ArticlePDF AvailableLiterature Review

A germ‐cell odyssey: fate, survival, migration, stem cells and differentiation: Meeting on Germ Cells

Authors:
EMBO reports VOL 4 | NO 4 | 2003 ©2003 EUROPEAN MOLECULAR BIOLOGY ORGANIZATION
meeting report
meeting report
352
A germ-cell odyssey: fate, survival, migration,
stem cells and differentiation
Meeting on Germ Cells
E. Jane Albert Hubbard1+ & Renee A. Reijo Pera2*
1Department of Biology, New York University,New York, USA, and 2Department of Obstetrics, Gynecology and Reproductive
Sciences, Department of Physiology and Department of Urology, Program in Human Genetics and Program in Developmental and
Stem Cell Biology,University of California at San Francisco,San Francisco, California,USA
Introduction
The function of the germ line in all sexually reproducing organisms is
to produce gametes that are able to contribute to the next generation
(Fig. 1). Thus, like many a great epic film, the germ-cell saga spans
generations and addresses such weighty themes as fate, immortality,
death and transformation. During the recent meeting (9–13 October
2002) at Cold Spring Harbor Laboratories, germ cells did not disap-
point. Well-known molecular ‘stars’ took their roles to new heights,
and new ‘actors’ were introduced. Meeting participants gained an
unprecedented overview of comparative germline development.
Here, we focus on the comparison of some of the mechanistic and
molecular characteristics of phylogenetic groups that were presented
at this meeting, in the context of key steps in germline development.
Germline specification: ‘getting into character’
The germ-cell story begins and ends with ‘maternal pronucleus meets
paternal pronucleus’. Soon after this encounter, the decision is made
as to which cells assume a somatic fate and which potentially con-
tribute to the next generation. So critical is this decision that, across
diverse phyla, cells that are destined to take on germ-cell fates are
physically separated from potential somatic cells early in embryo-
genesis, presumably to protect them from influences that would limit
their potential or direct them along the path to a somatic fate.
One of the earliest recognized common themes in germline
development among non-mammalian species is the presence of
germline-associated cytoplasm (‘germ plasm’), which contains
microscopically distinct, electron-dense granules. In some well-stud-
ied systems, such as Drosophila melanogaster, germ plasm in the
embryo (known as ‘pole plasm’ in this species) can confer a germ-
cell fate, whereas in other species, this ability does not exist or has
not been shown. Even in species in which germ plasm or germ-
associated granule localization has not been linked directly to the
acquisition of a germ-cell fate, these granules segregate with the
germ line and remain associated with it at later stages (Saffman &
Lasko, 1999). Thus, studies that address germ-granule components
and their assembly are of great interest, especially as some of these
molecules are widely conserved across animal species.
1Department of Biology, New York University, 1009 Silver Center, 100 Washington Square
East, New York, New York 10003-6688, USA
2Department of Obstetrics, Gynecology and Reproductive Sciences, University of
California at San Francisco, 513 Parnassus Avenue, HSE1659, San Francisco, California
94143-0556, USA
+Corresponding author.Tel: +1 212 998 8293; Fax: +1 212 995 4015;
E-mail: jane.hubbard@nyu.edu
*Corresponding author.Tel: +1 415 476 3178; Fax: +1 415 476 3121;
E-mail: reijo@itsa.ucsf.edu
Submitted 22 January 2003; accepted 21 February 2003
Published online 21 March 2003
The biennial meeting on Germ Cells at Cold Spring Harbor Laboratories
took place during 9–13 October 2002. It was organized by Ruth Lehmann
and David Page. The photograph is taken from the cover of the abstracts
book and shows a bag of marbles mutant ovary that has germ cells
labelled in green (anti-Vasa antibody) and somatic-cell membranes in
orange (anti-spectrin antibody). Spectrin labelling also indicates the
circular spectrosome in each undifferentiated germ cell. Photograph by
Lilach Gilboa, from the laboratory of Ruth Lehmann.
EMBO reports 4, 352–357 (2003)
doi:10.1038/sj.embor.embor807
reviews
©2003 EUROPEAN MOLECULAR BIOLOGY ORGANIZATION EMBO reports VOL 4 | NO 4 | 2003
meeting report
353
germline factors is unclear. Several possibilities were debated infor-
mally at the meeting, including that: rapid evolution of reproductive
genes may be required for, or result from, speciation; group-specific
factors might have evolved under group-specific developmental
constraints; and constraints that shape the roles of various factors
may be derived from the non-germline roles that these molecules
have or once had during the course of evolution of a particular
species. Alternatively, given some of the anatomical and mechanis-
tic differences in germline development among the most studied
groups, it may be equally surprising that any conservation exists
across such large phylogenetic distances. Perhaps the conserved
factors form a core machinery, the functions of which have been
either maintained through evolution or modified for unique uses in
each species. Regardless, functional studies of both group-specific
genes and widely conserved genes are enriching the field of germ-
cell research.
The set of important non-conserved germline components
includes Xenopus Xpat, fly oskar, worm pie-1 and mammalian
Oct4. Xpat was originally identified as a transcript that localizes
to the vegetal pole, and it encodes a protein that localizes to the
germ plasm during the formation of stage 1 oocytes and during
the period of PGC movement into the mesentery (Hudson &
Woodland, 1998). Recent work discussed by H. Woodland
(Coventry, UK) suggests that the Xpat protein has a role in target-
ing and/or transporting granules to germinal particles. In addition,
Xpat is associated with centrosomes, although its precise function
there is unknown.
G. Seydoux (Baltimore, MD, USA) discussed two general mech-
anisms for the asymmetrical distribution of germ granules: asym-
metric enrichment before cell division, and degradation in nascent
somatic blastomeres. PIE-1 normally segregates with germ-cell
precursors and maintains germline fate. Video recordings of a
C. elegans PIE-1::GFP (green fluorescent protein) fusion protein
demonstrated its degradation in somatic blastomeres. Somatic
PIE-1 degradation depends on the first of its two CCCH-type zinc
fingers (ZF1) (Reese et al., 2000) and on zinc-finger-interacting
factor-1 (ZIF-1), a novel protein to which ZF1 binds. A yeast two-
hybrid screen with ZIF-1 identified a potential E3 ubiquitin ligase
subunit, Elongin C. Thus, ZIF-1 may provide a link between PIE-1
At this meeting, several possible germ-granule components were
introduced on the basis of their relationships with established com-
ponents such as the VASA proteins.VASA proteins are germline
DEAD-box RNA helicases that are found in primordial germ cells
(PGCs) in species ranging from hydra to humans. P. Lasko (Montreal,
Canada) provided a new piece of the Drosophila pole-plasm
assembly puzzle in the form of Gustavus (Gus), which is a protein
that is required for the localization of Vasa to the posterior of the
developing oocyte. Lasko pointed out that there are homologues of
Gus in evolutionarily distant organisms, including mammals (Styhler
et al., 2002). On a similar theme, K. Bennett (Columbia, MO, USA)
presented members of three conserved protein families, CSN-5,
KGB-1 and ZYX-1, that interact with the Vasa-related Caenorhabditis
elegans GLH proteins (Smith et al., 2002).
Nanos (Nos) and its related proteins have key roles in germ-cell
fate acquisition, migration and survival in various species. Because
Drosophila germ cells that lack nos die, it has been impossible to
assess their fate. S. Kobayashi (Okazaki, Japan) circumvented this
problem by using a deletion that removes three loci that are crucial
for embryonic programmed cell death. Surprisingly, a small but
significant percentage of nosgerm cells incorporated into somatic
tissues, and intermingled with and took on the appearance of
their neighbouring somatic cells. In many systems, the physical
mechanisms by which germinal particles are localized are unclear.
M.L. King (Miami, FL, USA) proposed that the localization of the
Nos-related Xenopus laevis protein Xcat2 into germinal particles
occurs by a ‘diffusion and entrapment’ mechanism. A conserved
zebrafish RNA-binding protein, Deadend, was introduced by J.
Stebler (Göttingen, Germany). This protein colocalizes with Vasa and
Nos in germ granules, and functional studies have suggested that it is
crucial for the migration of PGCs and for the maintenance of their
fate. Finally, three mouse homologues of Nos were revealed by M.
Tsuda and colleagues (Mishima, Japan); two of these proteins are
expressed and function in the germ line during embryogenesis.
In contrast to non-mammalian species, the determination of PGC
fate in mammals is independent of germline-specific granules; it
occurs instead through an inductive process. Because the precise
nature of germ granules is unknown, the question remains as to
whether they exist in mammals. Given that the factors that are associ-
ated with germ granules in non-mammalian species are also
expressed in mammalian germ cells, it is tempting to speculate that
all species share a basic germ-plasm machinery, and that this
machinery exists in mammals as submicroscopic complexes that are
rich in RNAs and RNA-binding proteins. One possibility is that this
machinery could be assembled in response to inductive signalling by
molecules such as bone morphogenetic protein 4 (BMP4; see below)
and could function in a manner analogous to that of germ granules in
model organisms, ensuring the formation and/or subsequent mainte-
nance of germ-cell populations.
Non-conserved genes in early development
Unlike the conserved Vasa and Nos families, several important
proteins involved in early germline development (in species such as
worms, flies, frogs, fish and mammals) lack obvious orthologues in
the genomes of species outside their phylogenetic group. In addi-
tion, some conserved proteins have different roles in different phyla.
Given the striking parallels in germline development among various
animals, one might expect conservation of protein families to be the
rule. Thus, the reason for the existence of crucial group-specific
Fig. 1 | Summary of the main events in germline development in animals. The
approximate temporal relationships between these events and the approximate
stage of their occurrence within the context of overall development are shown.
Details of germline development differ among species and between sexes
within species and may not be consistent with this generalized diagram.
Developmental stage
Embryonic Post-embryonic Adult
Germline specification Germ-cell differentiation Gametogenesis
Germ-cell migration
Germline sex
determination
Somatic gonad
development
Germline–soma
interactions
Meiosis
Stem-cell renewal
reviews
EMBO reports VOL 4 | NO 4 | 2003 ©2003 EUROPEAN MOLECULAR BIOLOGY ORGANIZATION
meeting report
354
To differentiate or not to differentiate
Once germ-cell fate is specified, germ cells must find, interact with
and populate the developing somatic gonad (Starz-Gaiano &
Lehmann, 2001). In organisms that produce germline stem cells,
either all or a subset of the germ cells that enter the somatic gonad
become stem cells. In the latter case, what determines which cells
acquire (or maintain) stem-cell potential? The results from several lab-
oratories suggest that this decision is regulated both spatially and
temporally. M. Asaoka, from the laboratory of H. Lin (Durham, NC,
USA), presented a lineage analysis in Drosophila using GFP-marked
pole cells to distinguish between two possible mechanisms of
stem-cell specification from the PGC pool: lineage dependent versus
position dependent. Position in the embryonic gonad appeared to be
critical. D. Godt (Toronto, Canada) discussed the role of Traffic jam, a
transcription factor from the musculo-aponeurotic fibrosarcoma
(Maf) family. In traffic jam mutants of Drosophila, the cells of the
somatic gonad and the germ line do not intermingle correctly, and
the premature separation of germ cells and somatic cells blocks
the proper differentiation of both cell types. Data presented by
E.J. Hubbard (New York, NY, USA) suggested that the timing and posi-
tion of the developmentally earliest meiotic entry in the C. elegans
hermaphrodite is determined by an interaction between the germ line
and the early-larval proximal somatic gonad, in addition to the well-
characterized interaction with the distal somatic gonad. An investiga-
tion of the PTEN lipid phosphatase in the mouse male germ line by
T. Nakano (Osaka, Japan) suggests that this protein is crucial for the
regulation of PGC proliferation. Male germ cells that lack PTEN fail to
arrest properly at the G1 stage on days E13.5–E14.5, resulting in an
increased number of bilateral testicular teratomas. It was proposed
that, in the absence of PTEN, PGCs may de-differentiate into pluripo-
tent ES or embryonic-germ-like cells, suggesting that PTEN is a
crucial regulator of germ-cell differentiation (Kimura et al., 2003).
Once germline stem cells are established, their self-renewal and
differentiation must be regulated. If differentiation exceeds self-
renewal, the germ line is depleted, and if self-renewal exceeds
differentiation, overproliferation may occur. Regulation of germline
proliferation involves input from neighbouring somatic cells,
although different signalling pathways are used for analogous pur-
poses in different organisms. Often, there is an anatomical axis of
proliferation and differentiation, which has stem cells at one end and
mature gametes at the other (open) end (Fig. 2). For example, in both
sexes of Drosophila, in C. elegans and in male mammals, localized
somatic cells signal to the germ line, and germline-localized factors
respond. The molecular details of these signalling pathways are now
under investigation in several systems.
In Drosophila males, the Jak/Stat pathway has been implicated in
the maintenance of germline stem-cell fate through a ‘hub-to-germ-
line’ signal (Kiger et al., 2001; Tulina & Matunis, 2001). Work pre-
sented by E. Matunis (Baltimore, MD, USA) suggests that the Jak/Stat
signalling pathway may also act in the hub itself, where it is activated
in an autocrine manner by the ligand Unpaired, to regulate the size of
the hub and the number of germ cells that it contacts. M. Fuller
(Stanford, CA, USA) presented a cell-biological approach, suggesting
that the orientation of stem-cell divisions is crucial for determining
the fate of the two daughter cells. Mutations in the gene that encodes
centrosomin, an integral centrosome component, result in spindle-
orientation defects that correlate with an increased number of
germline stem cells. In Drosophila females, a BMP homologue,
the product of the decapentaplegic gene, functions as the main
and a ubiquitin-mediated mechanism that ensures the degradation
of germ-specific components in somatic cells during early develop-
ment in C. elegans. Seydoux also made a comparison between
Drosophila and C. elegans, pointing out the role of a conserved
kinase, PAR-1, in relation to the non-conserved germ-plasm com-
ponents Oskar and PIE-1, respectively (Pellerrieri & Seydoux,
2002). In both systems, PAR-1 is required to stabilize the germ-
plasm proteins in the posterior of the embryo. In Drosophila, PAR-1
seems to stabilize Oskar by direct phosphorylation (Riechmann
et al., 2002). In C. elegans, PAR-1 seems to stabilize PIE-1 through
two intermediates, MEX-5 and MEX-6, that must be downregulated
in the posterior of the embryo for stabilization to occur (Cuenca
et al., 2003). This comparison illustrates the complex interplay
between conserved and non-conserved factors in analogous
processes.
The mouse protein Oct4 contains a DNA-binding POU domain
and is expressed in the totipotent embryo, the germ line and undiffer-
entiated embryonic stem (ES) cells. A. Tomilin (Freiburg, Germany)
reported on a conditional knockout of the Oct4 gene in PGCs. In
Oct4PGCs, proliferation was halted between days E9.5 and E10, and
the germ-cell population was virtually eliminated by apoptosis. The
loss of Oct4 in ES cells causes a different outcome—loss of pleuripo-
tency, not cell death as in PGCs—so Tomilin suggested that Oct4 has
different targets during development.
As noted by M.A. Surani (Cambridge, UK), because there is no
evidence for the presence of germline determinants in mammals,
and only a subset of BMP4-responding cells go on to acquire a
germ-cell fate, other factors must be required for germline deter-
mination. Surani and colleagues therefore compared gene
expression between single PGC founders and their somatic
neighbours that share a common ancestry, and identified fragilis,
a gene that encodes an interferon-inducible transmembrane pro-
tein. Its expression is dependent on the dose of BMP4 and the
protein may have a role in establishing germ-cell competence.
They also identified stella, a gene that is first expressed in the
nascent PGCs, starting at the centre of the fragilis-positive region,
and thereafter in migrating PGCs. A key finding is that, unlike
somatic cells, nascent PGCs repress region-specific Hox genes
such as Hoxb1. Thus, the expression of stella and Hoxb1 is mutu-
ally exclusive in germline and somatic neighbours (Saitou et al.,
2002). Y. Matsui (Osaka, Japan) and colleagues used a primary
culture system of mouse epiblast cells, from which PGCs arise, to
show that inductive signals from the extra-embryonic ectoderm
are required between days E5.25 and E5.5; inductive signalling
with BMP4 alone stimulated PGC formation at day E5.5 but not
before that stage. The investigators also identified mil-1 and mil-2
by virtue of their differential expression between early and
migrating PGCs. Both genes encode members of a family of
interferon-induced transmembrane proteins, one of which is
identical to fragilis (Tanaka & Matsui, 2003).
The analysis of early events in mammalian germline develop-
ment is enhanced by new germ-cell markers such as the Oct4::GFP
fusion presented by C. Wylie (Cincinnati, OH, USA; Anderson et al.,
2000; Molyneaux et al., 2001). The use of this marker was shown in
the analysis of mouse germ cells that lack the pro-apoptotic gene
Bax. Similar to the ectopic localization of PGCs that lack nos and
other cell-death genes in Drosophila, mammalian germ cells that
survived early fetal development were found in ectopic, somatic
locations, but still expressed germ-cell markers.
reviews
©2003 EUROPEAN MOLECULAR BIOLOGY ORGANIZATION EMBO reports VOL 4 | NO 4 | 2003
meeting report
355
soma-to-germ-line signal that maintains stem cells (Spradling et al.,
2001). The bag of marbles (bam) gene is required for the differenti-
ated germ-cell fate and antagonizes stem-cell fate. D. McKearin
(Dallas, TX, USA) presented evidence that a transcriptional silencer
element is active specifically in stem cells and prevents bam expres-
sion, blocking differentiation and allowing the maintenance of stem
cells (Chen & McKearin, 2003). On the basis of an analysis of mutants
in the zero population growth gene, which encodes a gap-junction
protein (Tazuke et al., 2002), and of observations from wild-type flies,
L. Gilboa (New York, NY, USA) suggested the existence of an interme-
diate (pre-cystoblast) population of cells. She further suggested that
pre-cystoblasts are probably an intermediate in the differentiation
process rather than a ‘transit-amplifying’ cell population, and that
germ-cell tumours observed in bam mutants (or under similar condi-
tions) may primarily contain pre-cystoblasts rather than stem cells.
In C. elegans, GLP-1, a Notch-family receptor, is activated in the
germ line in response to a somatically produced ligand to promote
mitosis and/or inhibit meiosis. The RNA-binding protein GLD-1 acts
genetically downstream of, and in opposition to, the activity of the
GLP-1-mediated pathway. GLD-1 and GLD-2, an atypical poly(A)
polymerase, are redundant in this meiosis-promoting role. In some
way, these—and probably other—components demarcate the
mitosis–meiosis boundary at the proximal edge of the stem-cell
population. Although the picture is incomplete, a complex web of
regulation that involves several RNA-based controls is emerging. The
results from a genetic strategy presented by T. Schedl (St Louis, MO,
USA) suggest that the nos-related gene nos-3 is redundant with gld-2
in promoting meiosis. Analysis of GLD-1 protein levels revealed that
part of the spatial mechanism that determines the position of the
mitosis–meiosis border involves inhibition of GLD-1 levels by glp-1
and elevation of GLD-1 by gld-2/nos-3. J. Kimble (Madison, WI, USA)
presented an analysis of GLD-2 and its activating protein GLD-3,
which is another RNA-binding molecule. These studies led to the
proposal that meiotic RNAs are inactive and that poly(A) addition
by GLD-2/GLD-3 enables translation. The plot is thickened further
by the redundant activity of the PUF (PUMILIO and FBF) family
members FBF-1 and FBF-2, which maintain stem-cell proliferation in
late-larval and adult stages and which interact physically with GLD-3
and NOS-3 (Crittenden et al., 2002; Eckmann et al., 2002; Kraemer
et al., 1999).
Other reports reflected the importance of regulating germline
development through the control of RNA in many phyla. R. Reijo
Pera (San Francisco, CA, USA) brought the PUF family into the
mammalian germline picture with her report on the deleted in
azoospermia (DAZ) gene family and interacting factors (Moore et
al., 2003). Previous work had shown that DAZ genes are required
for germ-cell development in diverse organisms, ranging from
worms to humans. In recent studies, PUM2, a human PUF family
protein, was identified as one of six DAZ-interacting proteins.
PUM2, and several other interacting proteins, are expressed specifi-
cally in ES cells, PGCs and mature germ cells in both sexes. All six
proteins contain RNA-binding motifs. J. Richter (Worcester, MA,
USA) described findings from mice that were homozygous for a
null mutation of CPEB (cytoplasmic-polyadenylation-element-
binding protein). Although ovaries initially form and contain
numerous immature germ cells, as these cells enter meiosis, the
oocyte chromatin becomes diffuse and messenger RNAs that
encode synaptonemal-complex proteins are not correctly
polyadenylated or translated (Tay & Richter, 2001). Studies present-
ed by M. Hengartner (Zurich, Switzerland) revealed a role for a
CPEB-related protein in cell death in the C. elegans hermaphrodite
germ line. Thus, CPEB family members are continuing to be found
throughout the animal kingdom in various germline roles (Mendez
& Richter, 2001).
Fig. 2 | A schematic representation of gonad anatomy in (left to right) the Drosophila male, Drosophila female, Caenorhabditis elegans hermaphrodite and
male mammal. This diagram emphasizes the relative positions of stem cells (green) and differentiated germ cells. The green-to-yellow arrow indicates the
direction of proliferation to differentiation. In each case, the direction of the arrow corresponds to an anatomical axis that can be described as ‘apical-tip to
basal-lumen’ and the ‘anterior-to-posterior’ axes in the Drosophila male and female, respectively,‘distal-to-proximal’ in C. elegans, and a radial axis of‘basal-
lamina-to-lumen’in male mammals. Somatic cells that are important for stem-cell maintenance are shown in red.
Proliferation
Differentiation
Drosophila Drosophila C. elegans
Sperm Oocyte Sperm
or
oocyte
Sperm
Hub Terminal filament
and cap cells
Distal tip cell Sertoli cell
Lumen
Mammalian
reviews
EMBO reports VOL 4 | NO 4 | 2003 ©2003 EUROPEAN MOLECULAR BIOLOGY ORGANIZATION
meeting report
356
Transcriptional control, silencing and epigenetics
Germ cells are the vehicles through which DNA is passed from one
generation to the next. What molecular mechanisms underlie the
ability of germ cells to resist somatic fates, and yet undergo a
dramatic differentiation to gametes, while maintaining their ability
to contribute to the formation of the totipotent zygote?
Transcriptional and epigenetic controls that govern maintenance
and reprogramming of the germ line in diverse organisms are of
increasing interest as they may hold the key to some of these ques-
tions (Sassone-Corsi, 2002; Surani, 2001; Pirrotta, 2002). In
Drosophila and C. elegans, transcriptional silence is a hallmark of
nascent germ cells. Evidence that transcriptional silence may be a
characteristic of vertebrates was presented by M.L. King. The
expression of zygotic genes in nascent Xenopus PGCs was assessed
by RNA PolII CTD phosphorylation, and was shown to start at the
neurula stage. These data suggest that the failure of PGCs to adopt
endodermal fates, despite the fact that they inherit endodermal
determinants such as VegT, may be due to transcriptional repression
in nascent PGCs at a time when VegT targets would normally be
activated. Transcriptional silence in the early Drosophila germ line
was discussed by T. Jongens (Philadelphia, PA, USA) in the context
of the gene germcell-less (gcl).The removal of gcl activity results in
both a failure to establish transcriptional silencing in the germ line-
destined nuclei and a reduction in the number of pole cells formed,
whereas ectopic GCL expression causes an ectopic decrease in
transcriptional activity. Thus, transcriptional quiescence seems to
be imposed before pole-cell formation (Leatherman et al., 2002).
GCL-interacting factors include PHO-like, a protein with a zinc-
finger domain that is similar to PHO, a member of the Polycomb
group (see next paragraph). Jongens presented a model in which
GCL may tether chromatin to the nuclear envelope, thereby silenc-
ing transcription in a manner similar to telomeric silencing.
R. Martinho (New York, NY, USA) reported that tailless (tll) and
zerknullt (zen) in Drosophila, two genes that are normally transcrip-
tionally repressed before and during gastrulation, are not repressed
in the absence of the non-coding polar granule component RNA in
early embryonic germ cells. Furthermore, Osa, a component of the
Swi/Snf chromatin-remodelling protein complex, is required for
repression of zen transcription in germ cells during gastrulation,
suggesting that several independent but sequential mechanisms act
to repress transcription in early germ cells.
In C. elegans, at least two systems of germline silencing are at
work: the PIE-1 system, which represses transcription in early
germline blastomeres until the ~100-cell stage, and the MES system,
which is named for the maternal effect sterile (mes) mutants that led
to its discovery (Seydoux & Schedl, 2001; Pirrotta, 2002). Several mes
genes share homology with genes of the fly Polycomb group, which
are known for their role in the repression of homeotic genes. Several
laboratories (Fong et al., 2002; Kelly et al., 2002) have shown
germline repression of chromatin on the X chromosome and its corre-
lation with certain histone modifications. MES proteins participate in
this silencing. L. Bender (Bloomington, IN, USA) reported on
SET-2, one of several SET-domain proteins that are involved in
MES-mediated silencing, suggesting that this protein marks active
chromatin. C. Bean (Atlanta, GA, USA) elaborated on recent findings,
suggesting that the C. elegans paternal X chromosome is preferential-
ly inactivated in early XX embryos. A model was suggested in which
an epigenetic imprint is established in the male germ line, and its
decay in the early embryo regulates the onset of somatic dosage
compensation. Remarkably, mammalian dosage compensation can
involve non-random X-chromosome inactivation limited to the pater-
nally inherited X chromosome (in marsupials and the extra-embryonic
tissues of eutherian mammals). Because both processes may involve
proteins that are related to Esc (extra sexcombs) in Drosophila, MES-6
in worms and Eed (embryonic ectoderm development) in mouse, it is
tempting to speculate that an ancient large-scale silencing system may
have been co-opted by diverse phylogenetic groups to establish
and/or maintain active and inactive chromatin states over large con-
tiguous stretches of the genome (Pirrotta, 2002).
Although global silencing in germ cells has not been demon-
strated in mammals, epigenetic reprogramming is known to occur in
each generation. A better understanding of these processes is
becoming especially important as issues are raised about problems
in animal cloning and in therapeutic cloning for humans. In a com-
pelling talk by R. Jaenisch (Cambridge, MA, USA), these processes
were discussed with respect to cloning by somatic-cell nuclear
transfer. Jaenisch and colleagues addressed the questions: Does the
tissue source of the nucleus matter in nuclear transfer? And are
common problems in cloned embryos, such as large size and
defects in organogenesis, of genetic or epigenetic origin? Using
nuclear transfer in mice, they found that ES-cell-derived clones had
better survival rates than clones derived from mature somatic cells.
Furthermore, the group showed that problems with nuclear transfer
are likely to be epigenetic rather than genetic in origin. Strikingly,
4–5% of all genes, but 30–50% of imprinted genes, were aberrantly
expressed in cloned embryos. Jaenisch concluded that the lengthy
processes of normal gametogenesis and fertilization may give the
zygotic nucleus the ability to direct normal development, whereas
somatic-cell nuclear transfer may not provide the same opportunity
(Hochedlinger & Jaenisch, 2002; Humpherys et al., 2002).
And the ‘Oskar’ goes to…
Although they cannot be covered here, several other themes
emerged at this meeting, including molecular and evolutionary
aspects of meiosis, gametogenesis, germline sex determination,
sex-chromosome evolution, sperm–female (both sperm–egg and
sperm–soma) interactions, differential gamete success, the
evolution of reproductive proteins, the trade-off between male
reproductive success and female health in Drosophila, and the
astonishing (if not alarming) effects of the bacterium Wolbachia on
reproduction and sex determination. In short, there is every indica-
tion that more surprises, insights and potentially beneficial discov-
eries are in store as we explore the nature of germ cells in different
organisms.
Despite a host of worthy nominees, the unofficial consensus
was that the ‘Oskar for best picture’ of the 2002 Germ Cell
Meeting goes to “GFP-tagged zebrafish germ cells” (E. Raz,
Göttingen, Germany).
ACKNOWLEDGEMENTS
Thanks go to the organizers, Cold Spring Harbor Laboratories and all
participants for a tremendous meeting. We apologize to our many colleagues
whose excellent and fascinating work could not be covered here, and we are
grateful to the meeting presenters who allowed us to cite unpublished work.
The meeting was funded in part by the National Institute of Child Health and
Human Development (NICHD), a branch of the National Institutes of Health
(NIH), USA, the Lalor Foundation,and the March of Dimes. E.J.A.H. is
supported by the National Institute of General Medical Sciences/NIH, and
R.R.P. is supported by the NICHD/NIH and the Sandler Foundation.
reviews
©2003 EUROPEAN MOLECULAR BIOLOGY ORGANIZATION EMBO reports VOL 4 | NO 4 | 2003
meeting report
357
Dorfman, D.M. & Reijo Pera, R.A. (2003) A human homologue of
Drosophila Pumilio is expressed in embryonic stem cells and germ cells
and interacts with DAZ (Deleted in AZoospermia) protein. Proc. Natl
Acad. Sci. USA, 100, 538–543.
Pellettieri, J. & Seydoux, G. (2002) Anterior–posterior polarity in C.elegans
and Drosophila—PARallels and differences. Science, 298, 1946–1950.
Pirrotta, V. (2002) Silence in the germ. Cell, 110, 661–664.
Reese, K.J., Dunn, M.A., Waddle, J.A. & Seydoux, G. (2000) Asymmetric
segregation of PIE-1 in C. elegans is mediated by two complementary
mechanisms that act through separate PIE-1 protein domains. Mol. Cell,
6, 445–455.
Riechmann, V., Gutierrez, G.J., Filardo, P., Nebreda, A.R. & Ephrussi, A.
(2002) Par-1 regulates stability of the posterior determinant Oskar by
phosphorylation. Nature Cell Biol., 4, 337–342.
Saffman, E.E. & Lasko, P. (1999) Germline development in vertebrates and
invertebrates. Cell. Mol. Life Sci., 55, 1141–1163.
Saitou, M., Barton, S.C. & Surani, M.A. (2002) A molecular programme for the
specification of germ cell fate in mice. Nature, 418, 293–300.
Sassone-Corsi, P. (2002) Unique chromatin remodeling and transcriptional
regulation in spermatogenesis. Science, 296, 2176–2178.
Seydoux, G. & Schedl, T. (2001) The germline in C. elegans: origins,
proliferation, and silencing. Int. Rev. Cytol., 203, 139–185.
Smith, P., Leung-Chiu, W.-M., Montgomery, R., Orsborn, A., Kuznicki, K.,
Gressman-Coberly, E., Mutapcic, L. & Bennett, K. (2002) The GLH
proteins, Caenorhabditis elegans P granule components, associate with
CSN-t and KGB-1, proteins necessary for fertility, and with ZYX-1, a
predicted cytoskeletal protein. Dev. Biol., 251, 333–347.
Spradling, A., Drummond-Barbosa, D. & Kai, T. (2001) Stem cells find their
niche. Nature, 414, 98–104.
Starz-Gaiano, M. & Lehmann, R. (2001) Moving towards the next generation.
Mech. Dev., 105, 5–18.
Styhler, S., Nakamura, A. & Lasko, P. (2002) VASA localization requires the
SPRY-domain and SOCS-box containing protein, GUSTAVUS. Dev. Cell,
3, 865–876.
Surani, M.A. (2001) Reprogramming of genome function through epigenetic
inheritance. Nature, 414, 122–128.
Tanaka, S.S. & Matsui, Y. (2003) Developmentally regulated expression of mil-
1and mil-2, mouse interferon-induced transmembrane protein like
genes, during formation and differentiation of primordial germ cells.
Mech. Dev. (in the press).
Tay, J. & Richter, J.D. (2001) Germ cell differentiation and synaptonemal
complex formation are disrupted in CPEB knockout mice. Dev. Cell, 1,
201–213.
Tazuke, S.L., Schulz, C., Gilboa, L., Fogarty, M., Mahowald, A.P., Guichet, A.,
Ephrussi, A., Wood, C.G., Lehmann, R. & Fuller, M.T. (2002) A germline-
specific gap junction protein required for survival of differentiating early
germ cells. Development, 129, 2529–2539.
Tulina, N. & Matunis, E. (2001) Control of stem cell self-renewal in
Drosophila spermatogenesis by JAK–STAT signaling. Science, 294,
2546–2549.
REFERENCES
Anderson, R., Copeland, T.K., Scholer, H., Heasman, J. & Wylie, C. (2000) The
onset of germ cell migration in the mouse embryo. Mech. Dev., 91,
61–68.
Chen, D. & McKearin, D. M. (2003) A discrete transcriptional silencer in the
bam gene determines asymmetric division of the Drosophila germline
stem cell. Development, 130, 1159–1170.
Crittenden, S.L., Bernstein, D., Bachorik, J., Thompson, B., Moulder, G.,
Barstead, R., Wickens, M. & Kimble, J. (2002) FBF and control of germline
stem cells in Caenorhabditis elegans. Nature, 417, 660–663.
Cuenca, A.A., Schetter, A., Aceto, D., Kemphues, K. & Seydoux, G. (2003)
Polarization of the C. elegans zygote proceeds via distinct establishment
and maintenance phases. Development, 130, 1255–1265.
Eckmann, C.R., Kraemer, B., Wickens, M. & Kimble, J. (2002) GLD-3, a
bicaudal-C homolog that inhibits FBF to control germline sex
determination in C. elegans. Dev.Cell, 3, 697–710.
Fong, Y., Bender, L., Wang, W. & Strome, S. (2002) Regulation of the different
chromatin states of autosomes and X chromosomes in the germ line of C.
elegans. Science, 296, 2235–2238.
Hochedlinger, K. & Jaenisch, R. (2002) Nuclear transplantation: lessons from
frogs and mice. Curr.Opin. Cell Biol., 14, 741–748.
Hudson, C. & Woodland, H.R. (1998) Xpat, a gene expressed specifically in
germ plasm and primordial germ cells of Xenopus laevis. Mech. Dev., 73,
159–168.
Humpherys, D., Eggan, K., Akutsu, H., Friedman, A., Hochedlinger, K.,
Yanagimachi, R., Lander, E.S., Golub, T.R. & Jaenisch, R. (2002) Abnormal
gene expression in cloned mice derived from embryonic stem cell and
cumulus cell nuclei. Proc. Natl Acad. Sci.USA, 99, 12889–12894.
Kelly, W.G., Schaner, C.E., Dernburg, A.F., Lee, M.H., Kim, S.K., Villeneuve,
A.M. & Reinke, V. (2002) X-chromosome silencing in the germline of C.
elegans. Development, 129, 479–492.
Kiger, A.A., Jones, D.L., Schulz, C., Rogers, M.B. & Fuller, M.T. (2001) Stem
cell self-renewal specified by JAK–STAT activation in response to a
support cell cue. Science, 294, 2542–2545.
Kimura, T., Suzuki, A., Fujita, Y., Yomogida, K., Lomeli, H., Asada, N., Ikeuchi,
M., Nagy, A., Mak, T.W. & Nakano, T. (2003) Conditional loss of PTEN
leads to testicular teratoma and enhances embryonic germ cell
production. Development, 130, 1691–1700.
Kraemer, B., Crittenden, S., Gallegos, M., Moulder, G., Barstead, R., Kimble, J.
& Wickens, M. (1999) NANOS-3 and FBF proteins physically interact to
control the sperm–oocyte switch in Caenorhabditis elegans. Curr. Biol., 9,
1009–1018.
Leatherman, J.L., Levin, L., Boero, J. & Jongens, T.A. (2002) Germ cell-less acts
to repress transcription during the establishment of the Drosophila germ
cell lineage. Curr.Biol., 12, 1681–1685.
Mendez, R. & Richter, J.D. (2001) Translational control by CPEB: a means to
the end. Nature Rev. Mol. Cell Biol., 2, 521–529.
Molyneaux, K.A., Stallock, J., Schaible, K. & Wylie, C. (2001) Time-
lapse analysis of living mouse germ cell migration. Dev. Biol., 240, 488–498.
Moore, F.L., Jaruzelska, J., Fox, M., Urano, J., Firpo, M.T., Turek, P.J.,
... Germ granules may be as diverse as P-granules of somatic cells, with functions ranging from RNA localization and decay to translational activation and repression (Anderson & Kedersha, 2006; Seydoux & Braun, 2006). It is also possible, that in some cases, this machinery exists in mammals as submicroscopic complexes that are rich in RNAs and RNA-binding proteins (Hubbard & Pera, 2003). The Oskar protein nucleates the formation of polar granules de novo, from cytoplasmic pools of the components shared with nuage. ...
... Therefore, it is possible that the Arabidopsis male meiocyte is less specialized than the mouse male meiocyte. Unlike the early separation of germline cells from somatic tissues in animals (Hubbard and Pera, 2003), plant meiocytes are only a small number of mitoses from totipotent stem cells of the meristem (Ma, 2005), and possibly retain some stem cell properties. Furthermore, at both the sequence and expression levels, there is a greater degree of conservation between Arabidopsis and mouse meiocyte transcriptomes than of either to those of yeasts, even though fungi are evolutionarily closer to animals. ...
Article
Meiosis is essential for eukaryotic sexual reproduction, with two consecutive rounds of nuclear divisions, allowing production of haploid gametes. Information regarding the meiotic transcriptome should provide valuable clues about global expression patterns and detailed gene activities. Here we used RNA sequencing to explore the transcriptome of a single plant cell type, the Arabidopsis male meiocyte, detecting the expression of approximately 20 000 genes. Transcription of introns of >400 genes was observed, suggesting previously unannotated exons. More than 800 genes may be preferentially expressed in meiocytes, including known meiotic genes. Of the 3378 Pfam gene families in the Arabidopsis genome, 3265 matched meiocyte-expressed genes, and 18 gene families were over-represented in male meiocytes, including transcription factor and other regulatory gene families. Expression was detected for many genes thought to encode meiosis-related proteins, including MutS homologs (MSHs), kinesins and ATPases. We identified more than 1000 orthologous gene clusters that are also expressed in meiotic cells of mouse and fission yeast, including 503 single-copy genes across the three organisms, with a greater number of gene clusters shared between Arabidopsis and mouse than either share with yeast. Interestingly, approximately 5% transposable element genes were apparently transcribed in male meiocytes, with a positive correlation to the transcription of neighboring genes. In summary, our RNA-Seq transcriptome data provide an overview of gene expression in male meiocytes and invaluable information for future functional studies.
... Looijenga and colleagues have suggested that SOX17 may fulfil this role instead (de Jong et al., 2008). These data are interesting given the suggestion that OCT-3/4 may fulfil different functions in germ cells and stem cells through modulating different sets of target genes (Hubbard and Pera, 2003; Kehler et al., 2004). The differences in localization and onset of expression of these marker genes relative to the stage of germ cell maturation suggest significant differences may exist in the control of germ cell development between mouse and human. ...
Article
Full-text available
Germ cells have a critical role in mediating the generation of genetic diversity and transmitting this information across generations. Furthermore, gametogenesis is unique as a developmental process in that it generates highly-specialized haploid gametes from diploid precursor stem cells through meiosis. Despite the importance of this process, progress in elucidating the molecular mechanisms underpinning mammalian germ cell development has been retarded by the lack of an efficient and reproducible system of in vitro culture for the expansion and trans-meiotic differentiation of germline cells. The dearth of such a culture system has rendered the study of germ cell biology refractory to the application of new high-throughput technologies such as RNA interference, leaving in vivo gene-targeting approaches as the only option to determine the function of genes believed to be involved in gametogenesis. Recent reports detailing the derivation of gametes in vitro from stem cells may provide the first steps in developing new tools to solve this problem. This review considers the developments made in modelling germ cell development using stem cells, and some of the challenges that need to be overcome to make this a useful tool for studying gametogenesis and to realize any future clinical application.
... Despite great strides in our understanding of the genetic regulation of germ cell determination in recent years [1], the size of the founding germ cell population in humans remains obscure. Due to this uncertainty, it is difficult in a clinical environment to estimate the probability that a mutant allele known to be mosaic in the somatic tissues of a parent will be transmitted to offspring. ...
Article
Full-text available
The number of founding germ cells (FGCs) in mammals is of fundamental significance to the fidelity of gene transmission between generations, but estimates from various methods vary widely. In this paper we obtain a new estimate for the value in humans by using a mathematical model of germ cell development that depends on available oocyte counts for adult women. The germline-development model derives from the assumption that oogonial proliferation in the embryonic stage starts with a founding cells at t = 0 and that the subsequent proliferation can be defined as a simple stochastic birth process. It follows that the population size X(t) at the end of germline expansion (around the 5th month of pregnancy in humans; t = 0.42 years) is a random variable with a negative binomial distribution. A formula based on the expectation and variance of this random variable yields a moment-based estimate of a that is insensitive to the progressive reduction in oocyte numbers due to their utilization and apoptosis at later stages of life. In addition, we describe an algorithm for computing the maximum likelihood estimation of the FGC population size (a), as well as the rates of oogonial division and loss to apoptosis. Utilizing both of these approaches to evaluate available oocyte-counting data, we have obtained an estimate of a = 2 - 3 for Homo sapiens. The estimated number of founding germ cells in humans corresponds well with values previously derived from chimerical or mosaic mouse data. These findings suggest that the large variation in oocyte numbers between individual women is consistent with a smaller founding germ cell population size than has been estimated by cytological analyses.
Article
Objective: The current study was designed to identify whether transcript of VASA was changed in oligospermatism compared to normal fertile men, and to determine whether VASA and BMP-4 affect spermatogenesis, as well as the relationship between VASA and BMP-4, I type receptor Alk-3 on spermatogenesis. Methods: Ejaculates of normal and oligospermatism were collected by masturbation and subsequently segregation through discontinuous gradient of Percoll to remove the residual somatic cells and purified spermatozoa, immature spermatogenic cells (spermatogonium and spermatocyte). Total RNA was extracted, and RT-PCR, Real-time PCR, immunofluorescence and Western Blot (WB) were used to assess the gene transcript. Results: VASA expression was significantly reduced in oligospermatism. Real-time PCR analysis to indicate that transcription of VASA was greater by approximately five fold in normal spermatozoa than in oligospermatism group. VASA protein was detected on spermatozoa of both normal and abnormal groups by immunofluorescence and Western Blot, lower level was detected on the oligospermatism group. BMP-4 and Alk-3 were undetectable by RT-PCR in both normal and abnormal spermatozoa, but expressed on the immature spermatogenic cells. Normal human testis expressed BMP-4 and Alk-3. Conclusion: Identification of VASA various expression profile in normal and oligospermatism reveals that VASA was associated with spermatogenesis. BMP-4 and receptor Alk-3 do not directly upregulate expression of VASA. Testes transcript included the counterpart of spermatozoa. The expression of BMP-4 and Alk-3 from the testes may be due to spermatogonium and spermatocyte. Our findings suggest VASA play a role in the spermatogenesis or serve as valuable markers of fertility status and provide a genetic fingerprint of normal fertile men.
Article
George Williams indicated that he would not expect senescence to evolve in organisms that lack a distinction between germ line and soma. Escherichia coli--long assumed to lack even a hint of this distinction--is now known to senesce, posing what would seem to be a challenge to Williams's well-known theory of the evolution of senescence. However, in this review, I will show that cell division in E. coli produces a degree of germ-soma modularity sufficient to generate age structure and antagonistic pleiotropic effects, thereby satisfying the requirements of Williams's theory. From this perspective, senescence in E. coli is supportive and points the way to a better understanding of the pleiotropies that connect adaptive complexity and senescence. Sexual reproduction is but one of the complex adaptations illuminated by this approach.
Article
The genome of germline committed cells is thought to be protected by mechanisms of transcriptional silencing, posing a barrier to transgenesis using cultured germline cells. We found that selection for transgene integration into the primordial germ cell genome required that the transgenes be flanked by the chicken beta-globin insulator. However, integration frequency was low, and sequencing of the insertion sites revealed that the transgenes preferentially inserted into active promoter regions, implying that silencing prohibited recovery of insertions in other regions. Much higher frequencies of integration were achieved when the phiC31 integrase was used to insert transgenes into endogenous pseudo attP sites. Despite the evidence for transcriptional silencing in PGCs, gene targeting of a nonexpressed gene was also achieved. The ability to make genetic modifications in PGCs provides unprecedented opportunities to study the biology of PGCs, as well as produce transgenic chickens for applications in biotechnology and developmental biology.
Article
Full-text available
The regulated translation of messenger RNA is essential for cell-cycle progression, establishment of the body plan during early development, and modulation of key activities in the central nervous system. Cytoplasmic polyadenylation, which is one mechanism of controlling translation, is driven by CPEB--a highly conserved, sequence-specific RNA-binding protein that binds to the cytoplasmic polyadenylation element, and modulates translational repression and mRNA localization. What are the features and functions of this multifaceted protein?
Article
cDNAs specific to vegetal poles of Xenopus gastrula embryos were used as a probe to screen a gastrula vegetal pole cDNA library. One of the novel clones isolated had an RNA expression pattern consistent with it being a component of germ plasm and it was thus named Xpat (Xenopus primordial germ cell associated transcript). The open reading frame encodes a 35 kDa protein with no clear homologies. The RNA is localised to the vegetal pole throughout oogenesis and early cleavage. During gastrulation cells containing this message move internally and at tailbud stages they migrate in an antero-dorsal direction. Xpat mRNA is not detectable once the dorsal mesentery forms. We show that the 3'-UTR is required and is sufficient for localisation of exogenous RNA to the vegetal pole of oocytes. We propose that Xpat UTR-containing transcripts can be localised by the Vg1 or late pathway of mRNA localisation during stage III of oogenesis, but endogenous Xpat appears to be localised earlier by a mitochondrial cloud mechanism similar to that proposed for Xcat-2.
Article
In all animals information is passed from parent to offspring via the germline, which segregates from the soma early in development and undergoes a complex developmental program to give rise to the adult gametes. Many aspects of germline development have been conserved throughout the animal kingdom. Here we review the unique properties of germ cells, the initial determination of germ cell fates, the maintenance of germ cell identity, the migration of germ cells to the somatic gonadal primordia and the proliferation of germ cells during development invertebrates and invertebrates. Similarities in germline development in such diverse organisms as Drosophila melanogaster, Caenorhabditis elegans, Xenopus laevis and Mus musculus will be highlighted.
Article
The Caenorhabditis elegans FBF protein and its Drosophila relative, Pumilio, define a large family of eukaryotic RNA-binding proteins. By binding regulatory elements in the 3' untranslated regions (UTRs) of their cognate RNAs, FBF and Pumilio have key post-transcriptional roles in early developmental decisions. In C. elegans, FBF is required for repression of fem-3 mRNA to achieve the hermaphrodite switch from spermatogenesis to oogenesis. We report here that FBF and NANOS-3 (NOS-3), one of three C. elegans Nanos homologs, interact with each other in both yeast two-hybrid and in vitro assays. We have delineated the portions of each protein required for this interaction. Worms lacking nanos function were derived either by RNA-mediated interference (nos-1 and nos-2) or by use of a deletion mutant (nos-3). The roles of the three nos genes overlap during germ-line development. In certain nos-deficient animals, the hermaphrodite sperm-oocyte switch was defective, leading to the production of excess sperm and no oocytes. In other nos-deficient animals, the entire germ line died during larval development. This germ-line death did not require CED-3, a protease required for apoptosis. The data suggest that NOS-3 participates in the sperm-oocyte switch through its physical interaction with FBF, forming a regulatory complex that controls fem-3 mRNA. NOS-1 and NOS-2 also function in the switch, but do not interact directly with FBF. The three C. elegans nanos genes, like Drosophila nanos, are also critical for germ-line survival. We propose that this may have been the primitive function of nanos genes.
Article
Mouse primordial germ cells (PGCs) are specified between embryonic day 6.5 (E6.5) and E7.5, when they have been visualized as an alkaline phosphatase-positive (AP+) cell population in the developing allantois. By E8.5, they are embedded in the hind-gut epithelium. Previous experiments have suggested different sites for PGCs' origin, and it is unclear how they reach the gut epithelium. We have used transgenic mice expressing GFP under a truncated Oct4 promoter to visualize living PGCs. We find GFP+/AP+ cells in the posterior end of the primitive streak as a dispersed population of cells actively migrating into the allantois, and directly into the adjacent embryonic endoderm. Time-lapse analysis shows these cells to be actively migratory from the time they exit the primitive streak.
Article
The CCCH finger protein PIE-1 is a regulator of germ cell fate that segregates with the germ lineage in early embryos. At each asymmetric division, PIE-1 is inherited preferentially by the germline daughter and is excluded from the somatic daughter. We show that this asymmetry is regulated at the protein level by two complementary mechanisms. The first acts before cell division to enrich PIE-1 in the cytoplasm destined for the germline daughter. The second acts after cell division to eliminate any PIE-1 left in the somatic daughter. The latter mechanism depends on PIE-1's first CCCH finger (ZF1), which targets PIE-1 for degradation in somatic blastomeres. ZF1s in two other germline proteins, POS-1 and MEX-1, are also degraded in somatic blastomeres, suggesting that localized degradation also acts on these proteins to exclude them from somatic lineages.
Article
Germ cells are essential for reproduction, yet the molecular mechanisms that underlie their unique development are only beginning to be understood. Here we review important events that lead to the establishment of the germline and the initiation of meiotic development in C. elegans. Formation of the germline begins in the pregastrulation embryo, where it depends on polarization along the anterior/posterior axis and on the asymmetric segregation of P granules and associated factors. During postembryonic development, the germline expands using the GLP-1/Notch signaling pathway to promote proliferation and regulate entry into meiosis. Throughout their development, germ cells also employ unique "silencing" mechanisms to regulate their genome and protect themselves against unwanted expression from repetitive sequences including transposable elements. Together these mechanisms preserve the health and reproductive potential of the germline.
Article
In most organisms, primordial germ cells are set aside from the cells of the body early in development. To form an embryonic gonad, germ cells often have to migrate along complex routes through and along diverse tissues until they reach the somatic part of the gonad. Recent advances have been made in the genetic analysis of these early stages of germ line development. Here we review findings from Drosophila, zebrafish, and mouse; each organism provides unique insight into the mechanisms that determine germ cell fate and the cues that may guide their migration.
Article
Most cells contain the same set of genes and yet they are extremely diverse in appearance and functions. It is the selective expression and repression of genes that determines the specific properties of individual cells. Nevertheless, even when fully differentiated, any cell can potentially be reprogrammed back to totipotency, which in turn results in re-differentiation of the full repertoire of adult cells from a single original cell of any kind. Mechanisms that regulate this exceptional genomic plasticity and the state of totipotency are being unravelled, and will enhance our ability to manipulate stem cells for therapeutic purposes.
Article
The concept that stem cells are controlled by particular microenvironments known as 'niches' has been widely invoked. But niches have remained largely a theoretical construct because of the difficulty of identifying and manipulating individual stem cells and their surroundings. Technical advances now make it possible to characterize small zones that maintain and control stem cell activity in several organs, including gonads, skin and gut. These studies are beginning to unify our understanding of stem cell regulation at the cellular and molecular levels, and promise to advance efforts to use stem cells therapeutically.