caveo- lin-1 both confer a protective effect against mammary hyperplasia and tumorigenesis: Caveolin-1 antagonizes cyclin D1 function in mammary epithelial cells. Am J Pathol 2006; 169:1784-801; PMID:17071600; http://dx.doi.org/10.2353/ajpath.2006.060590.

Deisseroth K, Mermelstein PG, Xia H, Tsien RW. Signaling from synapse to nucleus: the logic behind the mechanisms. Curr Opin Neurobiol. 2003; 13: 354-365.

Mellios N, Huang HS, Grigorenko A, Rogaev E, Akbarian S. A set of differentially expressed miRNAs, including miR-30a-5p, act as post-transcriptional inhibitors of BDNF in prefrontal cortex. Hum Mol Genet. 2008; 17: 3030-3042.

Lee J, Duan W, Mattson MP. Evidence that brain-derived neurotrophic factor is required for basal neurogenesis and mediates, in part, the enhancement of neurogenesis by dietary restriction in the hippocampus of adult mice. J Neurochem. 2002; 82: 1367-1375.

Mattson MP, Duan W, Wan R, Guo Z. Prophylactic activation of neuroprotective stress response pathways by dietary and behavioral manipulations. NeuroRx. 2004; 1: 111-116.

Martinez I, Cazalla D, Almstead LL, Steitz JA, DiMaio D. miR-29 and miR-30 regulate B-Myb expression during cellular senescence. Proc Natl Acad Sci USA 2011; 108:522-7; PMID:21187425; http://dx.doi.org/10.1073/pnas.1017346108.

Simon, A. F., Shih, C., Mack, A. & Benzer, S. Steroid control of longevity in Drosophila melanogaster. Science 299, 1407–1410 (2003).

de Lencastre, A. et al. MicroRNAs both promote and antagonize longevity in C. elegans. Curr. Biol. 20, 2159–2168 (2010).

Ibanez-Ventoso, C. et al. Modulated microRNA expression during adult lifespan in Caenorhabditis elegans. Aging Cell 5, 235–246 (2006).

Bak, M. et al. MicroRNA expression in the adult mouse central nervous system. RNA 14, 432–444 (2008).

Zovoilis, A. et al. microRNA-34c is a novel target to treat dementias. EMBO J. 30, 4299–4308 (2011).

Minones-Moyano, E. et al. MicroRNA profiling of Parkinson’s disease brains identifies early downregulation of miR-34b/c which modulate mitochondrial function. Hum. Mol. Genet. 20, 3067–3078 (2011).

Li, X., Khanna, A., Li, N. & Wang, E. Circulatory miR34a as an RNA based, noninvasive biomarker for brain aging. Aging 3, 985–1002 (2011).

Khanna, A., Muthusamy, S., Liang, R., Sarojini, H. & Wang, E. Gain of survival signaling by down-regulation of three key miRNAs in brain of calorie-restricted mice. Aging 3, 223–236 (2011).

Gaughwin, P. M. et al. Hsa-miR-34b is a plasma-stable microRNA that is elevated in pre-manifest Huntington’s disease. Hum. Mol. Genet. 20, 2225–2237 (2011).

de Lencastre, A. et al. MicroRNAs both promote and antagonize longevity in C. elegans. Curr. Biol. 20, 2159–2168 (2010).

Yang, J. et al. MiR-34 modulates Caenorhabditis elegans lifespan via repressing the autophagy gene atg9. Age. doi:10.1007/s11357-011-9324-3 (2011).

Kato M, Chen X, Inukai S, Zhao H, Slack FJ. Age-associated changes in expression of small, noncoding RNAs, including microRNAs, in C. elegans. RNA. 17(10), 1804–1820 (2011).

de Lencastre A, Pincus Z, Zhou K, Kato M, Lee SS, Slack FJ. MicroRNAs both promote and antagonize longevity in C. elegans. Curr. Biol. 20(24), 2159–2168 (2010).

Boulias K, Horvitz HR. The C. elegans microRNA mir-71 acts in neurons to promote germline-mediated longevity through regulation of DAF-16/FOXO. Cell Metab. 15(4), 439–450 (2012).

Elsharawy A, Keller A, Flachsbart F, et al. Genome-wide miRNA signatures of human longevity. Aging Cell. (2012).

Noren Hooten N, Abdelmohsen K, Gorospe M, Ejiogu N, Zonderman AB, Evans MK. microRNA expression patterns reveal differential expression of target genes with age. PLoS ONE. 5(5),

Inukai S, de Lencastre A, Turner M, Slack F. Novel MicroRNAs Differentially Expressed during Aging in the Mouse Brain. PLoS ONE. 7(7), e40028 (2012).

Lanceta J, Prough RA, Liang R, Wang E. MicroRNA group disorganization in aging. Exp Gerontol. 45(4), 269–278 (2010).

Inukai S, de Lencastre A, Turner M, Slack F. Novel MicroRNAs Differentially Expressed during Aging in the Mouse Brain. PLoS ONE. 7(7), e40028 (2012).

Persengiev S, Kondova I, Otting N, Koeppen AH, Bontrop RE. Genome-wide analysis of miRNA expression reveals a potential role for miR-144 in brain aging and spinocerebellar ataxia pathogenesis. Neurobiol Aging. 32(12), 2316.e17–27 (2011).

Li X, Khanna A, Li N, Wang E. Circulatory miR34a as an RNAbased, noninvasive biomarker for brain aging. Aging (Albany NY). 3(10), 985–1002 (2011).


Categories: Article, reST

Comment: Corrected unicode encoding errors.

See entry | Admin

Comment on This Data Unit