ArticlePDF Available

Small RNAs – their biogenesis, regulation and function in embryonic stem cells

Authors:

Abstract and Figures

Small non-coding RNAs, including microRNAs (miRNAs), endogenous small interfering RNAs (endo-siRNAs) and Piwi-interacting RNAs (piRNAs), play essential roles in mammalian development. The function and timing of expression of these three classes of small RNAs differ greatly. piRNAs are expressed and play a crucial role during male gametogenesis, whereas endo-siRNAs are essential for oocyte meiosis. By contrast, miRNAs are ubiquitously expressed in somatic tissues and function throughout post-implantation development. Surprisingly, however, miRNAs are non-essential during pre-implantation embryonic development and their function is suppressed during oocyte meiosis. Here, we review the roles of small non-coding RNAs during the early stages of mammalian development, from gamete maturation through to gastrulation.
Content may be subject to copyright.
1653
Summary
Small non-coding RNAs, including microRNAs (miRNAs),
endogenous small interfering RNAs (endo-siRNAs) and Piwi-
interacting RNAs (piRNAs), play essential roles in mammalian
development. The function and timing of expression of these
three classes of small RNAs differ greatly. piRNAs are expressed
and play a crucial role during male gametogenesis, whereas
endo-siRNAs are essential for oocyte meiosis. By contrast,
miRNAs are ubiquitously expressed in somatic tissues and
function throughout post-implantation development.
Surprisingly, however, miRNAs are non-essential during pre-
implantation embryonic development and their function is
suppressed during oocyte meiosis. Here, we review the roles of
small non-coding RNAs during the early stages of mammalian
development, from gamete maturation through to
gastrulation.
Key words: siRNA, MicroRNA, Stem cells, Oocyte, Embryonic
development
Introduction
The processes of growth and differentiation are kept in balance
during the development of multicellular organisms. Post-
transcriptional control of gene expression plays a key role in this
balance by coordinating the expression of selected genes at specific
times and places. The role of post-transcriptional regulation is
particularly apparent in early mammalian development, from
maturation of the germ line to initiation of gastrulation, when
controls on mRNA localization, stability and translation are the
fundamental means of gene regulation. Indeed, from the fully
grown oocyte stage until zygotic genome activation (ZGA), the
genome is transcriptionally silent (Abe et al., 2010). Therefore, all
mRNA regulation must occur post-transcriptionally. Following
ZGA, the embryo establishes populations of early stem cells (SCs)
within the inner cell mass (ICM, the collection of cells that
eventually will form the fetus) that begin to proliferate rapidly,
ensuring that stocks of unspecialized cells are established for future
differentiation into the three germ layers. This rapid growth and the
subsequent switch from unspecialized cells into specific cell types
is a highly regulated process involving much post-transcriptional
regulation (Lu et al., 2009). Similar regulation also occurs in the
extra-embryonic tissues. This review will focus on one group of
post-transcriptional regulators, the small non-coding RNAs. These
RNAs range in size from 18 to 32 nucleotides (nt) in length and
have emerged in the past decades as major players in post-
transcriptional regulation across many, if not most, multicellular
organisms.
Classes and biogenesis of mammalian small RNAs
Three major classes of functional small non-coding RNAs have
been found in mammals: microRNAs (miRNAs), endogenous
small interfering RNAs (endo-siRNAs) and Piwi-interacting RNAs
(piRNAs) (Babiarz and Blelloch, 2009; Kim et al., 2009; Thomson
and Lin, 2009). These classes differ in their biogenesis, i.e. their
maturation from transcribed forms to the active form of the RNA
(Fig. 1).
miRNAs can be divided into two subclasses: canonical and non-
canonical miRNAs. Canonical miRNAs are initially transcribed as
long RNAs that contain hairpins (Fig. 1A). The 60-75 nt hairpins
are recognized by the RNA-binding protein Dgcr8 (DiGeorge
syndrome critical region 8), which directs the RNase III enzyme
Drosha to cleave the base of the hairpin (Denli et al., 2004;
Gregory et al., 2004; Han et al., 2004; Han et al., 2006; Landthaler
et al., 2004; Lee et al., 2003). Following cleavage by the Drosha-
Dgcr8 complex, also called the microprocessor, the released
hairpin is transported to the cytoplasm, where Dicer, another
RNase III enzyme, then cleaves it into a single short 18-25 nt
dsRNA (Bernstein et al., 2001; Hutvagner et al., 2001; Ketting et
al., 2001; Knight and Bass, 2001). Non-canonical miRNAs bypass
processing by the microprocessor by using other endonucleases or
by direct transcription of a short hairpin. The resulting pre-
miRNAs are then exported from the nucleus and cleaved once by
Dicer (Babiarz et al., 2008; Okamura et al., 2007; Ruby et al.,
2007).
By contrast, siRNAs are derived from long dsRNAs (Fig. 1B) in
the form of either sense or antisense RNA pairs or as long hairpins,
which are then directly processed by Dicer consecutively along the
dsRNA to produce multiple siRNAs (Chung et al., 2008; Czech et
al., 2008; Ghildiyal et al., 2008; Kawamura et al., 2008; Okamura
et al., 2008a; Okamura et al., 2008b). Therefore, canonical
miRNAs, non-canonical miRNAs and endo-siRNAs all involve
Dicer processing and are ~21 nt in length. Furthermore, in all three
cases, one strand of the Dicer product associates with an Argonaute
protein (Ago 1-4 in mammals; also known as Eif2c1-4) to form the
active RISC (RNA-induced silencing complex, Fig. 1D)
(Filipowicz, 2005). These ribonucleoprotein complexes are able to
bind to and control the levels and translation of their target mRNAs:
if the match between the small RNA and its target is perfect, the
target is cleaved; if not, the mRNA is destabilized through as yet
unresolved mechanisms (Doench et al., 2003; Fabian et al., 2010;
Zeng et al., 2003).
The piRNAs differ from the miRNAs and endo-siRNAs in that
they do not require Dicer for their processing (Houwing et al.,
2007; Vagin et al., 2006). Indeed, how piRNAs are produced and
their mechanism of action remains poorly characterized (for a
review, see Klattenhoff and Theurkauf, 2008). piRNAs are 25-32
nt in length, and are expressed predominantly in the germline in
mammals (Aravin et al., 2006; Grivna et al., 2006; Watanabe et al.,
2006). They are defined by their interaction with the Piwi proteins,
a distinct family of Argonaute proteins (including Miwi, Miwi2
Development 138, 1653-1661 (2011) doi:10.1242/dev.056234
© 2011. Published by The Company of Biologists Ltd
Small RNAs in early mammalian development: from gametes
to gastrulation
Nayoung Suh and Robert Blelloch*
The Eli and Edythe Broad Center of Regeneration Medicine and Stem Cell Research,
Center for Reproductive Sciences, and Department of Urology, University of
California San Francisco, San Francisco, CA 94143, USA.
*Author for correspondence (Blellochr@stemcell.ucsf.edu)
REVIEW
DEVELOPMENT
1654
and Mili in mouse; also known as Piwil1, Piwil4 and Piwil2,
respectively). piRNAs are generated from long single-stranded
RNA precursors that are often encoded by complex and repetitive
intergenic sequences. One proposed model for their biogenesis is
the ‘ping-pong mechanism’ (Fig. 1C) (Aravin et al., 2007;
Brennecke et al., 2007; Gunawardane et al., 2007). In this model,
the Argonaute protein Mili cleaves the primary piRNA to define
the 5end of piRNAs, which is subsequently recognized by Miwi2.
Miwi2 then cleaves the other strand of the precursor, thereby
generating a 5end of the piRNA that can then bind to Mili, thus
forming a positive amplification loop. Many of the details of this
model remain to be uncovered. Furthermore, the ping-pong model
is likely to explain the biogenesis of only a subset of mammalian
piRNAs, those that are derived from repetitive sequences, such as
transposons, and that are associated with Miwi2 and Mili in the
early stages of spermatogenesis. The mechanism of biogenesis of
piRNAs derived from complex intergenic sequences, associated
with Miwi and Mili in the later stages of spermatogenesis, is
unclear.
Small RNAs in gametes
Most animals, including vertebrates, reproduce sexually and have
the ability to form gametes. The two types of gametes, the egg and
the sperm, arise from immature germ cells, undergo extensive
differentiation and ultimately fuse to create their progeny. One
fascinating aspect of this process is that, upon fertilization, the
highly specialized sperm and egg unite to produce the zygote,
which is totipotent, having the potential to produce all the cells of
the body. This switch from a singular function to a totipotent cell
involves massive molecular rewiring (Hemberger et al., 2009).
Based on recent findings, small RNAs are likely to play a very
important role during this transition.
Small RNAs in spermatogenesis
Spermatogenesis is a highly regulated process, during which
diploid spermatogonia differentiate into haploid spermatozoa
within the seminiferous epithelium of the testis. During the course
of differentiation into sperm, numerous mRNAs are regulated
post-transcriptionally (Lee et al., 2009). Recent studies using both
REVIEW Development 138 (9)
pri-miRNA
pre-miRNA
Micro-
processor
Transcription
Mirtron
Splicing
Endo-siRNA locus
Inverted
repeat
Transcription
pre-miRNA
Export
shRNA
pathway
A miRNA pathway
D Effector pathway
B endo-siRNA pathway
Canonical
pathway
Non-canonical
pathway
Drosha
Dgcr8
Dicer
Nucleus
Cytoplasm
Dicer
Ago2
C piRNA pathway
Nucleus
Cytoplasm
piRNA origin
Primary
processing
Miwi2
Mili
Secondary
processing
Endonucleolytic cleavage
Ago1-4
Translational repression or
de-adenylation
miRNA siRNA
Strand selection/
RNP assembly
RISC RISC
cis
endo-siRNA
trans
endo-siRNA
Fig. 1. Biogenesis of microRNAs (miRNAs), endogenous small interfering RNAs (endo-siRNAs) and Piwi-interacting RNAs (piRNAs).
(A)
Canonical miRNAs are processed from long primary miRNAs (pri-miRNAs) into short hairpin precursor miRNAs (pre-miRNAs) by the
microprocessor, a complex consisting of the RNA binding protein Dgcr8 and the RNase III enzyme Drosha. By contrast, non-canonical miRNAs
are transcribed directly as short hairpins (shRNAs) or derive from introns that can refold into shRNAs (mirtrons). (B)
Precursors of endo-siRNAs
are derived from long stem-loop structures (inverted repeat), opposing strand transcription (cis endo-siRNAs), or gene-pseudogene pairs (trans
endo-siRNAs). Both miRNAs and endo-siRNAs are then processed by the RNase III enzyme Dicer to produce double-stranded RNAs of ~21
nucleotides. (C)
piRNAs are processed from single-stranded RNA precursors that are often encoded by intergenic repetitive elements or
transposons. The mechanisms that drive piRNA biogenesis are not well understood, although a ‘ping-pong mechanism’ has been described
for a subset of piRNAs. In this model, Mili cleaves the primary piRNA, which is subsequently recognized by Miwi2. Miwi2 cleaves the other
strand of the precursor that can then bind to Mili, thus forming a positive amplification loop. (D)
Following their processing, miRNAs and
endo-siRNAs are assembled into ribonucleoprotein (RNP) complexes called RNA-induced silencing complexes (RISCs). The key components of
RISCs are proteins of the Argonaute (Ago) family. In mammals, four Ago proteins (Ago 1-4) function in miRNA repression but only Ago2
functions in siRNA repression. The fate of piRNAs is unknown. On the DNA, blue represents the positive strand and red represents the
negative strand.
DEVELOPMENT
1655REVIEWDevelopment 138 (9)
genetics and miRNA profiling on different populations of
spermatogenetic cells have identified an important role for
miRNAs during spermatogenesis (Hayashi et al., 2008; Tang et al.,
2007; Yu et al., 2005; Bouhallier et al., 2010; Yan et al., 2009).
Deletion of Dicer, for example, results in a loss of sperm (Hayashi
et al., 2008; Maatouk et al., 2008). This could be attributed to the
loss of either miRNAs or endo-siRNAs. However, the deletion of
Argonaute 2 (Ago2), a protein that is essential for cleavage of
mRNA targets by endo-siRNAs, has no obvious testis phenotype,
suggesting that the Dicer phenotype is predominantly an miRNA-
based phenotype (Hayashi et al., 2008). Roles for individual
miRNAs during spermatogenesis have also been described. For
example, Mir122a regulates Tnp2, a testis-specific gene involved
in chromatin remodeling during spermatogenesis (Yu et al., 2005).
In addition, Mir34c is highly expressed in germ cells and its
overexpression enhances spermatogenesis (Bouhallier et al.,
2010). Furthermore, Dead end 1 (Dnd1), an RNA-binding protein
that is implicated in prevention of miRNA access to cell cycle-
related target mRNAs, is essential for fetal male germ cell
development (Cook et al., 2010; Kedde et al., 2007). Recently, it
has been also shown that Mir18, a member of the Oncomir-1
cluster of miRNAs, directly targets heat shock factor 2 (Hsf2), a
transcription factor involved in spermatogenesis (Bjork et al.,
2010).
Genetic studies imply that, like miRNAs, piRNAs are also
essential for spermatogenesis. In mouse, piRNAs have been
separated into two classes based on the timing of their expression,
their repetitive versus nonrepetitive nature, and the Piwi proteins
with which they are associated (Aravin et al., 2006; Girard et al.,
2006; Grivna et al., 2006; Lau et al., 2006; Watanabe et al., 2006).
The first class is highly repetitive and is expressed before meiotic
pachytene. This class of piRNAs interacts with Mili and Miwi2
(Aravin et al., 2008; Aravin et al., 2007). The second class of
piRNAs is nonrepetitive, becomes abundant during the pachytene
stage and is associated with Mili and Miwi proteins (Aravin et al.,
2008; Aravin et al., 2007; Girard et al., 2006). Consistent with their
timing of expression, deletion of Mili and Miwi2 results in early
arrest in meiosis I (at the primary spermatocyte stage), whereas
deletion of Miwi results in arrest following meiosis II (the round
spermatid stage) (Carmell et al., 2007; Deng and Lin, 2002;
Kuramochi-Miyagawa et al., 2004).
The repetitive piRNAs are associated with the repression of
transposable elements during spermatogenesis (Malone and
Hannon, 2009) but exactly how repression is achieved is unclear.
For example, it is unclear whether this repression occurs
transcriptionally or post-transcriptionally. There is evidence to
suggest that repetitive piRNAs promote de novo demethylation of
the transposons (Aravin et al., 2008; Kuramochi-Miyagawa et al.,
2008); however the mechanism by which these small RNAs can
direct the DNA methylation machinery is unclear.
Small RNAs in oogenesis
Mammalian oogenesis is distinct from spermatogenesis (Matova
and Cooley, 2001). The oocyte pool is largely fixed by the end of
mouse embryogenesis and oocytes are then induced to mature in
response to cyclic waves of hormones, including follicle
stimulating hormone and luteinizing hormone. In the mouse, each
cycle induces a small pool of oocytes to mature before they are
released into the fallopian tube. During maturation, the oocytes
undergo germinal vesicle breakdown (GVBD), meiosis I and
finally meiosis II, which is only completed following fertilization.
Deep sequencing of small RNAs in mouse oocytes has uncovered
not only miRNAs and piRNAs, but also a large population of endo-
siRNAs (Tam et al., 2008; Watanabe et al., 2006; Watanabe et al.,
2008).
An essential role for endo-siRNAs in the mouse oocyte has been
inferred by comparing the knockout phenotypes of Dicer and Dgcr8
mutant mice. Dicer loss in the oocyte results in meiotic arrest with
severe spindle and chromosomal segregation defects (Murchison et
al., 2007; Suh et al., 2010; Tang et al., 2007). Furthermore, thousands
of mRNAs are misregulated. By contrast, Dgcr8 loss has no
phenotype and mRNA levels remain unchanged (Suh et al., 2010).
As Dicer processes both miRNAs and endo-siRNAs, whereas Dgcr8
is essential only for miRNA processing, these findings imply that
endo-siRNAs, and not miRNAs, underlie the meiotic defect of Dicer
knockout oocytes. The loss of Ago2 results in a similar phenotype to
that observed in Dicer knockouts, further supporting the role of
endo-siRNAs in regulating meiosis in oocytes (Kaneda et al., 2009).
The mechanism of action of endo-siRNAs in oocytes is unclear. In
Schizosaccharomyces pombe, endo-siRNAs are crucial for
heterochromatin formation in repeat regions of the genome (Moazed,
2009). However, no such function has been shown convincingly in
mammals. It seems more likely that endo-siRNAs in mammals are
acting post-transcriptionally (Tam et al., 2008).
It was surprising to find that in the absence of Dgcr8 in mouse
oocytes, mRNA levels are unchanged. miRNAs are present in
oocytes, as determined by both deep sequencing and multiplex
quantitative PCR-based profiling (Murchison et al., 2007; Tam et
al., 2008; Tang et al., 2007; Watanabe et al., 2006; Watanabe et al.,
2008). For example, Let-7, Mir22, Mir16-1 and Mir29 are all highly
expressed in the oocyte. miRNA profiling following Dgcr8 deletion
confirmed that these and all other miRNAs tested were indeed lost
in knockout oocytes (Suh et al., 2010). Furthermore, mRNA
profiling and bioinformatic analyses showed that many targets for
the expressed miRNAs are present in the oocyte. Therefore,
everything is in place for miRNA-based destabilization to occur, but
mRNA levels remain unchanged (Suh et al., 2010). Consistent with
these findings, reporter assays show robust siRNA activity in mature
oocytes, but little to no miRNA function (Ma et al., 2010). Even
with artificial 3untranslated regions (3UTRs) carrying multiple
target sites, and the introduction of supraphysiological doses of
miRNAs, little suppression in terms of mRNA stability or
translation was seen (Ma et al., 2010). Together, these surprising
results show that miRNA function is suppressed in fully grown
oocytes even though miRNA biogenesis is unaffected and miRNA
targets are present. Although the mechanism of suppression is
unknown, one hint comes from the finding that P-bodies (processing
bodies, see Box 1), in which miRNA destabilization normally
occurs (Parker and Sheth, 2007), are lost in maturing oocytes and
only reform at the blastocyst stage (Flemr et al., 2010; Swetloff et
al., 2009). Whether the loss of P-bodies is a primary or secondary
consequence of miRNA functional loss is unclear.
piRNAs are also expressed in mouse oocytes (Watanabe et al.,
2008), but the deletion of the Piwi proteins do not produce an
oocyte phenotype (Carmell et al., 2007; Deng and Lin, 2002;
Kuramochi-Miyagawa et al., 2004). Therefore, it is unclear
whether they play any role during oogenesis.
Small RNAs in early embryogenesis
Small RNA function during pre-implantation development
Zygotic deletion of Dgcr8 or Dicer in mice leads to embryonic
arrest shortly after implantation between embryonic day (E) 6.5 and
E7.5 (Bernstein et al., 2003; Morita et al., 2007; Wang et al., 2007).
However, development to E3.5, the blastocyst stage, occurs
DEVELOPMENT
1656
normally. Indeed, maternal and zygotic loss of Dgcr8 leads to no
discernable phenotype in E3.5 embryos (Suh et al., 2010).
Therefore, miRNA function must become essential sometime
between E3.5 and E7.5. Recently, careful characterization of the
zygotic Dicer knockout phenotype suggested that an epiblast forms
and there is an initiation of gastrulation with expression of the early
mesoderm maker brachyury in the posterior epiblast (Spruce et al.,
2010). However, the primitive streak fails to elongate and there is
a loss of expression of the definitive endoderm markers Hex
(Hhex – Mouse Genome Informatics) and Cerl1 (Cer1). A major
caveat of these findings is that it is unclear whether all miRNAs
were lost in the Dicer knockout embryos. Indeed, in situ
hybridization suggested equal levels of miRNAs from the Mir290
cluster in Dicer versus wild-type embryos. Expression of the
Mir290 cluster is initiated with zygotic gene activation (Tang et al.,
2007). Hence, their presence in the zygotic Dicer knockout
blastocyst suggests the perdurance of maternal Dicer rather than
the miRNAs themselves. Because of this caveat, the earliest roles
of miRNAs in mouse development remain unknown. However, the
strong proliferation defects seen in Dicer and Dgcr8 knockout
mouse embryonic stem (ES) cells (Kanellopoulou et al., 2005;
Murchison et al., 2005; Wang et al., 2008; Wang et al., 2007),
suggests that miRNAs are likely to play an important role in the
expansion of the epiblast. Interestingly, the loss of maternal
miRNAs alone results in a decrease in the average number of
progeny produced following fertilization by wild-type males (Suh
et al., 2010). Together with a lack of pre-implantation phenotype,
this finding suggests a role for maternally contributed miRNAs
during the peri- and/or post-implantation stages of development,
multiple days after fertilization.
The lack of phenotypes following miRNA removal in early
mammalian development parallels findings observed in zebrafish:
the loss of both maternal and zygotic miRNAs in zebrafish first
manifests phenotypes relatively late in development (Giraldez et
al., 2005). Indeed, following maternal-zygotic knockdown of Dicer,
zebrafish gastrulate and only begin to manifest clear morphogenetic
defects during organogenesis. Zygotic loss alone allows
organogenesis to proceed normally (Wienholds et al., 2003).
Although the phenotypes are more severe in mouse, with zygotic
deletion of Dicer resulting in arrest prior to gastrulation (Bernstein
et al., 2003; Morita et al., 2007; Wang et al., 2007), these studies
show how early development can proceed normally in the absence
of miRNAs. A striking difference between the mouse and zebrafish
phenotypes is the lack of Dicer requirement in the maturing fish
oocyte. It will be important to determine whether this difference is
secondary to a lack of endo-siRNA function in the zebrafish egg.
An interesting possibility is that the loss of miRNA function
during pre-implantation development is a key component of the
dramatic reprogramming that occurs during this stage (Hemberger
et al., 2009). Indeed, recent profiling experiments suggest a shift
from a dominant presence of endo-siRNAs and piRNAs in the
oocyte to an miRNA majority as pre-implantation development
proceeds (Fig. 2) (Ohnishi et al., 2010). The transition occurs with
zygotic gene activation, which follows resetting of the epigenome.
This resetting occurs in a remarkably short window of time,
between fertilization and E2.5 in mice and slightly later in humans
(de Vries et al., 2008). Therefore, it is tempting to speculate that the
suppression of miRNA function enables this massive epigenomic
reprogramming in preparation for new gene expression.
miRNAs and siRNAs use distinct silencing
machinery
A surprising conclusion arising from the comparison of the Dicer
and Dgcr8 knockout phenotypes in mouse oocytes along with the
miRNA versus siRNA reporter assays is that the effector pathways
REVIEW Development 138 (9)
Box 1. P-bodies
P-bodies (processing bodies) are discrete cytoplasmic foci that
contain proteins involved in mRNA degradation. They are found in
eukaryotic cells as well as in somatic cells in plants and yeast
(Bashkirov et al., 1997; Cougot et al., 2004; Sheth and Parker,
2003; Xu et al., 2006). P-body proteins are required for diverse
post-transcriptional processes: mRNA decay, translational repression,
nonsense-mediated mRNA decay and RNAi-mediated repression. In
particular, all four Ago proteins (Eystathioy et al., 2003; Liu et al.,
2005; Sen and Blau, 2005), GW182 (Eystathioy et al., 2003) and
two RNA helicases RCK/p54 (Chu and Rana, 2006) and MOV10
(Meister et al., 2005) have been found in P-bodies, suggesting that
miRNA suppression is localized to the P-body. However, it has been
proposed that P-body formation is a consequence rather than the
cause of miRNA-mediated gene silencing (Eulalio et al., 2007), as
when siRNA or miRNA silencing pathways are blocked, P-bodies are
not formed (Eulalio et al., 2007). Interestingly, it has been shown
that P-body foci are dynamic, increasing or decreasing in size and
number depending on the global state of RNA turnover in yeast
(Sheth and Parker, 2003). Indeed, recent studies in mouse oocytes
and of early mouse embryonic development demonstrated that
these foci are regulated developmentally (Flemr et al., 2010;
Swetloff et al., 2009) In particular, it was shown that P-bodies are
lost in fully grown oocytes and during pre-implantation
development (Flemr et al., 2010; Swetloff et al., 2009). It will thus
be interesting to know how and when P-bodies are lost or re-stored
during early development.
PGCs
Sperm
MII oocyte Zygote
miRNA siRNA
miRNA miRNA/piRNA
Blastocyst2-cell 8-cell
piRNA/siRNA miRNA
Zygotic
genome
activation
(ZGA)
Fig. 2. Small RNA functions during germ cell and
early embryonic development. Piwi-interacting
RNAs (piRNAs) and microRNAs (miRNAs) are essential
in the developing male germline, whereas endogenous
small interfering RNAs (endo-siRNAs) play their most
crucial role in oocyte maturation. There is a transition
from endo-siRNAs or piRNAs to miRNAs during pre-
implantation development. PGCs, primordial germ
cells.
DEVELOPMENT
1657REVIEWDevelopment 138 (9)
for miRNA and endo-siRNA activity are separable (Fig. 1). That
is, although miRNA-based destabilization (by translational
inhibition) function is lost, siRNA cleavage function remains. In
Drosophila, Ago1, together with Loquacious, is the primary driver
of miRNA function, whereas Ago2, together with its partner R2D2,
is primarily responsible for siRNA function (Ghildiyal and Zamore,
2009). Mammals, by contrast, have four Argonaute proteins: Ago1-
4 (Siomi and Siomi, 2009). Ago2 is the only mammalian
Argonaute with slicer activity and hence the only Argonaute
protein able to perform siRNA cleavage. Indeed, deletion of Ago2
in oocytes produces a phenotype very similar to that of Dicer
(Kaneda et al., 2009). However, the Argonautes are highly
redundant in terms of miRNA activity. Although deletion of all four
Argonautes in ES cells results in complete loss of miRNA function,
re-introduction of any one of the four Argonautes can fully rescue
miRNA activity (Su et al., 2009). Therefore, unlike the situation in
Drosophila, siRNA and miRNA function in mammalian cells is
unlikely to be compartmentalized at the level of the Argonaute
proteins. Instead, proteins associated with or mechanisms
downstream of Argonautes must be influencing the specific loss of
miRNA function. It is unclear what these mechanisms might be. P-
bodies are lost concurrently with miRNA functional loss (Flemr
et al., 2010; Swetloff et al., 2009); therefore, studies of the
components of the P-body might provide hints.
miRNA versus endo-siRNA activity in other tissues
Microprocessor components have also been knocked out in other
tissues and the resulting phenotypes compared with corresponding
Dicer phenotypes. In particular, Dgcr8 has been knocked out in
skin and cardiomyocytes (Rao et al., 2009; Yi et al., 2009), and
Drosha has been knocked out in T cells (Chong et al., 2008). In
these cases, the microprocessor-null phenotypes were very similar
to those of the corresponding Dicer-null phenotypes. An exception
appears to be in the adult brain, where the deletion of Dicer in post-
mitotic neurons produces a more severe phenotype than does
Dgcr8 loss (J. E. Babiarz, R. Hsu, C. Melton, E. M. Ullian and
R.B., unpublished). However, deep sequencing failed to uncover
any evidence of endo-siRNAs in the brain (J. E. Babiarz, R. Hsu,
C. Melton, E. M. Ullian and R.B., unpublished). Instead, many
non-canonical miRNAs were found, suggesting that these small
RNAs might underlie the differences observed. These results
suggest that endo-siRNAs might be specific to oocytes and early
embryonic development.
Small RNAs in stem cells
Embryonic stem cells
Three self-renewing cell types can be derived from the late mouse
blastocyst: ES cells, which represent the pluripotent epiblast
lineage; trophoblast stem (TS) cells, which represent the
trophoblast lineage; and extra-embryonic endoderm (XEN) cells,
which represent the primitive endoderm lineage (Fig. 3) (Rossant,
2008). Insights into small RNA regulation of stem cell maintenance
and differentiation have been gained mostly from studies of ES
cells lacking either Dgcr8 or Dicer (Kanellopoulou et al., 2005;
Murchison et al., 2005; Wang et al., 2007). Both Dgcr8- and Dicer-
deficient ES cells exhibit proliferation and differentiation defects.
The proliferation phenotype is associated with accumulation of
cells in the G1 phase of the cell cycle, whereas the differentiation
defect is associated with an inability to silence the self-renewal
machinery (Wang et al., 2007). By adding back individual miRNAs
into Dgcr8 knockout ES cells, a large family of miRNAs, including
members of the Mir290 and Mir302 clusters, were found to rescue
the prolonged G1 phenotype (Wang et al., 2008). These miRNAs
were termed the ESCC miRNAs (ESC cell cycle regulating
miRNAs). Using a similar add-back strategy, the Let-7 (also known
as Mirlet7) miRNA family was shown to rescue the ability to
silence self-renewal (Melton et al., 2010). Specifically, the addition
of Let-7 into Dgcr8 knockout cells led to the loss of expression of
multiple markers of ES cells and blocked the ability of cells to
reform colonies, functionally proving a loss of their self-renewal
capacity. However, Let-7 only silenced self-renewal in the Dgcr8
knockout, not wild-type ES cells, suggesting that miRNAs
normally expressed in ES cells suppress the capacity of Let-7 to
induce differentiation. Indeed, simultaneous introduction of the
ESCC miRNAs into the Dgcr8 knockout ES cells suppressed the
capacity of Let-7 to induce differentiation. Microarray profiling of
mRNAs and bioinformatic analyses showed that these two
antagonizing miRNA families function by having opposing effects
on members of the pluripotency regulatory network, including
Myc, Lin28 and Sall4 along with others (Fig. 4) Taken together,
these findings show that the two miRNA families, the Let-7 and
ESCC miRNAs, play opposing roles in controlling the balance
between ES cell self-renewal and differentiation (Melton et al.,
2010).
Endo-siRNAs have also been identified in ES cells (Babiarz et
al., 2008). Interestingly, there is no overlap between the specific
endo-siRNAs expressed in oocytes and those expressed in ES cells,
showing that they are developmentally regulated and likely to have
distinct functions (Babiarz et al., 2008). However, the role of ES
cell endo-siRNAs is unclear. A hint comes once again from the
deletion of members of the biogenesis and effector pathways.
Deletions of Dicer or of all four Argonaute genes have more severe
phenotypes than deletion of Dgcr8 (Kanellopoulou et al., 2005;
Murchison et al., 2005; Su et al., 2009; Wang et al., 2007). In
particular pan-Argonaute-deficient ES cells undergo apoptosis.
Dicer-null ES cells can survive, but during their derivation they go
through a phase of arrest, with escaper cells eventually
proliferating. By contrast, Dgcr8-null ES cells survive and show
Blastocyst
Epiblast
Trophectoderm TS cells
Primitive endoderm XEN cells
ES cells
Stem cells Gene deleted
Dicer
Dicer
Dicer or
Dgcr8
Phenotype Reference(s)
Proliferation defects
Proliferation defects
Kanellopoulou et al., 2005;
Murchison et al., 2005; Wang et al., 2007
Spruce et al., 2010
Spruce et al., 2010
Proliferation and
differentiation defects
Fig. 3. Small RNA functions in stem cells derived from the mouse blastocyst. By the time of implantation, the mammalian blastocyst has
developed three different cell lineages: trophectoderm, primitive endoderm and epiblast (shown on left). Three distinct self-renewing cell lines can
be derived from these lineages: trophoblast stem (TS) cells, extra-embryonic endoderm (XEN) cells and embryonic stem (ES) cells. Studies of stem
cell lines lacking either Dgcr8 or Dicer provide insights into small RNA-mediated regulation of stem cell maintenance, proliferation and
differentiation (right).
DEVELOPMENT
1658
no evidence of arrest during derivation. These differences in
phenotype suggest a likely role for endo-siRNAs in ES cells. By
contrast, a role for piRNAs in the embryonic stem cells is doubtful
as their levels are greatly diminished relative to those observed in
the germline (Ohnishi et al., 2010), and the knockout of the Piwi
genes in mice show no embryonic phenotypes (Carmell et al.,
2007; Deng and Lin, 2002; Kuramochi-Miyagawa et al., 2004).
Trophoblast stem cells
During the course of early embryogenesis, the separation of the
trophectoderm and ICM lineages is the first known definitive
differentiation event. Fundamental insights into the molecular
control of trophectoderm determination and differentiation have
been made in the past decade (Chen et al., 2010; Douglas et al.,
2009; Ralston and Rossant, 2005). A number of transcription factors
and signaling pathways have been identified as crucial players. Even
before the formation of the blastocyst, two transcription factors,
Oct4 (Pou5f1 – Mouse Genome Informatics) and Cdx2, act
antagonistically to establish the boundaries between the
trophectoderm and inner cell mass (Niwa et al., 2005). Oct4 is
expressed throughout pre-implantation development whereas Cdx2
is expressed starting around the time of morula compaction
(Dietrich and Hiiragi, 2007). At first, Cdx2 is co-expressed with
Oct4 in the cells of the morula, but its expression is then seen to
segregate to the future trophectoderm cells. Cdx2 protein binds
directly to Oct4 protein resulting in reciprocal inhibition of their
target genes (Niwa et al., 2005). Dominance of one protein over the
other eventually leads to the choice between the two lineages: inner
cell mass versus trophectoderm. A second transcription factor,
Eomes, which acts independently of Cdx2, is also essential early in
trophectoderm determination and maintenance, but little more is
known about its function (Russ et al., 2000). Following the
formation of the blastocyst, the inner cell mass produces fibroblast
growth factor 4 (Fgf4), which signals through the FGF receptor 2
(Fgfr2), to promote proliferation of the overlying polar
trophectoderm, thereby allowing the polar trophectoderm to provide
an ongoing source of trophoblasts both to the mural trophectoderm
and future placenta (Nichols et al., 1998; Tanaka et al., 1998).
In contrast to the ES cell studies, little is known about the role
of small RNAs in trophectoderm specification. miRNA expression
profiling of ES cells, ES cell-derived TS cells and progressive
stages of pre-implantation embryos, has identified a subset of
miRNAs that might play a role in trophectoderm specification:
Mir297, Mir96, Mir21, Mir29c, Let-7, Mir214, Mir125a, and
Mir424 (Viswanathan et al., 2009) Moreover, studies analyzing the
phenotype of Dicer-deficient embryos during early post-
implantation stages have shown an essential role for small RNAs
in trophectoderm development (Spruce et al., 2010). In particular,
expression of the TS cell markers Eomes, Cdx2 and Esrrb was
greatly downregulated in Dicer knockout embryos. Similar to ES
cells, Dicer knockout TS cells show proliferation defects, with an
accumulation of cells in G1. This finding is consistent with the fact
that the Mir290 cluster is also highly expressed in TS cells
(Houbaviy et al., 2005). Indeed, as seen in ES cells, a number of
inhibitors of the cyclin E/Cdk2 pathways were upregulated in TS
cells following miRNA loss (Spruce et al., 2010). Dicer removal in
XEN cells also influenced self-renewal and proliferation, but
potentially through different pathways. In particular, regulation of
ERK activity appears to be an important player in the phenotype.
Taken together, early experiments in TS and XEN cells suggest
overlap in miRNA roles across the three stem cell populations of
the embryo.
Induced pluripotent stem cells
In 2006, Yamanaka and co-workers showed that somatic cells
could be reprogrammed into induced pluripotent stem (iPS) cells
by retroviral introduction of genes encoding four transcription
factors: Oct3/4 (Pou5F1 – Mouse Genome Informatics), Klf4, Sox2
and Myc (Takahashi and Yamanaka, 2006). With improvements in
the methods, these iPS cells have become increasingly similar to
ES cells both in their self-renewal and differentiation potential (for
a review, see Amabile and Meissner, 2009). The realisation of the
importance and therapeutic potential of iPS cells has opened a new
era in regenerative medicine (Yamanaka, 2009).
A role for miRNAs in iPS cell production has recently been
uncovered. In particular, ESCC miRNAs can promote the de-
differentiation of somatic cells to iPS cells. They can replace Myc
and, based on chromatin immunoprecipitation (ChIP) sequence
data, function downstream of Myc (Judson et al., 2009).
Interestingly, ESCC miRNAs also upregulate Myc, albeit indirectly
(Melton et al., 2010). Therefore, ESCC miRNAs and Myc form a
self-reinforcing loop that maintains ES cell self-renewal and even
promotes de-differentiation. Furthermore, inhibition of Let-7
function, either through overexpression of Lin28, which blocks
Let-7 biogenesis, or through antagomirs, which directly target
mature Let-7, is able to promote iPS cell production (Melton et al.,
2010; Yu et al., 2007). This result is consistent with Let-7’s
capacity to suppress Myc and many of the downstream targets of
the pluripotency network of transcription factors. These findings
further emphasize the role of these miRNAs in regulating the
switch between self-renewal and differentiation.
Emerging mechanisms of miRNA regulation
In the past decade, much progress has been made in identifying
miRNAs, understanding miRNA biogenesis and predicting
miRNA targets. Furthermore, it is becoming evident that miRNAs
can exert their effects through single or multiple targets, allowing
them to regulate development, normal physiology, and
pathological processes. However, there remains much to be
learned about miRNA biology. A question of increasing interest is
REVIEW Development 138 (9)
Stem cells
Let-7 miRNA
ESCC miRNA
Stemness factors
(e.g. Lin28, Myc)
Oct4, Sox2,
Nanog, Tcf3
Let-7 miRNA
ESCC miRNA
Stemness factors
(e.g. Lin28, Myc)
Oct4, Sox2,
Nanog, Tcf3
Differentiated cells
Fig. 4. The opposing roles of ESCC and Let-7 microRNAs (miRNAs)
in the switch between self-renewal and differentiation. In mouse
ES cells (left panel), ESCC miRNAs (green) and stemness factors are
highly expressed. ESCC miRNAs are regulated by the core ES cell
transcription factors such as Oct4, Sox2, Nanog, Tcf3 and Myc. Upon
differentiation (right panel), the expression of Let-7 miRNAs (red)
increases and helps to repress stemness factors such as Lin28, Myc and
Sall4.
DEVELOPMENT
1659REVIEWDevelopment 138 (9)
how miRNAs themselves are regulated. However, the answers to
this question are almost as diverse as the miRNAs identified and,
therefore, below we discuss only those mechanisms that regulate
small RNAs with known roles in germ cells and in the gastrulating
embryo [for a broader review on the topic, see Krol et al. (Krol et
al., 2010)].
Most miRNAs are expressed through an RNA polymerase II
mechanism, and regulation of their expression thus occurs through
the common mechanisms that regulate developmental genes. For
example, the expression of ESCC miRNAs is regulated by the core
ES cell transcription factors Oct4, Sox2, Nanog, Tcf3 and Myc
(Judson et al., 2009; Marson et al., 2008). Furthermore, and as
observed for other pluripotency genes, the expression of ESCC
miRNAs is regulated by epigenetic modifications, which include
activating and suppressive histone marks (Judson et al., 2009;
Marson et al., 2008).
miRNAs are also regulated post-transcriptionally at the level of
their biogenesis and stability. For example, Lin28, an RNA binding
protein, regulates the biogenesis of the Let-7 family of miRNAs
(Heo et al., 2008; Newman et al., 2008; Piskounova et al., 2008;
Rybak et al., 2008; Viswanathan et al., 2008). Lin28 inhibits Dicer
cleavage and destabilizes the pre-miRNA form of Let-7 (Heo et al.,
2008; Rybak et al., 2008). The latter is achieved through Lin28
interacting with the loop region of Let-7 and directing a terminal
uridyl transferase (TUTase) to polyuridylate the 3end of the pre-
Let-7 miRNA, leading to pre-Let-7 degradation (Hagan et al.,
2009; Heo et al., 2009).
miRNA function can also be regulated by interactions with their
downstream targets. For example, the RNA binding protein dead
end (DND1) blocks the ability of MIR221 and MIR372 to
suppress p27 (CDKN1B – Human Gene Nomenclature Database)
and LATS2, respectively, in human cells (Kedde et al., 2007).
DND1 is essential for germline development in zebrafish and
mouse, where it probably plays a similar role to that seen in the
human cell lines (Kedde et al., 2007; Slanchev et al., 2009).
Another germline RNA binding protein, Dazl (deleted in
azoospermia-like), has also been shown to have a potential role in
regulating the interactions between miRNAs and mRNA targets in
zebrafish (Takeda et al., 2009). In particular, Dazl appears to
inhibit miRNA induced de-adenylation by binding to the 3UTRs
of specific miRNA targets.
Future directions
Over the past decade we have learned a great deal about the nature
of small RNAs, their biogenesis and their expression across tissues.
However, we are just beginning to appreciate how these RNAs fit
in to the overall molecular network of the cell. Indeed, we have a
very poor understanding of how the many mRNA targets of
individual miRNAs work together to influence a specific cellular
outcome. To date, the field has mostly limited the analysis of
miRNA targeting in biological processes to a small number of
targets, probably not reflecting the true nature of miRNA function.
Future studies involving more systematic approaches should teach
us about how miRNAs are involved in the control of development,
tissue physiology and disease.
Furthermore, there remains much to be learned about the
transcriptional, epigenetic and post-transcriptional regulation of
miRNAs. miRNA targets are likely to be heavily influenced by
cellular context. The influence of cellular context can, in part, be
explained by differences in expression of those targets. However,
a large part will probably be attributable to the regulation of
miRNA biogenesis and the interaction of miRNAs with their target
mRNAs. The most extreme example is the global suppression of
miRNAs in the oocyte. The reason and mechanistic details of this
global suppression remain unclear. How the global or focused
regulation of miRNA maturation and targeting fit in with the timing
and regulation of cell fate decisions and their conservation across
species should be a fruitful area of research. Finally, it will be
important to study the relationship of these regulatory mechanisms
to the biology and treatment of disease, including within the field
of cellular reprogramming.
Acknowledgements
We thank Matthew Cook, Raga Krishnakumar and Ronald Parchem for helpful
comments on the manuscript. We apologize to those colleagues whose work
could not be cited directly owing to space constraint. R.B. is funded by the
National Institutes of Health, California Institute of Regenerative Medicine
(CIRM), the American Health Assistance Foundation (formerly Stem Cell
Research Foundation) and the Pew Charitable Trust. Further funds to R.B. came
from the NICHD/NIH through a pilot project funded within a cooperative
agreement as part of the Specialized Cooperative Centers Program in
Reproduction and Infertility Research. Deposited in PMC for release after 12
months.
Competing interests statement
The authors declare no competing financial interests.
References
Abe, K., Inoue, A., Suzuki, M. G. and Aoki, F. (2010). Global gene silencing is
caused by the dissociation of RNA polymerase II from DNA in mouse oocytes. J.
Reprod. Dev. 56, 502-507.
Amabile, G. and Meissner, A. (2009). Induced pluripotent stem cells: current
progress and potential for regenerative medicine. Trends Mol. Med. 15, 59-68.
Aravin, A., Gaidatzis, D., Pfeffer, S., Lagos-Quintana, M., Landgraf, P.,
Iovino, N., Morris, P., Brownstein, M. J., Kuramochi-Miyagawa, S.,
Nakano, T. et al. (2006). A novel class of small RNAs bind to MILI protein in
mouse testes. Nature 442, 203-207.
Aravin, A. A., Sachidanandam, R., Girard, A., Fejes-Toth, K. and Hannon, G.
J. (2007). Developmentally regulated piRNA clusters implicate MILI in transposon
control. Science 316, 744-747.
Aravin, A. A., Sachidanandam, R., Bourc’his, D., Schaefer, C., Pezic, D., Toth,
K. F., Bestor, T. and Hannon, G. J. (2008). A piRNA pathway primed by
individual transposons is linked to de novo DNA methylation in mice. Mol. Cell
31, 785-799.
Babiarz, J. E. and Blelloch, R. (2009). Small RNAs - their biogenesis, regulation
and function in embryonic stem cells. In StemBook (ed. The Stem Cell Research
Community). doi/10.3824/stembook.1.47.1.
Babiarz, J. E., Ruby, J. G., Wang, Y., Bartel, D. P. and Blelloch, R. (2008).
Mouse ES cells express endogenous shRNAs, siRNAs, and other microprocessor-
independent, Dicer-dependent small RNAs. Genes Dev. 22, 2773-2785.
Bashkirov, V. I., Scherthan, H., Solinger, J. A., Buerstedde, J. M. and Heyer,
W. D . (1997). A mouse cytoplasmic exoribonuclease (mXRN1p) with preference
for G4 tetraplex substrates. J. Cell Biol. 136, 761-773.
Bernstein, E., Caudy, A. A., Hammond, S. M. and Hannon, G. J. (2001). Role
for a bidentate ribonuclease in the initiation step of RNA interference. Nature
409, 363-366.
Bernstein, E., Kim, S. Y., Carmell, M. A., Murchison, E. P., Alcorn, H., Li, M. Z.,
Mills, A. A., Elledge, S. J., Anderson, K. V. and Hannon, G. J. (2003). Dicer is
essential for mouse development. Nat. Genet. 35, 215-217.
Bjork, J. K., Sandqvist, A., Elsing, A. N., Kotaja, N. and Sistonen, L. (2010).
miR-18, a member of Oncomir-1, targets heat shock transcription factor 2 in
spermatogenesis. Development 137, 3177-3184.
Bouhallier, F., Allioli, N., Lavial, F., Chalmel, F., Perrard, M. H., Durand, P.,
Samarut, J., Pain, B. and Rouault, J. P. (2010). Role of miR-34c microRNA in
the late steps of spermatogenesis. RNA 16, 720-731.
Brennecke, J., Aravin, A. A., Stark, A., Dus, M., Kellis, M., Sachidanandam,
R. and Hannon, G. J. (2007). Discrete small RNA-generating loci as master
regulators of transposon activity in Drosophila. Cell 128, 1089-1103.
Carmell, M. A., Girard, A., van de Kant, H. J., Bourc’his, D., Bestor, T. H., de
Rooij, D. G. and Hannon, G. J. (2007). MIWI2 is essential for spermatogenesis
and repression of transposons in the mouse male germline. Dev. Cell 12, 503-
514.
Chen, L., Wang, D., Wu, Z., Ma, L. and Daley, G. Q. (2010). Molecular basis of
the first cell fate determination in mouse embryogenesis. Cell Res. 20, 982-993.
Chong, M. M., Rasmussen, J. P., Rudensky, A. Y. and Littman, D. R. (2008).
The RNAseIII enzyme Drosha is critical in T cells for preventing lethal
inflammatory disease. J. Exp. Med. 205, 2005-2017.
Chu, C. Y. and Rana, T. M. (2006). Translation repression in human cells by
microRNA-induced gene silencing requires RCK/p54. PLoS Biol. 4, e210.
DEVELOPMENT
1660
Chung, W. J., Okamura, K., Martin, R. and Lai, E. C. (2008). Endogenous RNA
interference provides a somatic defense against Drosophila transposons. Curr.
Biol. 18, 795-802.
Cook, M. S., Munger, S. C., Nadeau, J. H. and Capel, B. (2010). Regulation of
male germ cell cycle arrest and differentiation by DND1 is modulated by genetic
background. Development 138, 23-32.
Cougot, N., Babajko, S. and Seraphin, B. (2004). Cytoplasmic foci are sites of
mRNA decay in human cells. J. Cell Biol. 165, 31-40.
Czech, B., Malone, C. D., Zhou, R., Stark, A., Schlingeheyde, C., Dus, M.,
Perrimon, N., Kellis, M., Wohlschlegel, J. A., Sachidanandam, R. et al.
(2008). An endogenous small interfering RNA pathway in Drosophila. Nature
453, 798-802.
de Vries, W. N., Evsikov, A. V., Brogan, L. J., Anderson, C. P., Graber, J. H.,
Knowles, B. B. and Solter, D. (2008). Reprogramming and differentiation in
mammals: motifs and mechanisms. Cold Spring Harb. Symp. Quant. Biol. 73,
33-38.
Deng, W. and Lin, H. (2002). miwi, a murine homolog of piwi, encodes a
cytoplasmic protein essential for spermatogenesis. Dev. Cell 2, 819-830.
Denli, A. M., Tops, B. B., Plasterk, R. H., Ketting, R. F. and Hannon, G. J.
(2004). Processing of primary microRNAs by the Microprocessor complex. Nature
432, 231-235.
Dietrich, J. E. and Hiiragi, T. (2007). Stochastic patterning in the mouse pre-
implantation embryo. Development 134, 4219-4231.
Doench, J. G., Petersen, C. P. and Sharp, P. A. (2003). siRNAs can function as
miRNAs. Genes Dev. 17, 438-442.
Douglas, G. C., VandeVoort, C. A., Kumar, P., Chang, T. C. and Golos, T. G.
(2009). Trophoblast stem cells: models for investigating trophectoderm
differentiation and placental development. Endocr. Rev. 30, 228-240.
Eulalio, A., Behm-Ansmant, I., Schweizer, D. and Izaurralde, E. (2007). P-body
formation is a consequence, not the cause, of RNA-mediated gene silencing.
Mol. Cell. Biol. 27, 3970-3981.
Eystathioy, T., Jakymiw, A., Chan, E. K., Seraphin, B., Cougot, N. and Fritzler,
M. J. (2003). The GW182 protein colocalizes with mRNA degradation associated
proteins hDcp1 and hLSm4 in cytoplasmic GW bodies. RNA 9, 1171-1173.
Fabian, M. R., Sonenberg, N. and Filipowicz, W. (2010). Regulation of mRNA
translation and stability by microRNAs. Annu. Rev. Biochem. 79, 351-379.
Filipowicz, W. (2005). RNAi: the nuts and bolts of the RISC machine. Cell 122, 17-
20.
Flemr, M., Ma, J., Schultz, R. M. and Svoboda, P. (2010). P-body loss is
concomitant with formation of a messenger RNA storage domain in mouse
oocytes. Biol. Reprod. 82, 1008-1017.
Ghildiyal, M. and Zamore, P. D. (2009). Small silencing RNAs: an expanding
universe. Nat. Rev. Genet. 10, 94-108.
Ghildiyal, M., Seitz, H., Horwich, M. D., Li, C., Du, T., Lee, S., Xu, J., Kittler, E.
L., Zapp, M. L., Weng, Z. et al. (2008). Endogenous siRNAs derived from
transposons and mRNAs in Drosophila somatic cells. Science 320, 1077-1081.
Giraldez, A. J., Cinalli, R. M., Glasner, M. E., Enright, A. J., Thomson, J. M.,
Baskerville, S., Hammond, S. M., Bartel, D. P. and Schier, A. F. (2005).
MicroRNAs regulate brain morphogenesis in zebrafish. Science 308, 833-838.
Girard, A., Sachidanandam, R., Hannon, G. J. and Carmell, M. A. (2006). A
germline-specific class of small RNAs binds mammalian Piwi proteins. Nature
442, 199-202.
Gregory, R. I., Yan, K. P., Amuthan, G., Chendrimada, T., Doratotaj, B.,
Cooch, N. and Shiekhattar, R. (2004). The Microprocessor complex mediates
the genesis of microRNAs. Nature 432, 235-240.
Grivna, S. T., Beyret, E., Wang, Z. and Lin, H. (2006). A novel class of small
RNAs in mouse spermatogenic cells. Genes Dev. 20, 1709-1714.
Gunawardane, L. S., Saito, K., Nishida, K. M., Miyoshi, K., Kawamura, Y.,
Nagami, T., Siomi, H. and Siomi, M. C. (2007). A slicer-mediated mechanism
for repeat-associated siRNA 5end formation in Drosophila. Science 315, 1587-
1590.
Hagan, J. P., Piskounova, E. and Gregory, R. I. (2009). Lin28 recruits the TUTase
Zcchc11 to inhibit let-7 maturation in mouse embryonic stem cells. Nat. Struct.
Mol. Biol. 16, 1021-1025.
Han, J., Lee, Y., Yeom, K. H., Kim, Y. K., Jin, H. and Kim, V. N. (2004). The
Drosha-DGCR8 complex in primary microRNA processing. Genes Dev. 18, 3016-
3027.
Han, J., Lee, Y., Yeom, K. H., Nam, J. W., Heo, I., Rhee, J. K., Sohn, S. Y., Cho,
Y., Zhang, B. T. and Kim, V. N. (2006). Molecular basis for the recognition of
primary microRNAs by the Drosha-DGCR8 complex. Cell 125, 887-901.
Hayashi, K., Chuva de Sousa Lopes, S. M., Kaneda, M., Tang, F., Hajkova, P.,
Lao, K., O’Carroll, D., Das, P. P., Tarakhovsky, A., Miska, E. A. et al. (2008).
MicroRNA biogenesis is required for mouse primordial germ cell development
and spermatogenesis. PLoS One 3, e1738.
Hemberger, M., Dean, W. and Reik, W. (2009). Epigenetic dynamics of stem
cells and cell lineage commitment: digging Waddington’s canal. Nat. Rev. Mol.
Cell Biol. 10, 526-537.
Heo, I., Joo, C., Cho, J., Ha, M., Han, J. and Kim, V. N. (2008). Lin28 mediates
the terminal uridylation of let-7 precursor MicroRNA. Mol. Cell 32, 276-284.
Heo, I., Joo, C., Kim, Y. K., Ha, M., Yoon, M. J., Cho, J., Yeom, K. H., Han, J.
and Kim, V. N. (2009). TUT4 in concert with Lin28 suppresses microRNA
biogenesis through pre-microRNA uridylation. Cell 138, 696-708.
Houbaviy, H. B., Dennis, L., Jaenisch, R. and Sharp, P. A. (2005).
Characterization of a highly variable eutherian microRNA gene. RNA 11, 1245-
1257.
Houwing, S., Kamminga, L. M., Berezikov, E., Cronembold, D., Girard, A.,
van den Elst, H., Filippov, D. V., Blaser, H., Raz, E., Moens, C. B. et al.
(2007). A role for Piwi and piRNAs in germ cell maintenance and transposon
silencing in Zebrafish. Cell 129, 69-82.
Hutvagner, G., McLachlan, J., Pasquinelli, A. E., Balint, E., Tuschl, T. and
Zamore, P. D. (2001). A cellular function for the RNA-interference enzyme Dicer
in the maturation of the let-7 small temporal RNA. Science 293, 834-838.
Judson, R. L., Babiarz, J. E., Venere, M. and Blelloch, R. (2009). Embryonic
stem cell-specific microRNAs promote induced pluripotency. Nat. Biotechnol. 27,
459-461.
Kaneda, M., Tang, F., O’Carroll, D., Lao, K. and Surani, M. A. (2009). Essential
role for Argonaute2 protein in mouse oogenesis. Epigenetics Chromatin 2, 9.
Kanellopoulou, C., Muljo, S. A., Kung, A. L., Ganesan, S., Drapkin, R.,
Jenuwein, T., Livingston, D. M. and Rajewsky, K. (2005). Dicer-deficient
mouse embryonic stem cells are defective in differentiation and centromeric
silencing. Genes Dev. 19, 489-501.
Kawamura, Y., Saito, K., Kin, T., Ono, Y., Asai, K., Sunohara, T., Okada, T. N.,
Siomi, M. C. and Siomi, H. (2008). Drosophila endogenous small RNAs bind to
Argonaute 2 in somatic cells. Nature 453, 793-797.
Kedde, M., Strasser, M. J., Boldajipour, B., Oude Vrielink, J. A., Slanchev, K.,
le Sage, C., Nagel, R., Voorhoeve, P. M., van Duijse, J., Orom, U. A. et al.
(2007). RNA-binding protein Dnd1 inhibits microRNA access to target mRNA.
Cell 131, 1273-1286.
Ketting, R. F., Fischer, S. E., Bernstein, E., Sijen, T., Hannon, G. J. and
Plasterk, R. H. (2001). Dicer functions in RNA interference and in synthesis of
small RNA involved in developmental timing in C. elegans. Genes Dev. 15, 2654-
2659.
Kim, V. N., Han, J. and Siomi, M. C. (2009). Biogenesis of small RNAs in animals.
Nat. Rev. Mol. Cell Biol. 10, 126-139.
Klattenhoff, C. and Theurkauf, W. (2008). Biogenesis and germline functions of
piRNAs. Development 135, 3-9.
Knight, S. W. and Bass, B. L. (2001). A role for the RNase III enzyme DCR-1 in
RNA interference and germ line development in Caenorhabditis elegans. Science
293, 2269-2271.
Krol, J., Loedige, I. and Filipowicz, W. (2010). The widespread regulation of
microRNA biogenesis, function and decay. Nat. Rev. Genet. 11, 597-610.
Kuramochi-Miyagawa, S., Kimura, T., Ijiri, T. W., Isobe, T., Asada, N., Fujita,
Y., Ikawa, M., Iwai, N., Okabe, M., Deng, W. et al. (2004). Mili, a
mammalian member of piwi family gene, is essential for spermatogenesis.
Development 131, 839-849.
Kuramochi-Miyagawa, S., Watanabe, T., Gotoh, K., Totoki, Y., Toyoda, A.,
Ikawa, M., Asada, N., Kojima, K., Yamaguchi, Y., Ijiri, T. W. et al. (2008).
DNA methylation of retrotransposon genes is regulated by Piwi family members
MILI and MIWI2 in murine fetal testes. Genes Dev. 22, 908-917.
Landthaler, M., Yalcin, A. and Tuschl, T. (2004). The human DiGeorge syndrome
critical region gene 8 and Its D. melanogaster homolog are required for miRNA
biogenesis. Curr. Biol. 14, 2162-2167.
Lau, N. C., Seto, A. G., Kim, J., Kuramochi-Miyagawa, S., Nakano, T., Bartel,
D. P. and Kingston, R. E. (2006). Characterization of the piRNA complex from
rat testes. Science 313, 363-367.
Lee, T. L., Pang, A. L., Rennert, O. M. and Chan, W. Y. (2009). Genomic
landscape of developing male germ cells. Birth Defects Res. C Embryo Today 87,
43-63.
Lee, Y., Ahn, C., Han, J., Choi, H., Kim, J., Yim, J., Lee, J., Provost, P.,
Radmark, O., Kim, S. et al. (2003). The nuclear RNase III Drosha initiates
microRNA processing. Nature 425, 415-419.
Liu, J., Valencia-Sanchez, M. A., Hannon, G. J. and Parker, R. (2005).
MicroRNA-dependent localization of targeted mRNAs to mammalian P-bodies.
Nat. Cell Biol. 7, 719-723.
Lu, R., Markowetz, F., Unwin, R. D., Leek, J. T., Airoldi, E. M., MacArthur, B.
D., Lachmann, A., Rozov, R., Ma’ayan, A., Boyer, L. A. et al. (2009).
Systems-level dynamic analyses of fate change in murine embryonic stem cells.
Nature 462, 358-362.
Ma, J., Flemr, M., Stein, P., Berninger, P., Malik, R., Zavolan, M., Svoboda, P.
and Schultz, R. M. (2010). MicroRNA activity is suppressed in mouse oocytes.
Curr. Biol. 20, 265-270.
Maatouk, D. M., Loveland, K. L., McManus, M. T., Moore, K. and Harfe, B. D.
(2008). Dicer1 is required for differentiation of the mouse male germline. Biol.
Reprod. 79, 696-703.
Malone, C. D. and Hannon, G. J. (2009). Small RNAs as guardians of the
genome. Cell 136, 656-668.
Marson, A., Levine, S. S., Cole, M. F., Frampton, G. M., Brambrink, T.,
Johnstone, S., Guenther, M. G., Johnston, W. K., Wernig, M., Newman, J.
REVIEW Development 138 (9)
DEVELOPMENT
1661REVIEWDevelopment 138 (9)
et al. (2008). Connecting microRNA genes to the core transcriptional regulatory
circuitry of embryonic stem cells. Cell 134, 521-533.
Matova, N. and Cooley, L. (2001). Comparative aspects of animal oogenesis.
Dev. Biol. 231, 291-320.
Meister, G., Landthaler, M., Peters, L., Chen, P. Y., Urlaub, H., Luhrmann, R.
and Tuschl, T. (2005). Identification of novel argonaute-associated proteins.
Curr. Biol. 15, 2149-2155.
Melton, C., Judson, R. L. and Blelloch, R. (2010). Opposing microRNA families
regulate self-renewal in mouse embryonic stem cells. Nature 463, 621-626.
Moazed, D. (2009). Small RNAs in transcriptional gene silencing and genome
defence. Nature 457, 413-420.
Morita, S., Horii, T., Kimura, M., Goto, Y., Ochiya, T. and Hatada, I. (2007).
One Argonaute family member, Eif2c2 (Ago2), is essential for development and
appears not to be involved in DNA methylation. Genomics 89, 687-696.
Murchison, E. P., Partridge, J. F., Tam, O. H., Cheloufi, S. and Hannon, G. J.
(2005). Characterization of Dicer-deficient murine embryonic stem cells. Proc.
Natl. Acad. Sci. USA 102, 12135-12140.
Murchison, E. P., Stein, P., Xuan, Z., Pan, H., Zhang, M. Q., Schultz, R. M. and
Hannon, G. J. (2007). Critical roles for Dicer in the female germline. Genes Dev.
21, 682-693.
Newman, M. A., Thomson, J. M. and Hammond, S. M. (2008). Lin-28
interaction with the Let-7 precursor loop mediates regulated microRNA
processing. RNA 14, 1539-1549.
Nichols, J., Zevnik, B., Anastassiadis, K., Niwa, H., Klewe-Nebenius, D.,
Chambers, I., Scholer, H. and Smith, A. (1998). Formation of pluripotent stem
cells in the mammalian embryo depends on the POU transcription factor Oct4.
Cell 95, 379-391.
Niwa, H., Toyooka, Y., Shimosato, D., Strumpf, D., Takahashi, K., Yagi, R.
and Rossant, J. (2005). Interaction between Oct3/4 and Cdx2 determines
trophectoderm differentiation. Cell 123, 917-929.
Ohnishi, Y., Totoki, Y., Toyoda, A., Watanabe, T., Yamamoto, Y., Tokunaga,
K., Sakaki, Y., Sasaki, H. and Hohjoh, H. (2010). Small RNA class transition
from siRNA/piRNA to miRNA during pre-implantation mouse development.
Nucleic Acids Res. 38, 5141-5151.
Okamura, K., Hagen, J. W., Duan, H., Tyler, D. M. and Lai, E. C. (2007). The
mirtron pathway generates microRNA-class regulatory RNAs in Drosophila. Cell
130, 89-100.
Okamura, K., Balla, S., Martin, R., Liu, N. and Lai, E. C. (2008a). Two distinct
mechanisms generate endogenous siRNAs from bidirectional transcription in
Drosophila melanogaster. Nat. Struct. Mol. Biol. 15, 998.
Okamura, K., Chung, W. J., Ruby, J. G., Guo, H., Bartel, D. P. and Lai, E. C.
(2008b). The Drosophila hairpin RNA pathway generates endogenous short
interfering RNAs. Nature 453, 803-806.
Parker, R. and Sheth, U. (2007). P bodies and the control of mRNA translation
and degradation. Mol. Cell 25, 635-646.
Piskounova, E., Viswanathan, S. R., Janas, M., LaPierre, R. J., Daley, G. Q.,
Sliz, P. and Gregory, R. I. (2008). Determinants of microRNA processing
inhibition by the developmentally regulated RNA-binding protein Lin28. J. Biol.
Chem. 283, 21310-21314.
Ralston, A. and Rossant, J. (2005). Genetic regulation of stem cell origins in the
mouse embryo. Clin. Genet. 68, 106-112.
Rao, P. K., Toyama, Y., Chiang, H. R., Gupta, S., Bauer, M., Medvid, R.,
Reinhardt, F., Liao, R., Krieger, M., Jaenisch, R. et al. (2009). Loss of cardiac
microRNA-mediated regulation leads to dilated cardiomyopathy and heart
failure. Circ. Res. 105, 585-594.
Rossant, J. (2008). Stem cells and early lineage development. Cell 132, 527-531.
Ruby, J. G., Jan, C. H. and Bartel, D. P. (2007). Intronic microRNA precursors that
bypass Drosha processing. Nature 448, 83-86.
Russ, A. P., Aparicio, S. A. and Carlton, M. B. (2000). Open-source work even
more vital to genome project than to software. Nature 404, 809.
Rybak, A., Fuchs, H., Smirnova, L., Brandt, C., Pohl, E. E., Nitsch, R. and
Wulczyn, F. G. (2008). A feedback loop comprising lin-28 and let-7 controls
pre-let-7 maturation during neural stem-cell commitment. Nat. Cell Biol. 10,
987-993.
Sen, G. L. and Blau, H. M. (2005). Argonaute 2/RISC resides in sites of mammalian
mRNA decay known as cytoplasmic bodies. Nat. Cell Biol. 7, 633-636.
Sheth, U. and Parker, R. (2003). Decapping and decay of messenger RNA occur
in cytoplasmic processing bodies. Science 300, 805-808.
Siomi, H. and Siomi, M. C. (2009). On the road to reading the RNA-interference
code. Nature 457, 396-404.
Slanchev, K., Stebler, J., Goudarzi, M., Cojocaru, V., Weidinger, G. and Raz,
E. (2009). Control of Dead end localization and activity-implications for the
function of the protein in antagonizing miRNA function. Mech. Dev. 126, 270-
277.
Spruce, T., Pernaute, B., Di-Gregorio, A., Cobb, B. S., Merkenschlager, M.,
Manzanares, M. and Rodriguez, T. A. (2010). An early developmental role for
miRNAs in the maintenance of extraembryonic stem cells in the mouse embryo.
Dev. Cell 19, 207-219.
Su, H., Trombly, M. I., Chen, J. and Wang, X. (2009). Essential and overlapping
functions for mammalian Argonautes in microRNA silencing. Genes Dev. 23,
304-317.
Suh, N., Baehner, L., Moltzahn, F., Melton, C., Shenoy, A., Chen, J. and
Blelloch, R. (2010). MicroRNA function is globally suppressed in mouse oocytes
and early embryos. Curr. Biol. 20, 271-277.
Swetloff, A., Conne, B., Huarte, J., Pitetti, J. L., Nef, S. and Vassalli, J. D.
(2009). Dcp1-bodies in mouse oocytes. Mol. Biol. Cell 20, 4951-4961.
Takahashi, K. and Yamanaka, S. (2006). Induction of pluripotent stem cells from
mouse embryonic and adult fibroblast cultures by defined factors. Cell 126, 663-
676.
Takeda, Y., Mishima, Y., Fujiwara, T., Sakamoto, H. and Inoue, K. (2009).
DAZL relieves miRNA-mediated repression of germline mRNAs by controlling
poly(A) tail length in zebrafish. PLoS One 4, e7513.
Tam, O. H., Aravin, A. A., Stein, P., Girard, A., Murchison, E. P., Cheloufi, S.,
Hodges, E., Anger, M., Sachidanandam, R., Schultz, R. M. et al. (2008).
Pseudogene-derived small interfering RNAs regulate gene expression in mouse
oocytes. Nature 453, 534-538.
Tanaka, S., Kunath, T., Hadjantonakis, A. K., Nagy, A. and Rossant, J. (1998).
Promotion of trophoblast stem cell proliferation by FGF4. Science 282, 2072-
2075.
Tang, F., Kaneda, M., O’Carroll, D., Hajkova, P., Barton, S. C., Sun, Y. A.,
Lee, C., Tarakhovsky, A., Lao, K. and Surani, M. A. (2007). Maternal
microRNAs are essential for mouse zygotic development. Genes Dev. 21, 644-
648.
Thomson, T. and Lin, H. (2009). The biogenesis and function of PIWI proteins and
piRNAs: progress and prospect. Annu. Rev. Cell Dev. Biol. 25, 355-376.
Vagin, V. V., Sigova, A., Li, C., Seitz, H., Gvozdev, V. and Zamore, P. D. (2006).
A distinct small RNA pathway silences selfish genetic elements in the germline.
Science 313, 320-324.
Viswanathan, S. R., Daley, G. Q. and Gregory, R. I. (2008). Selective blockade
of microRNA processing by Lin28. Science 320, 97-100.
Viswanathan, S. R., Mermel, C. H., Lu, J., Lu, C. W., Golub, T. R. and Daley, G.
Q. (2009). microRNA expression during trophectoderm specification. PLoS One
4, e6143.
Wang, Y., Medvid, R., Melton, C., Jaenisch, R. and Blelloch, R. (2007). DGCR8
is essential for microRNA biogenesis and silencing of embryonic stem cell self-
renewal. Nat. Genet. 39, 380-385.
Wang, Y., Baskerville, S., Shenoy, A., Babiarz, J. E., Baehner, L. and Blelloch,
R. (2008). Embryonic stem cell-specific microRNAs regulate the G1-S transition
and promote rapid proliferation. Nat. Genet. 40, 1478-1483.
Watanabe, T., Takeda, A., Tsukiyama, T., Mise, K., Okuno, T., Sasaki, H.,
Minami, N. and Imai, H. (2006). Identification and characterization of two
novel classes of small RNAs in the mouse germline: retrotransposon-derived
siRNAs in oocytes and germline small RNAs in testes. Genes Dev. 20, 1732-
1743.
Watanabe, T., Totoki, Y., Toyoda, A., Kaneda, M., Kuramochi-Miyagawa, S.,
Obata, Y., Chiba, H., Kohara, Y., Kono, T., Nakano, T. et al. (2008).
Endogenous siRNAs from naturally formed dsRNAs regulate transcripts in mouse
oocytes. Nature 453, 539-543.
Wienholds, E., Koudijs, M. J., van Eeden, F. J., Cuppen, E. and Plasterk, R. H.
(2003). The microRNA-producing enzyme Dicer1 is essential for zebrafish
development. Nat. Genet. 35, 217-218.
Xu, J., Yang, J. Y., Niu, Q. W. and Chua, N. H. (2006). Arabidopsis DCP2, DCP1,
and VARICOSE form a decapping complex required for postembryonic
development. Plant Cell 18, 3386-3398.
Yamanaka, S. (2009). A fresh look at iPS cells. Cell 137, 13-17.
Yan, N., Lu, Y., Sun, H., Qiu, W., Tao, D., Liu, Y., Chen, H., Yang, Y., Zhang, S.,
Li, X. et al. (2009). Microarray profiling of microRNAs expressed in testis tissues
of developing primates. J. Assist. Reprod. Genet. 26, 179-186.
Yi, R., Pasolli, H. A., Landthaler, M., Hafner, M., Ojo, T., Sheridan, R., Sander,
C., O’Carroll, D., Stoffel, M., Tuschl, T. et al. (2009). DGCR8-dependent
microRNA biogenesis is essential for skin development. Proc. Natl. Acad. Sci.
USA 106, 498-502.
Yu, J., Vodyanik, M. A., Smuga-Otto, K., Antosiewicz-Bourget, J., Frane, J.
L., Tian, S., Nie, J., Jonsdottir, G. A., Ruotti, V., Stewart, R. et al. (2007).
Induced pluripotent stem cell lines derived from human somatic cells. Science
318, 1917-1920.
Yu, Z., Raabe, T. and Hecht, N. B. (2005). MicroRNA Mirn122a reduces
expression of the posttranscriptionally regulated germ cell transition protein 2
(Tnp2) messenger RNA (mRNA) by mRNA cleavage. Biol. Reprod. 73, 427-433.
Zeng, Y., Yi, R. and Cullen, B. R. (2003). MicroRNAs and small interfering RNAs
can inhibit mRNA expression by similar mechanisms. Proc. Natl. Acad. Sci. USA
100, 9779-9784.
DEVELOPMENT
... Recent advances regarding the stem cell-based therapies for SCI have provided enormous potential for biological and future *Address correspondence to these authors at the Department of Stem Cells and Developmental Biology, Cell Science Research Center, Royan Institute for Stem Cell Biology and Technology, ACECR, Tehran, Iran; E-mail: Maryamfarzaneh2013@yahoo.com Physiology Research Center, Department of Physiology, Ahvaz Jundishapur University of Medical Sciences, Ahvaz, Iran; Tel: +989171491729; Fax: +986133738248; E-mail: Esmaeil.khoshnam1392@gmail.com clinical applications [13,14]. Basically, stem cells are defined by their abilities to divide asymmetrically and ultimately differentiate into various cell lineages depending on their potency [15,16]. On the basis of regenerative applications, stem cells are classified into pluripotent stem cells (embryonic stem cells (ESC) [17,18], multipotent stem cells ((neural stem cells (NSC) [19,20] or mesenchymal stem cells (MSC)) [21,22], and unipotent stem cells [23][24][25]. ...
Article
Full-text available
Spinal cord injury (SCI) as a serious public health issue and neurological insult is one of the most causes of long-term disability. To date, a variety of techniques have been widely developed to treat central nervous system injury. Currently, clinical treatments are limited to surgical decompression and pharmacotherapy. Because of their negative effects and inefficiency, novel therapeutic approaches in the management of SCI are required. Improvement and innovation of stem cell-based therapies have allowed a huge potential for biological and future clinical applications. Human pluripotent stem cells (hPSCs) including embryonic stem cells (ESCs) and induced pluripotent stem cells (iPSCs) are defined by their abilities to divide asymmetrically, self-renew and ultimately differentiate into various cell lineages. There are considerable research efforts to use various types of stem cells such as ESCs, neural stem cells (NSCs), and mesenchymal stem cells (MSCs) in the treatment of patients with SCI. Moreover, the use of patient-specific iPSCs hold great potential as an unlimited cell source for generating in vivo models of SCI. In this review we focused on the potential of hPSCs in treating SCI
... The majority of miRNAs are transcribed by RNA polymerase II [71], whereas others are transcribed by RNA polymerase III [84]. RNA polymerase III can also transcribe viral miRNAs and some endogenous miRNA-like small RNAs derived from transfer RNAs (tRNAs) [85,86]. ...
Article
Full-text available
MicroRNAs (miRNAs) are a subset of endogenous, small, non-coding RNA molecules involved in the post-transcriptional regulation of eukaryotic gene expression. Dysregulation in miRNA-related pathways in the central nervous system (CNS) is associated with severe neuronal injury and cell death, which can lead to the development of neurodegenerative disorders, such as amyotrophic lateral sclerosis (ALS). ALS is a fatal adult onset disease characterized by the selective loss of upper and lower motor neurons. While the pathogenesis of ALS is still largely unknown, familial ALS forms linked to TAR DNA-binding protein 43 (TDP-43) and fused in sarcoma (FUS) gene mutations, as well as sporadic forms, display changes in several steps of RNA metabolism, including miRNA processing. Here, we review the current knowledge about miRNA metabolism and biological functions and their crucial role in ALS pathogenesis with an in-depth analysis on different pathways. A more precise understanding of miRNA involvement in ALS could be useful not only to elucidate their role in the disease etiopathogenesis but also to investigate their potential as disease biomarkers and novel therapeutic targets.
... Sequence analyzed (Okamura and Lai, 2008;Tam et al., 2008;Watanabe et al., 2008), which require Dicer, but not Drosha/Dgcr8 for processing (Babiarz et al., 2008(Babiarz et al., , 2009). Endo-siRNAs have been described in murine embryonic stem cells (mESCs), oocytes (Babiarz et al., 2008;Tam et al., 2008;Watanabe et al., 2008) and in murine embryonic skin cells (Yi et al., 2009). ...
Article
In 2010, we described many similar DNA sequences in human and viral genomes, including herpesviral ones. The data obtained allowed us to suggest that these motifs may provide the antiviral protection by mating with a complementary potential target and destroying it by the catalytic way like small interfering RNA, siRNA. Since we have analyzed these viruses as a group, two major issues seemed to us curious: (1) the number of such motifs in genomes of various herpesvirus types, and (2) distribution of these motifs in an individual viral genome.
Article
Full-text available
Simple Summary Hepatocellular carcinoma (HCC) is one of the most frequently occurring cancers, and the prognosis for late-stage HCC remains poor. A better understanding of the pathogenesis of HCC is expected to improve outcomes. MicroRNAs (miRNAs) are small, noncoding, single-stranded RNAs that regulate the expression of various target genes, including those in cancer-associated genomic regions or fragile sites in various human cancers. We summarize the central roles of miRNAs in the pathogenesis of HCC and discuss their potential utility as valuable biomarkers and new therapeutic agents for HCC. Abstract Hepatocellular carcinoma (HCC) is the seventh most frequent cancer and the fourth leading cause of cancer mortality worldwide. Despite substantial advances in therapeutic strategies, the prognosis of late-stage HCC remains dismal because of the high recurrence rate. A better understanding of the etiology of HCC is therefore necessary to improve outcomes. MicroRNAs (miRNAs) are small, endogenous, noncoding, single-stranded RNAs that modulate the expression of their target genes at the posttranscriptional and translational levels. Aberrant expression of miRNAs has frequently been detected in cancer-associated genomic regions or fragile sites in various human cancers and has been observed in both HCC cells and tissues. The precise patterns of aberrant miRNA expression differ depending on disease etiology, including various causes of hepatocarcinogenesis, such as viral hepatitis, alcoholic liver disease, or nonalcoholic steatohepatitis. However, little is known about the underlying mechanisms and the association of miRNAs with the pathogenesis of HCC of various etiologies. In the present review, we summarize the key mechanisms of miRNAs in the pathogenesis of HCC and emphasize their potential utility as valuable diagnostic and prognostic biomarkers, as well as innovative therapeutic targets, in HCC diagnosis and treatment.
Article
When embryonic stem cells (ESCs) differentiate, they must both silence the ESC self-renewal program and activate new tissue-specific programs. In the absence of DGCR8 (Dgcr8(-/-)), a protein required for microRNA (miRNA) biogenesis, mouse ESCs are unable to silence self-renewal. Here we show that the introduction of let-7 miRNAs-a family of miRNAs highly expressed in somatic cells-can suppress self-renewal in Dgcr8(-/-) but not wild-type ESCs. Introduction of ESC cell cycle regulating (ESCC) miRNAs into the Dgcr8(-/-) ESCs blocks the capacity of let-7 to suppress self-renewal. Profiling and bioinformatic analyses show that let-7 inhibits whereas ESCC miRNAs indirectly activate numerous self-renewal genes. Furthermore, inhibition of the let-7 family promotes de-differentiation of somatic cells to induced pluripotent stem cells. Together, these findings show how the ESCC and let-7 miRNAs act through common pathways to alternatively stabilize the self-renewing versus differentiated cell fates.
Article
Full-text available
In human prostate cancer, the microRNA biogenesis machinery increases with prostate cancer progression. Here, we show that deletion of the Dgcr8 gene, a critical component of this complex, inhibits tumor progression in a Pten-knockout mouse model of prostate cancer. Early stages of tumor development were unaffected, but progression to advanced prostatic intraepithelial neoplasia was severely inhibited. Dgcr8 loss blocked Pten null-induced expansion of the basal-like, but not luminal, cellular compartment. Furthermore, while late-stage Pten knockout tumors exhibit decreased senescence-associated beta-galactosidase activity and increased proliferation, the simultaneous deletion of Dgcr8 blocked these changes resulting in levels similar to wild type. Sequencing of small RNAs in isolated epithelial cells uncovered numerous miRNA changes associated with PTEN loss. Consistent with a Pten-Dgcr8 association, analysis of a large cohort of human prostate tumors shows a strong correlation between Akt activation and increased Dgcr8 mRNA levels. Together, these findings uncover a critical role for microRNAs in enhancing proliferation and enabling the expansion of the basal cell compartment associated with tumor progression following Pten loss. © 2015 The Authors.
Article
Full-text available
Understanding origin, evolution and functions of small RNA (sRNA) genes has been a great challenge in the past decade. Molecular mechanisms underlying sexual reversal in vertebrates, particularly sRNAs involved in this process, are largely unknown. By deep-sequencing of small RNA transcriptomes in combination with genomic analysis, we identified a large amount of piRNAs and miRNAs including over 1,000 novel miRNAs, which were differentially expressed during gonad reversal from ovary to testis via ovotesis. Biogenesis and expressions of miRNAs were dynamically changed during the reversal. Notably, phylogenetic analysis revealed dynamic expansions of miRNAs in vertebrates and an evolutionary trajectory of conserved miR-17-92 cluster in the Eukarya. We showed that the miR-17-92 cluster in vertebrates was generated through multiple duplications from ancestor miR-92 in invertebrates Tetranychus urticae and Daphnia pulex from the Chelicerata around 580 Mya. Moreover, we identified the sexual regulator Dmrt1 as a direct target of the members miR-19a and -19b in the cluster. These data suggested dynamic biogenesis and expressions of small RNAs during sex reversal and revealed multiple expansions and evolutionary trajectory of miRNAs from invertebrates to vertebrates, which implicate small RNAs in sexual reversal and provide new insight into evolutionary and molecular mechanisms underlying sexual reversal.
Article
microRNAs (miRNAs) are important modulators of development. Owing to their ability to simultaneously silence hundreds of target genes, they have key roles in large-scale transcriptomic changes that occur during cell fate transitions. In somatic stem and progenitor cells - such as those involved in myogenesis, haematopoiesis, skin and neural development - miRNA function is carefully regulated to promote and stabilize cell fate choice. miRNAs are integrated within networks that form both positive and negative feedback loops. Their function is regulated at multiple levels, including transcription, biogenesis, stability, availability and/or number of target sites, as well as their cooperation with other miRNAs and RNA-binding proteins. Together, these regulatory mechanisms result in a refined molecular response that enables proper cellular differentiation and function.
Article
Post-transcriptional gene regulation is one mechanism that occurs 'above the genome', allowing the cells of an organism to have dramatically different phenotypes and functions. Non-coding ribonucleic acid (ncRNA) molecules regulate transcript and protein abundance above the level of transcription, and appear to play substantial roles in regulation of reproductive tissues. Three primary classes of small ncRNA are microRNA (miRNA), endogenous small interfering RNA (endo-siRNA), and PIWI-interacting RNA (piRNA). These RNA classes have similarities and clear distinctions between their biogenesis and in the interacting protein machinery that facilitates their effects on cellular phenotype. Characterization of the expression and importance of the critical components for the biogenesis of each class in different tissues is continuously contributing a better understanding of each of these RNA classes in different reproductive cell types. Here, we discuss the expression and potential roles of miRNA, endo-siRNA, and piRNA in reproduction from germ-cell development to pregnancy establishment and placental function. Additionally, the potential contribution of RNA binding proteins, long ncRNAs, and the more recently discovered circular RNAs (circRNAs) in relation to small RNA function is discussed. Mol. Reprod. Dev. © 2013 Wiley Periodicals, Inc.
Article
Full-text available
Small RNAs bound to Argonaute proteins recognize partially or fully complementary nucleic acid targets in diverse gene-silencing processes. A subgroup of the Argonaute proteins-known as the 'Piwi family'-is required for germ- and stem-cell development in invertebrates, and two Piwi members-MILI and MIWI-are essential for spermatogenesis in mouse. Here we describe a new class of small RNAs that bind to MILI in mouse male germ cells, where they accumulate at the onset of meiosis. The sequences of the over 1,000 identified unique molecules share a strong preference for a 5' uridine, but otherwise cannot be readily classified into sequence families. Genomic mapping of these small RNAs reveals a limited number of clusters, suggesting that these RNAs are processed from long primary transcripts. The small RNAs are 26-31 nucleotides (nt) in length-clearly distinct from the 21-23 nt of microRNAs (miRNAs) or short interfering RNAs (siRNAs)-and we refer to them as 'Piwi-interacting RNAs' or piRNAs. Orthologous human chromosomal regions also give rise to small RNAs with the characteristics of piRNAs, but the cloned sequences are distinct. The identification of this new class of small RNAs provides an important starting point to determine the molecular function of Piwi proteins in mammalian spermatogenesis.
Article
Full-text available
MicroRNAs (miRNAs) are small noncoding RNAs that posttranscriptionally regulate gene expression. Hundreds of miRNAs are expressed in mammals; however, their functions are just starting to be uncovered. MicroRNAs are processed from a long hairpin mRNA transcript, down to a approximately 23-nucleotide duplex. The enzyme Dicer1 is required for miRNA processing, and mouse knockouts of Dicer1 are embryonic lethal before 7.5 days postcoitus. To examine the function of miRNAs specifically in the germline, we used a mouse model that expresses Cre recombinase from the TNAP locus and a floxed Dicer1 conditional allele. Removal of Dicer1 from germ cells resulted in male infertility. Germ cells were present in adult testes, but few tubules contained elongating spermatids. Germ cells that did differentiate to elongating spermatids exhibited abnormal morphology and motility. Rarely, sperm lacking Dicer1 could fertilize wild-type eggs to generate viable offspring. These results show that Dicer1 and miRNAs are essential for proper differentiation of the male germline.
Article
Full-text available
P bodies are cytoplasmic domains that contain proteins involved in diverse posttranscriptional processes, such as mRNA degradation, nonsense-mediated mRNA decay (NMD), translational repression, and RNA-mediated gene silencing. The localization of these proteins and their targets in P bodies raises the question of whether their spatial concentration in discrete cytoplasmic domains is required for posttranscriptional gene regulation. We show that processes such as mRNA decay, NMD, and RNA-mediated gene silencing are functional in cells lacking detectable microscopic P bodies. Although P bodies are not required for silencing, blocking small interfering RNA or microRNA silencing pathways at any step prevents P-body formation, indicating that P bodies arise as a consequence of silencing. Consistently, we show that releasing mRNAs from polysomes is insufficient to trigger P-body assembly: polysome-free mRNAs must enter silencing and/or decapping pathways to nucleate P bodies. Thus, even though P-body components play crucial roles in mRNA silencing and decay, aggregation into P bodies is not required for function but is instead a consequence of their activity.
Article
Full-text available
Human germ cell tumors show a strong sensitivity to genetic background similar to Dnd1(Ter/Ter) mutant mice, where testicular teratomas arise only on the 129/SvJ genetic background. The introduction of the Bax mutation onto mixed background Dnd1(Ter/Ter) mutants, where teratomas do not typically develop, resulted in a high incidence of teratomas. However, when Dnd1(Ter/Ter); Bax(-/-) double mutants were backcrossed to C57BL/6J, no tumors arose. Dnd1(Ter/Ter) germ cells show a strong downregulation of male differentiation genes including Nanos2. In susceptible strains, where teratomas initiate around E15.5-E17.5, many mutant germ cells fail to enter mitotic arrest in G0 and do not downregulate the pluripotency markers NANOG, SOX2 and OCT4. We show that DND1 directly binds a group of transcripts that encode negative regulators of the cell cycle, including p27(Kip1) and p21(Cip)(1). P27(Kip1) and P21(Cip1) protein are both significantly decreased in Dnd1(Ter/Ter) germ cells on all strain backgrounds tested, strongly suggesting that DND1 regulates mitotic arrest in male germ cells through translational regulation of cell cycle genes. Nonetheless, in C57BL/6J mutants, germ cells arrest prior to M-phase of the cell cycle and downregulate NANOG, SOX2 and OCT4. Consistent with their ability to rescue cell cycle arrest, C57BL/6J germ cells overexpress negative regulators of the cell cycle relative to 129/SvJ. This work suggests that reprogramming of pluripotency in germ cells and prevention of tumor formation requires cell cycle arrest, and that differences in the balance of cell cycle regulators between 129/SvJ and C57BL/6 might underlie differences in tumor susceptibility.
Article
Full-text available
MicroRNAs (miRNAs) are a large family of post-transcriptional regulators of gene expression that are approximately 21 nucleotides in length and control many developmental and cellular processes in eukaryotic organisms. Research during the past decade has identified major factors participating in miRNA biogenesis and has established basic principles of miRNA function. More recently, it has become apparent that miRNA regulators themselves are subject to sophisticated control. Many reports over the past few years have reported the regulation of miRNA metabolism and function by a range of mechanisms involving numerous protein-protein and protein-RNA interactions. Such regulation has an important role in the context-specific functions of miRNAs.
Article
Cis-natural antisense transcripts (cis-NATs) have been speculated to be substrates for endogenous RNA interference (RNAi), but little experimental evidence for such a pathway in animals has been reported. Analysis of massive Drosophila melanogaster small RNA data sets now reveals two mechanisms that yield endogenous small interfering RNAs (siRNAs) via bidirectional transcription. First, >100 cis-NATs with overlapping 3' exons generate 21-nt, and, based on previously published small RNA data [corrected] Dicer-2 (Dcr-2)-dependent, 3'-end modified siRNAs. The processing of cis-NATs by RNA interference (RNAi) seems to be actively restricted, and the selected loci are enriched for nucleic acid-based functions and include Argonaute-2 (AGO2) itself. Second, we report that extended intervals of the thickveins and klarsicht genes generate exceptionally abundant siRNAs from both strands. These siRNA clusters derive from atypical cis-NAT arrangements involving introns and 5' or internal exons, but their biogenesis is similarly Dcr-2- and AGO2-dependent. These newly recognized siRNA pathways broaden the scope of regulatory networks mediated by small RNAs.
Article
The evolutionarily conserved Argonaute/PIWI (AGO/PIWI, also known as PAZ-PIWI domain or PPD) family of proteins is crucial for the biogenesis and function of small noncoding RNAs (ncRNAs). This family can be divided into AGO and PIWI subfamilies. The AGO proteins are ubiquitously present in diverse tissues. They bind to small interfering RNAs (siRNAs) and microRNAs (miRNAs). In contrast, the PIWI proteins are predominantly present in the germline and associate with a novel class of small RNAs known as PIWI-interacting RNAs (piRNAs). Tens of thousands of piRNA species, typically 24-32 nucleotide (nt) long, have been found in mammals, zebrafish, and Drosophila. Most piRNAs appear to be generated from a small number of long single-stranded RNA precursors that are often encoded by repetitive intergenic sequences in the genome. PIWI proteins play crucial roles during germline development and gametogenesis of many metazoan species, from germline determination and germline stem cell (GSC) maintenance to meios
Article
Small RNAs associate with Argonaute proteins and serve as sequence-specific guides for regulation of mRNA stability, productive translation, chromatin organization, and genome structure. In animals, the Argonaute superfamily segregates into two clades. The Argonaute clade acts in RNAi and in microRNA-mediated gene regulation in partnership with 21-22 nt RNAs. The Piwi clade, and their 26-30 nt piRNA partners, have yet to be assigned definitive functions. In mice, two Piwi-family members have been demonstrated to have essential roles in spermatogenesis. Here, we examine the effects of disrupting the gene encoding the third family member, MIWI2. Miwi2-deficient mice display a meiotic-progression defect in early prophase of meiosis I and a marked and progressive loss of germ cells with age. These phenotypes may be linked to an inappropriate activation of transposable elements detected in Miwi2 mutants. Our observations suggest a conserved function for Piwi-clade proteins in the control of transposons in the germline.
Article
The two first cell fate decisions taken in the mammalian embryo generate three distinct cell lineages: one embryonic, the epiblast, and two extraembryonic, the trophoblast and primitive endoderm. miRNAs are essential for early development, but it is not known if they are utilized in the same way in these three lineages. We find that in the pluripotent epiblast they inhibit apoptosis by blocking the expression of the proapoptotic protein Bcl2l11 (Bim) but play little role in the initiation of gastrulation. In contrast, in the trophectoderm, miRNAs maintain the trophoblast stem cell compartment by directly inhibiting expression of Cdkn1a (p21) and Cdkn1c (p57), and in the primitive endoderm, they prevent differentiation by maintaining ERK1/2 phosphorylation through blocking the expression of Mapk inhibitors. Therefore, we show that there are fundamental differences in how stem cells maintain their developmental potential in embryonic and extraembryonic tissues through miRNAs.