Sie sind auf Seite 1von 22

Germ Layer Induction in UNIT 1D.

1
ESC—Following the Vertebrate Roadmap
Jim Smith,1 Fiona Wardle,1 Matt Loose,2 Ed Stanley,3 and Roger Patient4
1
Wellcome Trust/Cancer Research UK Gurdon Institute, University of Cambridge,
Cambridge, United Kingdom
2
Institute of Genetics, University of Nottingham, Nottingham, United Kingdom
3
Monash Immunology and Stem Cell Laboratories, Monash University, Clayton,
Victoria, Australia
4
Weatherall Institute of Molecular Medicine, University of Oxford, Oxford,
United Kingdom

ABSTRACT
Controlled differentiation of pluripotential cells takes place routinely and with great success in
developing vertebrate embryos. It therefore makes sense to take note of how this is achieved
and use this knowledge to control the differentiation of embryonic stem cells (ESCs). An added
advantage is that the differentiated cells resulting from this process in embryos have proven
functionality and longevity. This unit reviews what is known about the embryonic signals that drive
differentiation in one of the most informative of the vertebrate animal models of development, the
amphibian Xenopus laevis. It summarizes their identities and the extent to which their activities
are dose-dependent. The unit details what is known about the transcription factor responses to
these signals, describing the networks of interactions that they generate. It then discusses the
target genes of these transcription factors, the effectors of the differentiated state. Finally, how
these same developmental programs operate during germ layer formation in the context of ESC
differentiation is summarized. Curr. Protoc. Stem Cell Biol. 1:1D.1.1-1D.1.22.  C 2007 by John
Wiley & Sons, Inc.
Keywords: embryonic signals r transcription factors r differentiation r Xenopus r
zebrafish r mouse r mesendoderm r mesoderm r endoderm r ectoderm r neural r Wnt r
TGF-β r nodal r activin r BMP r FGF r antagonists r T-box r VegT r brachyury r Mix r
Sox r GATA r ES cells

LESSONS FROM FROGS (ESC). The unit focuses on early developmen-


(AND FISH) tal decisions, in which cells become commit-
ted first to endoderm, mesoderm, or ectoderm,
Inductive Interactions in the Early and then to particular regions of these germ
Embryo layers, such as dorsal versus ventral meso-
The most important mechanism by which
derm or neural versus non-neural ectoderm.
cell fate is determined in the developing verte-
This is done because it is necessary that cul-
brate embryo involves inductive interactions,
tured ESC successfully negotiate these early
in which one group of cells makes a signal that
decisions before they follow pathways leading
is received, and acted upon, by adjacent tissue.
to the formation of specialized cell types such
Interactions of this sort are most easily studied
as pancreas, heart, or liver.
in embryos that are accessible to the experi-
menter, including those of the chick, zebrafish,
and frog, and among these species, most in- The first step: Dorsalization of the
sights have come from the frog, Xenopus lae- embryo
vis. In this introductory unit the authors sum- Most of the understanding on inductive in-
marize these results, discuss how they might teractions during early vertebrate development
apply to other species, including mammalian derives from work on the amphibian species,
embryos, and then consider to what extent Xenopus laevis. Xenopus offers many advan- Germ Layer
they might allow the experimenter to manipu- tages to the developmental biologist: the em- Induction/
late the differentiation of embryonic stem cells bryo is large, and therefore easy to inject, Differentiation
of Embryonic
Stem Cells

Current Protocols in Stem Cell Biology 1D.1.1-1D.1.22 1D.1.1


Published online June 2007 in Wiley Interscience (www.interscience.wiley.com).
DOI: 10.1002/9780470151808.sc01d01s1 Supplement 1
Copyright C 2007 John Wiley & Sons, Inc.
dissect, and manipulate; it is available in large large amounts of blood (Cooke and Smith,
numbers, and therefore suitable for biochemi- 1987). The loss of anterior tissue is not due to
cal approaches; it has, at early stages, a reliable nonspecific effects of UV irradiation, because
fate map, so that one can reliably interpret ex- rotation can be restored by tipping the fer-
periments that are designed to alter cell fates; tilized egg to one side, and this completely
and its cells are laden with yolk, so that they rescues the embryo, such that it forms a per-
can survive in a very simple salts solution in fectly normal tadpole (Scharf and Gerhart,
the absence of poorly characterized compo- 1980).
nents such as fetal bovine serum. Together, It is not known how rotation specifies dorsal
these advantages have made Xenopus the or- and anterior regions of the embryo, although it
ganism nonpareil for the analysis of inductive is clear that signaling by members of the Wnt
interactions. pathway is involved. Thus, ectopic expression
When it is laid, the Xenopus egg is a sphere of members of the Wnt family in the early
about 1.4-mm in diameter. It has a heavily Xenopus embryo causes the formation of an
pigmented animal hemisphere, which comes additional head (McMahon and Moon, 1989;
to lie uppermost in the water, and a paler, Smith and Harland, 1991; Sokol et al., 1991),
yolkier, vegetal half that forms the southern and inhibition or stimulation of Wnt signal-
hemisphere. At this early stage, and, as far as ing causes, respectively, the loss or enhance-
is known, in contrast to the mammalian egg, ment of anterior/dorsal structures (Heasman
some RNAs are differentially localized within et al., 1994; Kofron et al., 2001). Most signif-
this large cell. Two such RNAs are of particular icantly, depletion of maternal Wnt11 mRNA
note. One encodes VegT (also known as Brat, causes embryos to develop without a head or
Xombi, and Antipodean), a member of the anterior structures, a phenotype that can be
T-box family of transcription factors (Lustig rescued by the re-introduction of exogenous
et al., 1996; Stennard et al., 1996; Zhang and Wnt11 mRNA (Tao et al., 2005). Rotation of
King, 1996; Horb and Thomsen, 1997), and the cortical layer of cytoplasm is now thought
the other encodes Vg1, a member of the trans- to reposition maternal Wnt11 mRNA from its
forming growth factor type β (TGF-β) family original vegetal location to a position closer
(Weeks and Melton, 1987). More of these will to the equator of the embryo, where it is free
be discussed later in the unit. For now, suffice to diffuse into deeper cytoplasm. This allows
it to say that these RNAs move to their veg- Wnt11 protein to accumulate in dorsal veg-
etal positions during oogenesis, and that their etal cells and to activate the canonical Wnt
position in this region of the embryo is crucial signal transduction pathway in this region of
for normal development. the embryo. It is also possible that cortical
Immediately after being laid the egg ap- rotation causes the dorsal enrichment of Wnt
pears, and is, radially symmetrical about the co-receptors, or components of the Wnt sig-
animal-vegetal axis. Polarity is imposed upon nal transduction pathway, in a mechanism that
the egg at the time of fertilization, when the po- would further enhance the level of Wnt signal-
sition of sperm entry defines the side of the egg ing in this region of the embryo (Dominguez
that will eventually form posterior and ventral and Green, 2000).
structures (Vincent and Gerhart, 1987). The One of the consequences of Wnt signaling
mechanism by which this occurs is not well in the dorso-vegetal region of the embryo is
understood, but it is associated with the ro- the activation of the Siamois gene, which is
tation, of ∼30◦ , of a shell of cortical cyto- involved in the formation of dorsal and axial
plasm, just beneath the egg plasma membrane tissues (Lemaire et al., 1995; Fan and Sokol,
(Vincent and Gerhart, 1987). The position of 1997; Engleka and Kessler, 2001).
sperm entry defines the orientation of this ro-
tation, and the significance of the rotation for The second step: Mesoderm induction
the specification of the anterior/dorsal to pos- As is remarked above, there is no evidence
terior/ventral axis is demonstrated by experi- yet for localized RNAs or other determinants
ments in which the rotation is inhibited. For in the mammalian egg, and indeed the ex-
example, this can be done by irradiating the tent to which the early mammalian embryo
vegetal hemisphere of the newly fertilized em- is patterned at preblastocyst stages remains
bryo with UV light (Scharf and Gerhart, 1980; controversial (Hiiragi et al., 2006; Zernicka-
Holwill et al., 1987). This prevents rotation Goetz, 2006). However, at later stages, sim-
and the embryo develops without a head or ilarities between amphibian and amniote de-
Germ Layer axial structures. All that is formed is a mass velopment are more obvious, and the analysis
Induction in ESC
of ventral tissue that includes, for example, of mesoderm induction in Xenopus has indeed
1D.1.2
Supplement 1 Current Protocols in Stem Cell Biology
informed studies of mouse development and cell types, and fail to induce the expression
of ESC differentiation (Gadue et al., 2006). of genes that are expressed in prospective an-
The mesoderm of the amphibian embryo terior and dorsal tissues, such as goosecoid
derives from the equatorial region of the em- (Green et al., 1990). In contrast, the members
bryo. Pioneering experiments by Nieuwkoop of the TGF-β family tend to induce a wide
showed that this germ layer (and part of the spectrum of tissues, from ventral and poste-
endoderm) forms in this position as the re- rior to dorsal and anterior, and they do this in a
sult of an inductive interaction in which cells concentration-dependent manner, with lower
of the vegetal hemisphere (which inherit, for concentrations inducing posterior/ventral tis-
example, Vg1 RNA and protein from the veg- sues and high doses inducing anterior and dor-
etal region of the fertilized egg) act on over- sal structures (Green and Smith, 1990; Green
lying cells (Nieuwkoop, 1969; Sudarwati and et al., 1992; Green, 1994; Gurdon et al., 1994,
Nieuwkoop, 1971). Nieuwkoop demonstrated 1995, 1996). These dose-dependent effects of
this interaction by dissecting cells from the the TGF-β family are discussed below.
animal pole region of the embryo (which The FGF and the TGF-β families employ
normally become ectoderm), and juxtapos- different signal transduction pathways, the for-
ing them with vegetal cells (which normally mer inducing mesoderm through the MAP ki-
become endoderm). Individually neither cell nase pathway (Gotoh et al., 1995; LaBonne
population would form mesoderm, but the et al., 1995; Umbhauer et al., 1995) and the
combination of cells produced large amounts latter through the Smad family (Hill, 2001).
of mesodermal tissue such as muscle, and lin- Limited knowledge of the transcriptional reg-
eage labeling experiments revealed that the ulation of known FGF and TGF-β target genes
mesoderm derived from the ectodermal tis- (Watabe et al., 1995; Howell and Hill, 1997;
sue, indicating that the inducing signal derives Latinkic et al., 1997; Howell et al., 1999; Ger-
from the endoderm. main et al., 2000; Lerchner et al., 2000) pro-
vides a molecular basis for the difference in the
Mesoderm inducing factors inducing activities of the two classes of sig-
Work in the late 1980s and early 1990s re- naling molecules. Most significant, however,
vealed that two classes of signaling molecules is the conclusion confirmed by one of the first
have the ability to mimic the effect of the experiments to employ transgenesis in Xeno-
vegetal cells. These include members of the pus species that the role of FGF signaling is
fibroblast growth factor (FGF; Slack et al., to maintain mesodermal identity rather than to
1987; Kimelman et al., 1988) and transform- induce it (Kroll and Amaya, 1996).
ing growth factor type β (TGF-β) families (Al-
bano et al., 1990; Asashima et al., 1990; Smith Concentration-dependent effects of inducing
et al., 1990; Thomsen et al., 1990). The most factors
powerful inducers proved to be members of It is interesting that members of the TGF-
the TGF-β family, and especially activin and β family are able to induce different types
the Xenopus nodal-related (Xnr) genes (Jones of mesoderm, and to activate the expression
et al., 1995; Joseph and Melton, 1997; Taka- of different mesoderm- and endoderm-specific
hashi et al., 2000), and since this time two genes, at different concentrations. Unfortu-
other related proteins have been characterized nately, rather little is known about the mech-
as mesoderm-inducing factors: Vg1, whose anism through which this occurs. Work by
transcripts are expressed in the vegetal region Gurdon and colleagues has shown that 100
of the oocyte and egg (Weeks and Melton, molecules of activin bound to a single ani-
1987; Dale et al., 1993; Thomsen and Melton, mal pole cell are sufficient to induce expres-
1993; Birsoy et al., 2006), and derrière, which, sion of the pan-mesodermal gene brachyury,
like activin and the nodal-related genes, is ex- while 300 molecules are required to extin-
pressed zygotically (Sun et al., 1999a). guish brachyury and to activate goosecoid
When applied to isolated animal pole re- (Dyson and Gurdon, 1998). It is also known
gions, all of these inducing factors, including that protein synthesis is required for the down-
members of both the FGF and TGF-β fami- regulation of brachyury expression that oc-
lies, can cause the activation of mesoderm- and curs at high doses of activin, suggesting that
endoderm-specific genes and the differenti- these high doses of a TGF-β family member
Germ Layer
ation of mesodermal cell types. One differ- induce a repressor of brachyury (Papin and Induction/
ence between the two families of signaling Smith, 2000). One candidate for such a re- Differentiation
molecules is that FGF family members in- pressor is Goosecoid itself, and indeed muta- of Embryonic
Stem Cells
duce predominantly ventral and posterior tion of a Goosecoid binding site in a 381-base
1D.1.3
Current Protocols in Stem Cell Biology Supplement 1
pair brachyury promoter fragment prevents the route such as transcytosis (Williams et al.,
down-regulation of a brachyury reporter con- 2004).
struct that occurs at high activin concentra- Refining the pattern: Inhibiting BMP
tions (Latinkic et al., 1997). However, it is signaling
unlikely that Goosecoid is the only repressor Different concentrations of TGF-β fam-
of brachyury that is induced at high levels of ily members, such as activin and the nodal-
activin, because inhibition of Goosecoid ac- related genes, can establish differences be-
tivity does not affect the down-regulation of tween the prospective anterior/dorsal and pos-
endogenous brachyury that occurs at high ac- terior/ventral regions of the late blastula and
tivin concentrations (Papin and Smith, 2000). early gastrula regions of the embryo. These
This phenomenon requires further investiga- differences, however, are rather crude. Dissec-
tion, because, as is discussed below, it is likely tion of tissue from different regions along this
that pattern formation in the developing em- axis of the embryo reveals that there are only
bryo occurs in response to gradients of TGF-β two well-defined domains, a smaller one on
family members, and attempts to direct ESC the dorsal side of the embryo, where the dorsal
differentiation must take account of the con- lip of the blastopore will appear, and a larger
centrations of the inducing factors that are one that comprises the rest of the equatorial
used. region of the embryo (Dale et al., 1985). Cul-
The most direct evidence that gradients of ture of the former region reveals that this is
inducing factors activate different genes in dif- specified to form notochord and muscle, while
ferent regions of the embryo, and specify dif- the latter region forms predominantly blood.
ferent cell types, comes from experiments in Significantly, these results stand in contrast to
which the functions of the genes are inhibited. the fate map of the embryo at this stage, which
Developmental biologists know of eight TGF- shows that most of the muscle of the embryo
β family members that are expressed in the derives from the larger, ventral, region of the
embryo and that have mesoderm-inducing ac- prospective mesoderm (Dale and Slack, 1987).
tivity: Vg1, activin, Xnr1, Xnr2, Xnr4, Xnr5, How can these observations be reconciled?
Xnr6, and derrière (see references above). It Further grafting experiments show that the re-
has not yet proved possible to inhibit the ac- gion of the embryo that will form the dorsal
tion of all of these factors individually, but anti- lip of the blastopore is the source of signals
sense and dominant-negative approaches have that can dorsalize adjacent ventral mesoderm
demonstrated that the maternally expressed (Smith and Slack, 1983). Indeed, if this region,
gene Vg1 (Birsoy et al., 2006), and the zygoti- known as Spemann’s organizer, is grafted to
cally activated activin (Piepenburg et al., 2004) the ventral region of a host embryo, the host
and derrière (Sun et al., 1999a) are all required develops a secondary axis on its ventral side
for proper mesoderm formation. Of the nodal- (Spemann, 1938). Interactions between Spe-
related genes, only Xnr1 has been studied indi- mann’s organizer and the rest of the embryo re-
vidually (Toyoizumi et al., 2005), and it proves fine the spatial patterns of gene expression and
to be required for proper specification of the cell differentiation along the anterior/dorsal to
left-right axis of the embryo. However, simul- posterior/ventral axis.
taneous inhibition of all the Xenopus nodal- Identification of the signals produced by
related genes by increasing expression of a Spemann’s organizer that are responsible for
truncated form of Cerberus, termed Cerberus- dorsalization came from several types of ex-
short, causes the progressive loss of dorsal and periments. Expression cloning identified the
then ventral gene expression, consistent with gene noggin (Smith and Harland, 1992), which
the idea that gradients of these TGF-β family was capable of rescuing embryos that had
members specify cell types in the developing been ventralized by treatment with UV light
embryo (Agius et al., 2000). (see above). An in situ hybridization screen
It is not yet clear how such gradients are showed that chordin is expressed in the dor-
established or how the inducing factors tra- sal region of the embryo, with subsequent ex-
verse fields of responding cells, although the periments revealing that it too had dorsalizing
higher expression of the nodal-related genes activity (Sasai et al., 1994); and analysis of
at the dorsal side of the embryo is thought follistatin, whose gene product was already
to derive from the enhanced VegT, Vg1, and known to inhibit activin signaling, showed
Wnt signaling in this region of the embryo that this gene is also expressed dorsally and
(Agius et al., 2000), and it seems likely that is capable of dorsalizing ventral mesoderm
Germ Layer inducing factors travel in the extracellular (Hemmati-Brivanlou et al., 1994). All of these
Induction in ESC
milieu rather than through an intracellular factors have in common the fact that they can
1D.1.4
Supplement 1 Current Protocols in Stem Cell Biology
inhibit signaling by bone morphogenetic pro- Piccolo et al., 1999). Ectopic expression of
teins (BMPs), factors which previously had Cerberus in the Xenopus embryo causes the
been shown to have powerful ventralizing ac- formation of an extra head (Bouwmeester
tivity in the Xenopus embryo (Dale et al., et al., 1996), indicating, remarkably, that for-
1992; Jones et al., 1992), and subsequent work mation of the head does not require special or
demonstrated that these inhibitors can set up novel inducing activities but rather the inhibi-
a gradient in the embryo, high in the ante- tion of several signaling pathways.
rior/dorsal region and low in posterior/ventral
tissues (Jones and Smith, 1998). BMP fam- Later inductive interactions and
ily members such as BMP4 are expressed implications for embryonic stem cells
in a widespread fashion throughout the em- This description of the induction and pat-
bryo (with the exception of the dorsal region; terning of the mesoderm has emphasized (1)
Hemmati-Brivanlou and Thomsen, 1995), so the importance of intercellular signaling by se-
that the gradient of BMP inhibitors establishes creted polypeptide growth factors, (2) the fact
a reverse gradient of BMP activity. It is this that these molecules can exert concentration-
gradient that creates positional information dependent effects, and (3) the observation that
along the anterior/dorsal to posterior/ventral effective gradients of such growth factors can
axis of the embryo. be established by reverse gradients of secreted
Like Vg1, activin, the Xnr proteins, and inhibitors. These principles also apply to later
derrière, BMPs are members of the TGF-β stages of embryonic development. For exam-
family, but they differ because they signal ple, neural induction by Spemann’s organizer
through different cell surface receptors and in the gastrula-stage embryo is largely a con-
different members of the Smad family. Activin sequence of the inhibition of BMP signaling in
and the other mesoderm-inducing factors sig- the prospective ectoderm (Harland, 1994), to-
nal through Smad2 or Smad3, which form het- gether with contributions from FGF and prob-
eromeric complexes with Smad4; BMPs sig- ably other signaling families (Launay et al.,
nal through Smad1, Smad5, and Smad8, which 1996; Sasai et al., 1996; Linker and Stern,
also form complexes with Smad4 (Hill, 2001; 2004). At later stages and in other tissues,
Schier, 2003). other families of signaling molecules also play
a role, including hedgehog and notch signal-
The organizer produces many inhibitors ing. A thorough knowledge of these signaling
The discovery of noggin, chordin, and events will be invaluable in coming to under-
follistatin as factors that pattern the ante- stand the ways in which one can direct and
rior/dorsal to posterior/ventral axis of the em- influence ESC differentiation.
bryo by inhibiting BMP signaling was quickly
followed by the identification of other in- Integration of Intracellular Responses
hibitors of extracellular signaling molecules. to Embryonic Induction
These include Frzb-1 (Leyns et al., 1997), The distinct combinations and levels of em-
a secreted protein containing a domain sim- bryonic signals experienced by different cells
ilar to the putative Wnt-binding region of the in the embryo differentially induce the expres-
frizzled family of transmembrane receptors, sion or activities of transcription factors (TFs);
and Dickkopf-1 (Dkk1; Glinka et al., 1998), and the particular combination of TFs active in
a ligand for the Wnt co-receptor LRP6. Both a given cell determines its phenotype. The re-
of these molecules inhibit signaling by Wnt sponses of TFs to embryonic signals together
family members, and have the effect of in- with interactions between TFs constitute a ge-
hibiting the ventralizing effects of proteins netic regulatory network (GRN), describing
such as Wnt8, which is expressed through- the state of a given cell at a particular time
out the lateral and posterior/ventral regions of (Loose and Patient, 2004; Koide et al., 2005).
the early gastrula-staged embryo. They there- To understand germ layer induction, it is nec-
fore establish an effective reverse gradient of essary to fully understand the construction,
Wnt signaling, further refining positional in- dynamics, and stability of these networks as
formation along the dorso-ventral axis of the differentiation proceeds.
embryo. Here, because of the volume of data avail-
Perhaps the most remarkable inhibitor able, the authors concentrate on mesendoderm
Germ Layer
produced by Spemann’s organizer, however, induction in Xenopus, but much of what has Induction/
is Cerberus, which inhibits the activities been found in Xenopus is the same or very Differentiation
of BMP, Wnt, and nodal-related signaling similar in zebrafish, making it very likely that of Embryonic
Stem Cells
(Bouwmeester et al., 1996; Glinka et al., 1997; the fundamental mechanisms discussed here
1D.1.5
Current Protocols in Stem Cell Biology Supplement 1
Figure 1D.1.1 Legend at right.

serve as a model for vertebrate germ layer for- In order to begin to appreciate the impli-
mation generally. In vertebrates, the three pri- cations of the Xenopus network for our un-
mary germ layers, ectoderm, mesoderm, and derstanding of ESC programming, the authors
endoderm, are classically regarded as the first have built a simplified version of this network,
signs of differentiation in the embryo proper, which takes account of reduced gene numbers
with differentiation to extra-embryonic tis- in mammals and illustrates the key principles
sues taking place concomitantly. However, a (Fig. 1D.1.1A). GRNs are composed of many
number of observations suggest that the ini- smaller networks, termed motifs (for a review
tial separations may not be as clean as pre- see Lee et al., 2002 or Babu et al., 2004).
viously thought (Rodaway and Patient, 2001; These motifs occur more often than would
Wardle and Smith, 2004). In addition, subdivi- be expected in a random network with the
sions within the mesoderm become apparent same degree of connectivity as the network
in some cases before mesodermal and endo- under study. Such motifs are easiest to identify
dermal fates have been distinguished. Thus, computationally, and the program mFinder has
skeletal muscle and notochord progenitors be- been used to identify motifs in a recent update
come distinguishable from endodermal pre- of the Xenopus mesendoderm network used
cursors before blood and cardiovascular pro- (Milo et al., 2002; Loose and Patient, 2004).
genitors. Observations of this nature have led The complete network is scale-free, that is, all
to the use of the term mesendoderm, and com- genes do not display equal connectivity. Some
parisons with more primitive organisms indi- of the TFs studied to date display substantially
Germ Layer cate that the distinction may in fact be quite more connectivity than others making them
Induction in ESC
ancient. potentially master regulators. The following
1D.1.6
Supplement 1 Current Protocols in Stem Cell Biology
will highlight some of these and the motifs are of the form that would be expected to ac-
found in the Xenopus GRN, and will try to celerate the onset of expression of the target.
draw out the lessons to be learned for ESC The targets of the feed forward motifs in the
differentiation. Xenopus GRN include both mesodermal and
endodermal TFs, suggesting that cells in which
Feed forward loops maternal VegT is expressed have the potential
The most striking thing about mesendo- to become both mesoderm and endoderm, i.e.,
derm formation in Xenopus is the extent to a molecular demonstration of their mesendo-
which the maternal T-box TF, VegT, domi- dermal potential. Of the remaining 22 feed for-
nates proceedings (Zhang et al., 1998; Xanthos ward loops identified in the updated Xenopus
et al., 2001; Taverner et al., 2005; Heasman, mesendodermal network, 13 are initiated by
2006). Of 54 genes currently in the network, nodal-related signals, 5 by β-catenin, driven
26 are regulated directly by VegT. A clue to maternally by Wnt11 signaling (Tao et al.,
its function in the network emerges when con- 2005), and 4 by Mix family TFs or GATA
sidering which network motifs feature VegT. factors.
VegT features heavily in a common motif, the Thus, in the Xenopus embryo, the local-
feed forward loop (see Fig. 1D.1.1B). Feed ization of maternal determinants is a crucial
forward loops consist of three genes, X, Y, and aspect of mesendoderm formation. The sig-
Z, where X regulates Y and Z, and Y regulates Z nificance of this for programming ESCs re-
(Milo et al., 2002). Studies of the properties of mains to be seen. For example, even though
feed forward loops reveal two likely functions: zebrafish express a maternal T-box TF, its
either accelerated or delayed expression of the activities have not yet been demonstrated to
target Z, depending on the nature of the inter- be equivalent to VegT in Xenopus (Bjornson
actions between the three genes (activating or et al., 2005), so it is possible that this is a de-
repressing) and the way in which the X and rived characteristic of Xenopus. Alternatively,
Y genes interact at Z (both or either required the loss of such a wide range of activities
for expression; see Mangan et al., 2003). Of might be a derived characteristic of zebrafish,
the 63 such feed forward motifs identified in in which case, one should be looking for the
the Xenopus network, 41 are initiated by VegT. distribution of T-box transcription factors in
Furthermore, the majority of these 41 motifs ESCs and determining the extent to which they

Figure 1D.1.1 (at left) (A) A reduced network representing the key interactions occurring within
the Xenopus mesendoderm network at the start of gastrulation (approximately stage 10.5; Loose
and Patient, 2004). In order to reduce the complexity of the network, the authors have combined
the interactions of key gene family members into one representative gene. Thus, the nodal family
members (and derrière) are combined into one nodal gene, the mix family members into one mix
and the GATA genes into one GATA gene. For the nodal and mix families, this reduction reflects
the fact that there is only one mix and one nodal homolog in mouse and human. In contrast, the
reduction of the GATA family into one representative reflects the difficulty of distinguishing the
targets of individual GATA factors at this time, although recent work has begun to make some
headway (Afouda et al., 2005). The interactions shown in the diagram represent the combined
input and output of all family members in the original network. No explicit conflicts were found
within families (i.e., one family member positively regulating, another negatively), although for
some interactions, individual members have been shown to regulate a subset of the targets of the
whole family. In these cases the authors have assumed that the ability to regulate a given target
has been lost by the individual family member as opposed to having been gained by several family
members. In order to clearly represent the topology of the network, genes that do not feed back
into the network have not been shown. The exception to this is the two targets of the GATA family,
sox17 and HNF1β. At later time-points these genes will feed back into the network. sox17 and
HNF1β also represent endoderm within the network. Dashed lines illustrate interactions that are
initiated by a maternal message, although later, zygotic expression will take over. Examples of
feed forward loops (B) can be traced in this network, for example VegT to nodal to mix (illustrated
in C) and mix to GATA to sox17. (B) Illustration of a feed forward motif. Gene X activates both
Y and Z, gene Y activates Z. A feed forward loop can have any combination of activation and
repression, resulting in 8 different sub-classes of feed forward loop. (C) A sample feed forward
loop involving VegT, nodal, and mix. Note that nodal is able to autoregulate and can therefore Germ Layer
maintain mix expression in the absence of VegT. (D) This motif is extracted from panel A and Induction/
Differentiation
shows the antagonism between mix and brachyury. In addition, goosecoid is activated by mix and of Embryonic
in turn repressed brachyury expression. Stem Cells

1D.1.7
Current Protocols in Stem Cell Biology Supplement 1
correlate with mesendoderm differentiation. In addition to this, each factor antagonizes the
the absence of a maternal T-box TF, expres- other’s activities and so inactivates the alter-
sion of a T-box TF could be induced by FGF native pathway. Examples of this can be found
as seen for Brachyury and FGF4 (eFGF) in the in lineage decisions in the blood. Here, GATA-
Xenopus mesendoderm network (Isaacs et al., 1 drives the erythroid/megakaryocyte program
1994; Casey et al., 1998; see below). Thus, and the Ets factor, Pu.1, drives other myeloid
in the absence of a maternal T-box transcrip- outcomes (Nerlov et al., 2000; Zhang et al.,
tion factor with the full range of activities in 2000; Swiers et al., 2006). This switch can
ESCs, FGF signaling may play a greater role be modified by altering the levels of the mas-
in establishing the mesoderm and the endo- ter regulators (Galloway et al., 2005; Rhodes
derm. As discussed below, nodal-related sig- et al., 2005).
naling is often downstream of VegT and the The closest motif with characteristics of
nodal-related family members share many of cross-antagonism in the Xenopus mesendo-
the targets of VegT. Therefore, another possi- derm network involves Mix.1 and Brachyury.
ble substitute for a maternal T-box determi- These TFs repress each other’s expression and
nant could be nodal signaling itself. Finally, in likely in part underpin the choice between en-
the case of maternal β-catenin, whose nuclear doderm and mesoderm from the mesendoder-
localization is driven by maternal Wnt11, the mal layer (Lemaire et al., 1998). Although
evidence suggests that this pathway is intact in there is no evidence for direct autoregula-
ESCs (Lindsley et al., 2006). tion of Mix.1 or Brachyury, the autoregula-
tory loop between Brachyury and FGF4 pro-
Autoregulation vides this function. The apparent absence of
An important consideration when driving autoregulation on the Mix.1/endoderm path-
a cell along a specific differentiation pathway way may be compensated for by the role
is how the cell will maintain the new pathway of Goosecoid, which throws the switch to-
once the driving stimulus has been removed. In wards Mix.1/endoderm, because it represses
addition to the example of FGF and Brachyury Brachyury and is itself downstream of Mix.1
mentioned above, the key initial driver of the (see Fig. 1D.1.1D; Latinkic et al., 1997; Ger-
Xenopus mesendoderm network, VegT, drives main et al., 2000). This relationship between
expression of the nodals, of which Xnr1, 2, Goosecoid and Brachyury likely also un-
and 4 can positively autoregulate (Osada et al., derpins the anteroposterior patterning of the
2000; Cha et al., 2006). Of the 63 feed-forward mesoderm, with continued Brachyury expres-
motifs identified thus far, 27 involve a nodal sion being restricted to the posterior mesoderm
family member in the Y position of the motif by Goosecoid repression. Mix.1 favors endo-
(see Fig. 1D.1.1B,C). Thus, the initial activat- derm differentiation by driving expression of
ing push that is provided by VegT is gradually GATA factors and Sox17. Both factors drive
replaced by nodal signaling, and this signal has expression of endodermal genes throughout
the ability to maintain its expression and the the future endoderm in Xenopus and zebrafish,
pathway remains active as VegT is depleted. and, later in the anterior of the embryo. Sox17
Thus, autoregulation is another key motif in also regulates Goosecoid positively, thereby
genetic regulatory networks contributing to reinforcing the endodermal lineage decision
cell memory and forward momentum. Buffer- there (Latinkic and Smith, 1999; Sinner et al.,
ing of positive autoregulation is provided by 2004, 2006; see below).
negative regulators, such as antivin/lefty, itself Both Mix.1 and Brachyury are downstream
a target of nodal signaling (Cha et al., 2006). of the same nodal-related signaling factors and
are initially found co-expressing in the same
Cross-antagonism cells (Lemaire et al., 1998), suggesting that
An important concept in differentiation, in cells are initially programmed with the po-
addition to activation of specific programs, is tential to differentiate into either germ layer
the shutting down of alternative pathways. A and subsequent downstream interactions se-
control motif unique to cells that have to make lect one germ layer or another. As already dis-
lineage decisions is cross-antagonism. In the cussed, the choice of germ layer is determined
most complete examples, this motif has two by the combination or concentration of the in-
master regulators each trying to drive differ- ducing signals. Thus, individual cells within
ent programs, with each regulator positively the mesendoderm may have indistinguishable
autoregulating its own expression, providing molecular phenotypes reflecting the broad po-
Germ Layer tential of these cells.
Induction in ESC a driving force for that particular program. In

1D.1.8
Supplement 1 Current Protocols in Stem Cell Biology
Network states ntl (no tail, the zebrafish Brachyury homolog)
Microarray based profiling of gene expres- and spt (tbx16) or zygotic oep reveal a role for
sion in embryos is revealing new interactions Brachyury in trunk somite and blood forma-
important for endoderm and mesoderm spec- tion in zebrafish (Schier et al., 1997; Amacher
ification, and these will need to be incorpo- et al., 2002).
rated into the GRN (Dickinson et al., 2006; Brachyury exerts its effects on mesoder-
Sinner et al., 2006). However, it is already mal cells both through regulating cell fate,
clear that GRNs are four dimensional, with as shown by down regulation of target genes
levels of individual transcription factors vary- when Brachyury activity is decreased or ab-
ing in both space and time between, and pos- sent, and through regulating cell movements
sibly also within, fields of cells. In vivo the during gastrulation, because cells do not move
network of an individual differentiating cell to their correct positions in brachyury mu-
will pass through various intermediate states tant embryos. Research over the last decade
and there may be a limited number of ways or so has also identified a handful of direct
of constructing the stable network responsible targets of Brachyury in Xenopus, including
for the phenotype of a specific cell type. A bet- Xwnt11, Vent2b, and FGF4 that mediate both
ter understanding of these phenomena in de- cell fate choices and cell movement (Casey
veloping embryos will likely enable enhanced et al., 1998; Tada et al., 1998; Tada and Smith,
control over ESC differentiation. 2000; Messenger et al., 2005). For instance,
Wnt11, acting through the noncanonical Wnt
Direct Target Genes pathway, regulates the convergent extension
The activity of early regulatory transcrip- movements of gastrulation in both Xenopus
tion factors must be translated into down- and zebrafish (Heisenberg et al., 2000; Tada
stream responses, such as the cell movements and Smith, 2000). Vent2b, on the other hand,
of gastrulation and the onset of cell differenti- is a transcriptional repressor that plays a role
ation. This is achieved through regulating the in BMP-mediated specification of ventral and
expression of downstream targets, which may paraxial mesoderm, at least in part through its
themselves be regulatory factors or signals giv- ability to repress goosecoid expression (Lad-
ing rise to GRNs as discussed above, or may be her et al., 1996; Onichtchouk et al., 1996;
direct effectors of cell movement or differen- Schmidt et al., 1996; Trindade et al., 1999;
tiation. Below the authors discuss some direct Melby et al., 2000). Brachyury binds the pro-
targets and the roles they play in mediating moter of Xvent2b in vivo, and probably does
the activities of a few of these early regulatory so in combination with Smad1, a transducer
transcription factors in mesendoderm. of BMP signals, in ventral lateral mesoderm
(Messenger et al., 2005).
Mesoderm FGFs are involved both in regulating gene
expression and cell movement. Inhibition of
Direct targets of Brachyury FGF signaling causes loss of posterior somites
Brachyury is expressed throughout the and notochord, and defects in gastrulation
mesoderm of early vertebrate embryos as an movements in Xenopus and zebrafish (Amaya
immediate response to mesoderm induction, et al., 1993; Griffin et al., 1995). Similarly,
and it is required for many aspects of meso- FGF signaling is required for normal meso-
derm formation. Inhibition of Brachyury func- derm formation in mouse embryos (Deng
tion and mutations in the gene reveal that its et al., 1994; Yamaguchi et al., 1994). Sev-
orthologs in mouse, frog, and zebrafish are eral different FGFs are expressed in the meso-
required during gastrulation for normal mor- derm of Xenopus (fgf3, fgf4 [efgf], and fgf8b)
phogenetic cell movements, for notochord and and zebrafish (fgf8, fgf17β, and fgf24) and
posterior mesoderm formation, and for the es- act in combination to regulate gene activ-
tablishment of left-right asymmetry (Smith, ity which patterns and maintains mesoderm
2001; Amack and Yost, 2004). Moreover, (Furthauer et al., 1997; Fisher et al., 2002;
ectopic activation of Xenopus Brachyury in Draper et al., 2003; Fletcher et al., 2006b). For
Xenopus ectodermal explants (animal caps) is instance, FGF4 and Brachyury are involved
sufficient to induce ventral and lateral meso- in a positive autoregulatory feedback loop
dermal cell fates in those cells (Cunliffe and (Isaacs et al., 1994; Casey et al., 1998), and
Smith, 1992). Double mutants uncover addi- Germ Layer
both FGF4 and FGF8b regulate MyoD expres- Induction/
tional roles for Brachyury that are not evident sion (Fisher et al., 2002; Fletcher et al., 2006b). Differentiation
in the single mutant due to redundancy with Inhibition of FGF signaling also causes de- of Embryonic
other factors. For instance double mutants in Stem Cells
fects in cell movement during gastrulation, and
1D.1.9
Current Protocols in Stem Cell Biology Supplement 1
evidence in chick embryos suggests that tional fox-related factors, the monorail muta-
different FGFs can act as chemoattractants tion not being a complete null, and/or other
or chemorepellents to ensure that migrating factors taking over the role of foxa2 in mutant
mesodermal cells arrive at their correct desti- zebrafish. Inhibiting HNF1β activity in Xeno-
nation during gastrulation (Yang et al., 2002). pus using an engrailed repressor construct in-
Interestingly, target genes of Brachyury hibits mesoderm induction by vegetal explants
that directly mediate notochord differentia- and causes defects in mesoderm formation if
tion have yet to be isolated. Furthermore, localized to the marginal zone, while inhibition
as mentioned above, during the early stages or augmentation of HNF1β activity in Xeno-
of mesendoderm formation, Brachyury is ex- pus, using mutated human alleles, leads to de-
pressed in cells that will eventually become fects in pronephros formation (Vignali et al.,
endoderm, although expression resolves into 2000; Wild et al., 2000). Hence, although these
mesodermal cells as development proceeds downstream targets play some role in aspects
(Rodaway et al., 1999; Wardle and Smith, of mesendoderm formation, it is not clear from
2004). Similarly Brachyury-expressing cells these experiments which aspects of Sox17 ac-
in embryoid bodies have the potential to go on tivity they are involved in, and it is evident that
and form both mesoderm and endodermal lin- more targets remain to be isolated.
eages (Fehling et al., 2003; Kubo et al., 2004).
Given this, it is possible that Brachyury also Targets of GATA factors
regulates the expression of early endodermal In Xenopus and zebrafish GATA 4 to 6 me-
genes, although targets are yet to be identified. diate endoderm formation. Ectopic expression
of GATA 4, 5, or 6 in Xenopus animal caps
Endoderm induces endodermal markers, such as sox17β
Direct targets of Sox17 and HNF1β, while knock down of GATA 5 and
In Xenopus, sox17 alleles (there are three) 6 activity in the whole embryo using antisense
are required in combination for normal endo- morpholinos leads to defects in gut morphol-
derm formation, while in zebrafish, Sox17 and ogy (Weber et al., 2000; Afouda et al., 2005).
the related factor, Sox32/Casanova, both play In zebrafish, mutations in gata5 (faust) lead to
a role in endoderm formation, with Casanova a decrease in early endodermal markers and
acting upstream of sox17. Inhibition of all subsequent defects in gut morphogenesis (Re-
Sox17 activity in Xenopus leads to down- iter et al., 1999, 2001).
regulation of endodermal markers, abnormal Little is known about direct targets of
gut formation, and inhibition of gastrulation GATA factors, although experiments in which
movements (Clements and Woodland, 2000). protein synthesis was inhibited suggest that
Similarly, mouse embryos null for sox17 have GATA6, and to some extent GATA5, is able
defects in gut formation (Kanai-Azuma et al., to directly activate expression of sox17β and
2002). HNF1β in Xenopus (Afouda et al., 2005).
Direct targets of Sox17 in the gastrula GATA factors are also involved in migration
stage Xenopus embryo include endodermin, of the leading edge mesendoderm across the
HNF1β, foxa1, and foxa2 which are expressed blastocoel roof during gastrulation (Fletcher
in the endoderm of early embryos (Clements et al., 2006a), although the direct targets that
and Woodland, 2003; Ahmed et al., 2004; mediate this activity are not known.
Sinner et al., 2004). foxa2 is also expressed
in the floorplate of the neural tube, and mu- Identification of direct targets
tants in zebrafish foxa2 (monorail) show a Clearly, in order to better understand how
defect in floor plate differentiation (Norton the actions of regulatory transcription factors
et al., 2005). On the other hand, inhibition of induced in response to mesendoderm-inducing
Xenopus Foxa2 in ventral tissues using an en- signals are translated into cell movements and
grailed repressor construct causes expression the onset of cell differentiation, it is necessary
of dorsoanterior mesendodermal markers and to have a better understanding of the direct
some decrease in anterior structures, while ec- targets that mediate those processes, and par-
topic expression of foxa2 in Xenopus inhibits ticularly those targets that directly affect those
the expression of mesoendodermal markers processes.
and causes severe defects in gastrulation (Suri One method of identifying direct targets is
et al., 2004). The difference in phenotypes by a candidate approach. For instance, sub-
in Xenopus and zebrafish may be due to the tractive screens or, more recently, microar-
Germ Layer
Induction in ESC engrailed repressor construct inhibiting addi- ray experiments can be used to identify genes

1D.1.10
Supplement 1 Current Protocols in Stem Cell Biology
whose expression is altered in response to ac- GERM LAYER INDUCTION
tivation or inhibition of a transcription factor DURING EMBRYONIC STEM CELL
(Saka et al., 2000; Taverner et al., 2005). Us- DIFFERENTIATION
ing protein synthesis inhibitors, such as cyclo-
heximide, one can then identify whether these
Developmental Potential of ESCs
ESCs are pluripotent immortal cells derived
genes are directly regulated transcriptional tar-
from the inner cell mass of the preimplantation
gets of that factor (Taverner et al., 2005).
mammalian blastocyst (Evans and Kaufman,
Another, more high-throughput method of
1981; Martin, 1981). The clearest demonstra-
identifying directly bound targets of a partic-
tion that ESCs are capable of generating all
ular transcription factor, or other DNA bind-
of the cell types found in the adult comes
ing protein, in embryos or ESCs is chromatin
from experiments in which normal fertile mice
immunoprecipitation combined with genomic
composed entirely of ESC-derived cells were
microarrays (Taverner et al., 2004; Boyer et al.,
generated by aggregating clusters of ESCs
2005; Wardle et al., 2006). In this method, of-
with tetraploid embryos (Nagy et al., 1993).
ten known as ChIP-chip or location analysis,
In these experiments, instructive signals from
cells or embryos are treated with formalde-
the tetraploid embryo–derived extraembryonic
hyde to cross-link the proteins to DNA. The
tissues initiated a cascade of developmental
DNA is then fragmented (ranging in size from
programs within the ESC aggregates, culmi-
∼0.2 to 2 kb) and antibodies to the factor
nating in the formation of a complete ani-
of interest are used to immunoprecipitate it
mal. Similarly, the application of exogenous
and identify the DNA bound to it in vivo.
instructive signals to aggregates of ESCs in
The isolated DNA is then amplified, labeled,
vitro also initiates a cascade of differentiation
and hybridized to a microarray containing ge-
resulting in the generation of cells representing
nomic promoter sequences. Additional meth-
multiple tissue types.
ods to identify genomic sequences directly
bound by a factor that have been used in other
systems such as Drosophila embryos, C. ele- Developmental Congruence and
gans, or mammalian cell lines include Dam-ID Directed Differentiation
(van Steensel et al., 2001; Greil et al., 2006) The factors governing establishment of the
and ChIP-cloning (Weinmann et al., 2001; Oh embryonic germ layers, ectoderm, mesoderm,
et al., 2006) or ChIP coupled with pair-end and endoderm, have been determined from
ditag sequencing (Wei et al., 2006). Cloning embryological studies, predominantly using
and sequencing the isolated fragments has the model systems such as zebrafish and Xeno-
advantage of identifying all genomic regions pus (Kimelman and Griffin, 2000; Loose and
that are associated with the factor, whereas Patient, 2004; Tam et al., 2006). These fac-
currently genomic microarrays for vertebrate tors form part of an evolutionarily conserved
organisms contain only a subset of the genome genetic regulatory network that coordinates
such as sequences around the gene promoter, gene expression, cell movement, and differen-
although this is likely to change as new tech- tiation. Within these networks, specific genes
nologies allow larger whole genomes to be mark or regulate sequential embryonic stages
studied. as cells pass through a series of progressive de-
These powerful direct genomic binding ap- velopmental restrictions culminating in their
proaches will eventually lead to the identifi- irreversible commitment to the germ layers,
cation of all the direct targets of each of the ectoderm, endoderm, and mesoderm. These
regulatory TFs involved in germ layer forma- same steps can be observed during the early
tion. These data will enhance the GRNs de- phases of ESC differentiation, with cells se-
scribed above and allow for a more complete quentially expressing genes representing in-
understanding of the control circuits of dif- ner cell mass, epiblast, primitive streak, meso-
ferentiation. They will in addition eventually derm, and endoderm (Hirst et al., 2006). Like-
identify the end game of each differentiation wise, the commitment of ESC differentiation
pathway with a complete readout of which reg- to a particular developmental program can be
ulatory TFs drive each of the terminal differ- precipitated by the same extracellular factors
entiation products. Together with the informa- found to be important for the execution of
tion on the embryonic signals that control the corresponding programs within the embryo
Germ Layer
activities of these regulatory TFs, the ability (see Fig. 1D.1.2). This correspondence be- Induction/
to control differentiation will be substantively tween what happens in the embryo and what Differentiation
improved. happens during ESC differentiation in vitro of Embryonic
Stem Cells

1D.1.11
Current Protocols in Stem Cell Biology Supplement 1
Figure 1D.1.2 Schematic representation of key steps and factors involved in the specification of the germ
layers from differentiating ESCs. BMP and Activin-like signals are required for expansion of cell numbers as
cells traverse the in vitro equivalent of an epiblast stage (Mishina et al., 1995; Song et al., 1999), marked by ex-
pression of FGF5. In the absence of BMP, Wnt, or Nodal signals, epiblast cells adopt a default neurectodermal
fate marked by expression of genes including Sox1 or Pax6. In the presence of BMP, Nodal, or Wnt signals,
cells commit to the in vitro equivalent of primitive streak formation, marked by expression of genes such as
brachyury, Mixl1, and FoxA2. FGF signals may play a role in both the neurectodermal and mesendodermal
differentiation pathways (Dell’Era et al., 2003; Sun et al., 1999b; Ying et al., 2003), although this has not yet
been assessed in conditions free from extraneous influences. Once mesendoderm formation is initiated, the
continued presence of a robust Activin-like signal promotes the formation of anterior-dorsal mesoderm (noto-
chord, somites, cardiac) and definitive endoderm. Whereas, high levels of BMP activity favor the generation
of posterior-ventral hematopoietic mesoderm. The overlapping triangles representing Activin, Wnt, and BMP
activity serve as a reminder that all of these signaling pathways must be intact for correct patterning of the
emerging mesendoderm. Activin* signifies that the identity of the ligand responsible for the activin-like signal
at this stage is unclear.

represents a developmental congruence be- perimenter more of the cell type they desire
tween the two systems. The existence of de- and a lower proportion of unwanted cell types.
velopmental congruence allows studies of em- Taken together, the ideas of developmental
bryo development in vivo to be used as a refer- congruence and directed differentiation pro-
ence for understanding and manipulating ESC vide a conceptual framework for the develop-
differentiation in vitro. ment of ESC differentiation protocols aimed at
A second concept underlying the manip- producing large numbers of specific cell types.
ulation of ESC differentiation systems is en-
capsulated in the phrase “directed differenti- Neural induction
Germ Layer ation.” This term has been used to describe A critical idea surrounding the process of
Induction in ESC ESC differentiation protocols that give the ex- germ layer formation is the concept of the
1D.1.12
Supplement 1 Current Protocols in Stem Cell Biology
default differentiation pathway; the commit- though the molecular mechanism by which RA
ment of cells to a particular lineage in the ab- exerts this affect has not been determined, it is
sence of signals to the contrary. This idea had possible that RA’s ability to induce expression
its origins in studies of Xenopus development of the Wnt antagonist, dkk1, may be impor-
in which it was demonstrated that ectoder- tant for RA-dependent neural differentiation
mal cells deprived of the opportunity for cell- (Verani et al., 2006). Finally, it is worth not-
cell interactions adopted a neural fate (Grunz ing that once neurectodermal fate has been es-
and Tacke, 1989; Green, 1994). Subsequent tablished, the same pathways whose signaling
work demonstrated that a blockade of BMP were blocked in order to achieve neural induc-
signaling or of the activity of Brachyury, a tion in the first instance are used to pattern the
critical target gene of BMP/Nodal/Wnt sig- nascent neurectodermal to specific cell fates
naling, was sufficient to permit neural dif- (Okada et al., 2004; Chiba et al., 2005; Irioka
ferentiation of naive ectoderm (Rao, 1994; et al., 2005; Kawaguchi et al., 2005; Watanabe
Hawley et al., 1995; Grunz, 1996; Hansen et al., 2005; Su et al., 2006). The use of the
et al., 1997). These studies were possible be- same signaling pathways to induce and then
cause cells of the Xenopus embryo survived in subsequently pattern cells of the nascent germ
a medium devoid of constituents with induc- layers is a theme common to both neurecto-
tive activities. Application of this principle ap- derm and mesendoderm formation.
plied to ESC differentiation required the devel-
opment of a medium capable of supporting cell Mesendoderm induction
growth and survival without imposing a prede- Gene targeting studies have demonstrated
termined differentiation outcome (Johansson that the secreted factors BMP4 (Winnier et al.,
and Wiles, 1995; Wiles and Johansson, 1997). 1995), Nodal (Conlon et al., 1994), Wnt3 (Liu
Experiments performed using such a chemi- et al., 1999), and FGF8 (Sun et al., 1999b)
cally defined media (CDM) showed that in the are required for primitive streak formation in
absence of serum, BMPs, or Activin A, dif- the mouse. Studies using chimeric embryos, in
ferentiating ESCs up-regulated expression of which the expression of these genes is absent
the neural marker Pax6, suggesting cells had from either the embryonic or extraembryonic
embarked on a neural differentiation pathway region, point to distinct roles for these proteins
(Wiles and Johansson, 1997). These and sub- during germ layer formation and patterning.
sequent studies demonstrated that BMP, Wnt, Analysis of chimeras of BMP4 null embryos
and Nodal signaling are able to block neu- and wild-type ESCs showed that BMP4 ex-
ral differentiation of ESCs, mirroring findings pression within extraembryonic ectoderm is
from studies of zebrafish and Xenopus devel- required for the initiation of gastrulation (Fu-
opment (Finley et al., 1999; Wiles and Jo- jiwara et al., 2001). Conversely, Nodal expres-
hansson, 1999; Kimelman and Griffin, 2000; sion within the epiblast is required for gastru-
Ikeda et al., 2005; Watanabe et al., 2005). That lation while its expression in extraembryonic
neurectoderm formation does indeed represent endoderm regulates induction and patterning
a default differentiation pathway for ESCs has of anterior neural tissues (Varlet et al., 1997).
since been confirmed using serum-free low- Recent work suggests that BMP4, Nodal, and
density embryoid body (EB) and monolayer Wnt3 form a regulatory loop whereby Nodal
culture differentiation systems where cell-cell signals from the epiblast maintain expression
interactions are minimized (Tropepe et al., of BMP4 in the extraembryonic ectoderm,
2001; Ying et al., 2003; Smukler et al., 2006). which in turn induces expression of Wnt3 in
Although it is unclear what factors in proximal epiblast (Ben-Haim et al., 2006).
serum are responsible for blocking neural Intriguingly, these studies also showed that
differentiation (and inducing mesendoderm Nodal is unable to directly induce expression
differentiation), the fact that the neural default of either Brachyury or Wnt3, implying Nodal
can be restored by BMP, Wnt, or Nodal an- acts indirectly through up-regulation of ex-
tagonists suggests that serum inhibits neural traembryonic BMP4, which in turn activates
development by inducing secretion of proteins embryonic Wnt3 expression. Wnt3 maintains
that activate these signaling pathways (Aubert Nodal expression in the epiblast thus resulting
et al., 2002; Gratsch and O’Shea, 2002; Pera in the establishment of a positive reinforcing
et al., 2004; Watanabe et al., 2005). Addition feedback loop (Ben-Haim et al., 2006).
Germ Layer
of retinoic acid (RA) during the early phases Paralleling findings from studies in the Induction/
of ESC differentiation can also promote neu- embryo, experiments using ESCs differenti- Differentiation
ral differentiation (Meyer et al., 2004; Okada ated in vitro suggest that BMP, Activin, and of Embryonic
Stem Cells
et al., 2004; Chiba et al., 2005). Again, al- Wnt signals are all capable of initiating the
1D.1.13
Current Protocols in Stem Cell Biology Supplement 1
formation of mesendoderm (Johansson and bryo has already mapped out differentiation
Wiles, 1995; Kubo et al., 2004; Park et al., pathways for the generation of every known
2004; D’Amour et al., 2005; Ng et al., 2005; cell type, it would seem prudent to begin with
Tada et al., 2005; Yasunaga et al., 2005; Gadue nature’s road map. In this regard, the most de-
et al., 2006; Lindsley et al., 2006). By and tailed maps are those generated from the study
large, findings from ESC differentiation stud- of frogs and fish, organisms that continue to
ies can be rationalized in the context of de- provide new insights into the molecular mech-
velopmental congruence between the in vitro anisms underlying vertebrate development.
and in vivo systems. For example, a number
of reports have now shown that sustained Ac- LITERATURE CITED
tivin signaling promotes formation of defini- Afouda, B.A., Ciau-Uitz, A., and Patient, R. 2005.
tive endoderm (DE) from ESCs (Kubo et al., GATA4, 5 and 6 mediate TGFβ maintenance
2004; D’Amour et al., 2005; Tada et al., of endodermal gene expression in Xenopus em-
bryos. Development 132:763-774.
2005; Yasunaga et al., 2005; Gadue et al.,
2006), consistent with the requirement for Agius, E., Oelgeschlager, M., Wessely, O., Kemp,
C., and De Robertis, E.M. 2000. Endodermal
Nodal signaling in gut endoderm formation
Nodal-related signals and mesoderm induction
and with lineage tracing experiments show- in Xenopus. Development 127:1173-1183.
ing that DE arises from the anterior primitive
Ahmed, N., Howard, L., and Woodland, HR. 2004.
streak, a region of robust Nodal expression Early endodermal expression of the Xenopus
(Lowe et al., 2001; Lawson and Schoenwolf, Endodermin gene is driven by regulatory
2003; Tam et al., 2003). Similarly, BMP4 pro- sequences containing essential Sox protein-
motes the formation of hematopoietic meso- binding elements. Differentiation 72:171-184.
derm (Johansson and Wiles, 1995; Wiles and Albano, R.M., Godsave, S.F., Huylebroeck, D.,
Johansson, 1997; Li et al., 2001; Park et al., Van Nimmen, K., Isaacs, H.V., Slack, J.M., and
Smith, J.C. 1990. A mesoderm-inducing fac-
2004; Ng et al., 2005) and cardiac mesoderm tor produced by WEHI-3 murine myelomono-
(Honda et al., 2006; Hosseinkhani et al., 2007), cytic leukemia cells is activin A. Development
the latter in a concentration-dependent man- 110:435-443.
ner, reflecting findings from studies in Xeno- Amacher, S.L., Draper, B.W., Summers, B.R., and
pus and zebrafish documenting the role of Kimmel, C.B. 2002. The zebrafish T-box genes
BMP4 as a morphogen (Dosch et al., 1997; no tail and spadetail are required for develop-
Neave et al., 1997). Finally, it has recently been ment of trunk and tail mesoderm and medial
floor plate. Development 129:3311-3323.
established that Nodal and Wnt signaling are
both required for induction of primitive streak Amack, J.D. and Yost, H.J. 2004. The T box tran-
scription factor no tail in ciliated cells con-
genes during ESC differentiation (Gadue et al., trols zebrafish left-right asymmetry. Curr. Biol.
2006), providing an in vitro correlate of ge- 14:685-690.
netic ablation studies in the developing mouse Amaya, E., Stein, P.A., Musci, T.J., and Kirschner,
embryo (Conlon et al., 1994; Liu et al., 1999). M.W. 1993. FGF signalling in the early speci-
fication of mesoderm in Xenopus. Development
CONCLUSIONS 118:477-487.
Although the similarities between ESC dif- Asashima, M., Nakano, H., Shimada, K., Kinoshita,
ferentiation and early mammalian develop- K., Ishii, K., Shibai, H., and Ueno, N. 1990.
ment are compelling, one must also bear in Mesodermal induction in early amphibian em-
mind that there is no a priori reason why bryos by activin A (erythroid differentiation
cells must behave according to embryological factor). Roux’s Arch. Dev. Biol. 198:330-335.
principles. Indeed, the ability to maintain im- Aubert, J., Dunstan, H., Chambers, I., and Smith, A.
mortal, pluripotent ESCs is a poignant re- 2002. Functional gene screening in embryonic
stem cells implicates Wnt antagonism in neural
minder that in vitro cultures can lead to differentiation. Nat. Biotechnol. 20:1240-1245.
the generation of nonphysiological cell types.
Babu, M.M., Luscombe, N.M., Aravind, L.,
Moreover, the outcome desired by the embryo Gerstein, M., and Teichmann, S.A. 2004. Struc-
is not the same as that sought by the experi- ture and evolution of transcriptional regula-
menter. Whereas the aim of embryogenesis is tory networks. Curr. Opin. Struct. Biol. 14:283-
to produce an animal, the aim of directed dif- 291.
ferentiation protocols is often to generate pure Ben-Haim, N., Lu, C., Guzman-Ayala, M.,
populations of a single cell type. Thus it is pos- Pescatore, L., Mesnard, D., Bischofberger, M.,
sible that many of the considerations that con- Naef, F., Robertson, E.J., and Constam, D.B.
2006. The nodal precursor acting via activin
strain the course of cell differentiation within receptors induces mesoderm by maintaining a
Germ Layer the embryo may not necessarily apply to ESC source of its convertases and BMP4. Dev. Cell
Induction in ESC differentiation. Nevertheless, because the em- 11:313-323.

1D.1.14
Supplement 1 Current Protocols in Stem Cell Biology
Birsoy, B., Kofron, M., Schaible, K., Wylie, C., and cient differentiation of human embryonic stem
Heasman, J. 2006. Vg 1 is an essential signal- cells to definitive endoderm. Nat. Biotechnol.
ing molecule in Xenopus development. Devel- 23:1534-1541.
opment 133:15-20. Dale, L. and Slack, J.M. 1987. Fate map for the
Bjornson, C.R., Griffin, K.J., Farr, G.H., 3rd, 32-cell stage of Xenopus laevis. Development
Terashima, A., Himeda, C., Kikuchi, Y., and 99:527-551.
Kimelman, D. 2005. Eomesodermin is a local- Dale, L., Smith, J.C., and Slack, J.M. 1985. Meso-
ized maternal determinant required for endo- derm induction in Xenopus laevis: A quantita-
derm induction in zebrafish. Dev. Cell 9:523- tive study using a cell lineage label and tissue-
533. specific antibodies. J. Embryol. Exp. Morphol.
Bouwmeester, T., Kim, S., Sasai, Y., Lu, B., and 89:289-312.
De Robertis, E.M. 1996. Cerberus is a head- Dale, L., Howes, G., Price, B.M., and Smith, J.C.
inducing secreted factor expressed in the ante- 1992. Bone morphogenetic protein 4: A ven-
rior endoderm of Spemann’s organizer. Nature tralizing factor in early Xenopus development.
382:595-601. Development 115:573-585.
Boyer, L.A., Lee, T.I., Cole, M.F., Johnstone, Dale, L., Matthews, G., and Colman, A. 1993.
S.E., Levine, S.S., Zucker, J.P., Guenther, M.G., Secretion and mesoderm-inducing activity of
Kumar, R.M., Murray, H.L., Jenner, R.G., the TGF-beta related domain of Xenopus Vg1.
Gifford, D.K., Melton, D.A., Jaenisch, R., and EMBO J. 12:4471-4480.
Young, R.A. 2005. Core transcriptional regula-
tory circuitry in human embryonic stem cells. Dell’Era, P., Ronca, R., Coco, L., Nicoli, S., Metra,
Cell 122:947-956. M., and Presta, M. 2003. Fibroblast growth fac-
tor receptor-1 is essential for in vitro cardiomy-
Casey, E.S., O’Reilly, M.A., Conlon, F.L., and ocyte development. Circ. Res. 93:414-420.
Smith, J.C. 1998. The T-box transcription fac-
tor Brachyury regulates expression of eFGF Deng, C.X., Wynshaw-Boris, A., Shen, M.M.,
through binding to a non-palindromic response Daugherty, C., Ornitz, D.M., and Leder, P. 1994.
element. Development 125:3887-3894. Murine FGFR-1 is required for early postim-
plantation growth and axial organization. Genes
Cha, Y.R., Takahashi, S., and Wright, C.V. 2006. Dev. 8:3045-3057.
Cooperative non-cell and cell autonomous regu-
lation of Nodal gene expression and signaling by Dickinson, K., Leonard, J., and Baker, J.C. 2006.
Lefty/Antivin and Brachyury in Xenopus. Dev. Genomic profiling of mixer and Sox17beta tar-
Biol. 290:246-264. gets during Xenopus endoderm development.
Dev. Dyn. 235:368-381.
Chiba, S., Kurokawa, M.S., Yoshikawa, H., Ikeda,
R., Takeno, M., Tadokoro, M., Sekino, H., Dominguez, I. and Green, J.B. 2000. Dorsal down-
Hashimoto, T., and Suzuki, N. 2005. Noggin and regulation of GSK3beta by a non-Wnt-like
basic FGF were implicated in forebrain fate and mechanism is an early molecular consequence
caudal fate, respectively, of the neural tube-like of cortical rotation in early Xenopus embryos.
structures emerging in mouse ES cell culture. Development 127:861-868.
Exp. Brain Res. 163:86-99. Dosch, R., Gawantka, V., Delius, H., Blumenstock,
Clements, D. and Woodland, H.R. 2000. Changes C., and Niehrs, C. 1997. Bmp-4 acts as a mor-
in embryonic cell fate produced by expression phogen in dorsoventral mesoderm patterning in
of an endodermal transcription factor, Xsox17. Xenopus. Development 124:2325-2334.
Mech. Dev. 99:65-70. Draper, B.W., Stock, D.W., and Kimmel, C.B. 2003.
Clements, D. and Woodland, H.R. 2003. VegT in- Zebrafish fgf24 functions with fgf8 to promote
duces endoderm by a self-limiting mechanism posterior mesodermal development. Develop-
and by changing the competence of cells to re- ment 130:4639-4654.
spond to TGF-beta signals. Dev. Biol. 258:454- Dyson, S. and Gurdon, J.B. 1998. The interpretation
463. of position in a morphogen gradient as revealed
Conlon, F.L., Lyons, K.M., Takaesu, N., Barth, by occupancy of activin receptors. Cell 93:557-
K.S., Kispert, A., Herrmann, B., and Robertson, 568.
E.J. 1994. A primary requirement for nodal in Engleka, M.J. and Kessler, D.S. 2001. Siamois co-
the formation and maintenance of the primitive operates with TGFbeta signals to induce the
streak in the mouse. Development 120:1919- complete function of the Spemann-Mangold
1928. organizer. Int. J. Dev. Biol. 45:241-250.
Cooke, J. and Smith, J.C. 1987. The midblastula cell Evans, M.J. and Kaufman, M.H. 1981. Establish-
cycle transition and the character of mesoderm ment in culture of pluripotential cells from
in u.v.-induced nonaxial Xenopus development. mouse embryos. Nature 292:154-156.
Development 99:197-210.
Fan, M.J. and Sokol, S.Y. 1997. A role for Siamois
Cunliffe, V. and Smith, J.C. 1992. Ectopic meso- in Spemann organizer formation. Development
derm formation in Xenopus embryos caused by 124:2581-2589.
widespread expression of a Brachyury homo- Germ Layer
logue. Nature 358:427-430. Fehling, H.J., Lacaud, G., Kubo, A., Kennedy, Induction/
M., Robertson, S., Keller, G., and Kouskoff, Differentiation
D’Amour, K.A., Agulnick, A.D., Eliazer, S., Kelly, V. 2003. Tracking mesoderm induction and of Embryonic
O.G., Kroon, E., and Baetge, E.E. 2005. Effi- its specification to the hemangioblast during Stem Cells

1D.1.15
Current Protocols in Stem Cell Biology Supplement 1
embryonic stem cell differentiation. Develop- Green, J.B.A. and Smith, J.C. 1990. Graded changes
ment 130:4217-4227. in dose of a Xenopus activin A homologue
Finley, M.F., Devata, S., and Huettner, J.E. 1999. elicit stepwise transitions in embryonic cell fate.
BMP-4 inhibits neural differentiation of murine Nature 347:391-394.
embryonic stem cells. J. Neurobiol. 40:271-287. Green, J.B., New, H.V., and Smith, J.C. 1992.
Fisher, M.E., Isaacs, H.V., and Pownall, M.E. 2002. Responses of embryonic Xenopus cells to ac-
eFGF is required for activation of XmyoD ex- tivin and FGF are separated by multiple dose
pression in the myogenic cell lineage of Xenopus thresholds and correspond to distinct axes of the
laevis. Development 129:1307-1315. mesoderm. Cell 71:731-739.

Fletcher, G., Jones, G., Patient, R., and Snape, A. Green, J.B.A., Howes, G., Symes, K., Cooke, J.,
2006a. A role for GATA factors in Xenopus and Smith, J.C. 1990. The biological effects
gastrulation movements. Mech. Dev. 123:730- of XTC-MIF: Quantitative comparison with
745. Xenopus bFGF. Development 108:229-238.

Fletcher, R.B., Baker, J.C., and Harland, R.M. Greil, F., Moorman, C., and Steensel, B.V. 2006.
2006b. FGF8 spliceforms mediate early meso- DamID: Mapping of in vivo protein-genome in-
derm and posterior neural tissue formation in teractions using tethered DNA adenine methyl-
Xenopus. Development 133:1703-1714. transferase. Methods Enzymol. 410:342-359.

Fujiwara, T., Dunn, N.R., and Hogan, B.L. 2001. Griffin, K., Patient, R., and Holder, N. 1995.
Bone morphogenetic protein 4 in the extraem- Analysis of FGF function in normal and no
bryonic mesoderm is required for allantois de- tail zebrafish embryos reveals separate mech-
velopment and the localization and survival of anisms for formation of the trunk and the tail.
primordial germ cells in the mouse. Proc. Natl. Development 121:2983-2994.
Acad. Sci. U.S.A. 98:13739-13744. Grunz, H. 1996. Factors responsible for the es-
Furthauer, M., Thisse, C., and Thisse, B. 1997. A tablishment of the body plan in the amphibian
role for FGF-8 in the dorsoventral patterning of embryo. Int. J. Dev. Biol. 40:279-289.
the zebrafish gastrula. Development 124:4253- Grunz, H. and Tacke, L. 1989. Neural differenti-
4264. ation of Xenopus laevis ectoderm takes place
Gadue, P., Huber, T.L., Paddison, P.J., and Keller, after disaggregation and delayed reaggregation
G.M. 2006. Wnt and TGF-beta signaling are re- without inducer. Cell Differ. Dev. 28:211-217.
quired for the induction of an in vitro model Gurdon, J.B., Harger, P., Mitchell, A., and Lemaire,
of primitive streak formation using embry- P. 1994. Activin signalling and response to a
onic stem cells. Proc. Natl. Acad. Sci. U.S.A. morphogen gradient. Nature 371:487-492.
103:16806-16811. Gurdon, J.B., Mitchell, A., and Mahony, D. 1995.
Galloway, J.L., Wingert, R.A., Thisse, C., Thisse, Direct and continuous assessment by cells of
B., and Zon, L.I. 2005. Loss of gata1 but not their position in a morphogen gradient. Nature
gata2 converts erythropoiesis to myelopoiesis in 376:520-521.
zebrafish embryos. Dev. Cell 8:109-116. Gurdon, J.B., Mitchell, A., and Ryan, K. 1996. An
Germain, S., Howell, M., Esslemont, G.M., and experimental system for analyzing response to
Hill, C.S. 2000. Homeodomain and winged- a morphogen gradient. Proc. Natl. Acad. Sci.
helix transcription factors recruit activated U.S.A. 93:9334-9338.
Smads to distinct promoter elements via a Hansen, C.S., Marion, C.D., Steele, K., George, S.,
common Smad interaction motif. Genes Dev. and Smith, W.C. 1997. Direct neural induction
14:435-451. and selective inhibition of mesoderm and epider-
Glinka, A., Wu, W., Onichtchouk, D., Blumenstock, mis inducers by Xnr3. Development 124:483-
C., and Niehrs, C. 1997. Head induction by 492.
simultaneous repression of Bmp and Wnt sig- Harland, R.M. 1994. Neural induction in Xenopus.
nalling in Xenopus. Nature 389:517-519. Curr. Opin. Genet. Dev. 4:543-549.
Glinka, A., Wu, W., Delius, H., Monaghan, Hawley, S.H., Wunnenberg-Stapleton, K.,
A.P., Blumenstock, C., and Niehrs, C. 1998. Hashimoto, C., Laurent, M.N., Watabe, T.,
Dickkopf-1 is a member of a new family of se- Blumberg, B.W., and Cho, K.W. 1995. Dis-
creted proteins and functions in head induction. ruption of BMP signals in embryonic Xenopus
Nature 391:357-362. ectoderm leads to direct neural induction. Genes
Gotoh, Y., Masuyama, N., Suzuki, A., Ueno, N., Dev. 9:2923-2935.
and Nishida, E. 1995. Involvement of the MAP Heasman, J. 2006. Maternal determinants of em-
kinase cascade in Xenopus mesoderm induction. bryonic cell fate. Semin. Cell Dev. Biol. 17:93-
EMBO J. 14:2491-2498. 98.
Gratsch, T.E. and O’Shea, K.S. 2002. Noggin and
Heasman, J., Crawford, A., Goldstone, K.,
chordin have distinct activities in promoting
Garnerhamrick, P., Gumbiner, B., McCrea,
lineage commitment of mouse embryonic stem
P., Kintner, C., Noro, C.Y., and Wylie, C.
(ES) cells. Dev. Biol. 245:83-94.
1994. Overexpression of cadherins and under-
Green, J.B. 1994. Borrowing thy neighbour’s genet- expression of β-catenin inhibit dorsal meso-
Germ Layer ics: Neural induction and a Brachyury mutant in derm induction in early Xenopus embryos. Cell
Induction in ESC Xenopus. Bioessays 16:539-540. 79:791-803.

1D.1.16
Supplement 1 Current Protocols in Stem Cell Biology
Heisenberg, C.P., Tada, M., Rauch, G.J., Saude, Y., Sasai, N., Yoshimura, N., Takahashi, M., and
L., Concha, M.L., Geisler, R., Stemple, D.L., Sasai, Y. 2005. Generation of Rx+/Pax6+ neural
Smith, J.C., and Wilson, S.W. 2000. Sil- retinal precursors from embryonic stem cells.
berblick/Wnt11 mediates convergent extension Proc. Natl. Acad. Sci. U.S.A. 102:11331-11336.
movements during zebrafish gastrulation. Na- Irioka, T., Watanabe, K., Mizusawa, H., Mizuseki,
ture 405:76-81. K., and Sasai, Y. 2005. Distinct effects of caudal-
Hemmati-Brivanlou, A. and Thomsen, G.H. 1995. izing factors on regional specification of embry-
Ventral mesodermal patterning in Xenopus em- onic stem cell-derived neural precursors. Brain
bryos: Expression patterns and activities of Res. Dev. Brain Res. 154:63-70.
BMP-2 and BMP-4. Dev. Genet. 17:78-89. Isaacs, H.V., Pownall, M.E., and Slack, J.M. 1994.
Hemmati-Brivanlou, A., Kelly, O.G., and Melton, eFGF regulates Xbra expression during Xenopus
D.A. 1994. Follistatin, an antagonist of activin, gastrulation. EMBO J. 13:4469-4481.
is expressed in the Spemann organizer and dis- Johansson, B.M. and Wiles, M.V. 1995. Evidence
plays direct neuralizing activity. Cell 77:283- for involvement of activin A and bone mor-
295. phogenetic protein 4 in mammalian mesoderm
Hiiragi, T., Alarcon, V.B., Fujimori, T., Louvet- and hematopoietic development. Mol. Cell Biol.
Vallee, S., Maleszewski, M., Marikawa, Y., 15:141-151.
Maro, B., and Solter, D. 2006. Where do we Jones, C.M. and Smith, J.C. 1998. Establishment
stand now? Mouse early embryo patterning of a BMP-4 morphogen gradient by long-range
meeting in Freiburg, Germany (2005). Int. J. inhibition. Dev. Biol. 194:12-17.
Dev. Biol. 50:581-586; discussion 586-587.
Jones, C.M., Lyons, K.M., Lapan, P.M., Wright,
Hill, C.S. 2001. TGF-β signaling pathways in early C.V., and Hogan, B.L. 1992. DVR-4 (bone mor-
Xenopus development. Curr. Opin. Genet. Dev. phogenetic protein-4) as a posterior-ventralizing
11:533-540. factor in Xenopus mesoderm induction. Devel-
Hirst, C.E., Ng, E.S., Azzola, L., Voss, A. K., opment 115:639-647.
Thomas, T., Stanley, E.G., and Elefanty, A.G. Jones, C.M., Kuehn, M.R., Hogan, B.L.M., Smith,
2006. Transcriptional profiling of mouse and hu- J.C., and Wright, C.V.E. 1995. Nodal-related
man ES cells identifies SLAIN1, a novel stem cell signals induce axial mesoderm and dorsal-
gene. Dev. Biol. 293:90-103. ize mesoderm during gastrulation. Development
Holwill, S., Heasman, J., Crawley, C.R., and 121:3651-3662.
Wylie, C.C. 1987. Axis and germ line defi- Joseph, E.M. and Melton, D.A. 1997. Xnr4: A
ciencies caused by UV irradiation of Xenopus Xenopus nodal-related gene expressed in the
oocytes cultured in vitro. Development 100:735- Spemann Organizer. Dev. Biol. 184:367-372.
743.
Kanai-Azuma, M., Kanai, Y., Gad, J. M., Tajima, Y.,
Honda, M., Kurisaki, A., Ohnuma, K., Okochi, H., Taya, C., Kurohmaru, M., Sanai, Y., Yonekawa,
Hamazaki, T.S., and Asashima, M. 2006. N- H., Yazaki, K., Tam, P.P., and Hayashi, Y. 2002.
cadherin is a useful marker for the progenitor Depletion of definitive gut endoderm in Sox17-
of cardiomyocytes differentiated from mouse ES null mutant mice. Development 129:2367-2379.
cells in serum-free condition. Biochem. Biophys.
Res. Commun. 351:877-882. Kawaguchi, J., Mee, P.J., and Smith, A.G. 2005.
Osteogenic and chondrogenic differentiation of
Horb, M.E. and Thomsen, G.H. 1997. A vegetally- embryonic stem cells in response to specific
localized T-box transcription factor in Xenopus growth factors. Bone 36:758-769.
eggs specifies mesoderm and endoderm and is
essential for embryonic mesoderm formation. Kimelman, D. and Griffin, K.J. 2000. Vertebrate
Development 124:1689-1698. mesendoderm induction and patterning. Curr.
Opin. Genet. Dev. 10:350-356.
Hosseinkhani, M., Hosseinkhani, H., Khademhos-
seini, A., Bolland, F., Kobayashi, H., and Kimelman, D., Abraham, J.A., Haaparanta, T.,
Prat Gonzalez, S. 2007. Bone morphogenetic Palisi, T.M., and Kirschner, M.W. 1988. The
protein-4 enhances cardiomyocyte differenti- presence of fibroblast growth factor in the frog
ation of cynomolgus monkey ES cells in egg: its role as a natural mesoderm inducer. Sci-
Knockout Serum Replacement medium. Stem ence 242:1053-1056.
Cells 25:571-580. Kofron, M., Klein, P., Zhang, F., Houston, D.W.,
Howell, M. and Hill, C.S. 1997. XSmad2 directly Schaible, K., Wylie, C., and Heasman, J. 2001.
activates the activin-inducible, dorsal mesoderm The role of maternal axin in patterning the
gene XFKH1 in Xenopus embryos. EMBO J. Xenopus embryo. Dev. Biol. 237:183-201.
16:7411-7421. Koide, T., Hayata, T., and Cho, K.W. 2005. Xenopus
Howell, M., Itoh, F., Pierreux, C.E., Valgeirsdottir, as a model system to study transcriptional reg-
S., Itoh, S., ten Dijke, P., and Hill, C.S. 1999. ulatory networks. Proc. Natl. Acad. Sci. U.S.A.
Xenopus Smad4beta is the co-Smad component 102:4943-4948.
of developmentally regulated transcription fac- Kroll, K.L. and Amaya, E. 1996. Transgenic
tor complexes responsible for induction of early Germ Layer
Xenopus embryos from sperm nuclear trans- Induction/
mesodermal genes. Dev. Biol. 214:354-369. plantations reveal FGF signaling requirements Differentiation
Ikeda, H., Osakada, F., Watanabe, K., Mizuseki, K., during gastrulation. Development 122:3173- of Embryonic
Haraguchi, T., Miyoshi, H., Kamiya, D., Honda, 3183. Stem Cells

1D.1.17
Current Protocols in Stem Cell Biology Supplement 1
Kubo, A., Shinozaki, K., Shannon, J.M., Kouskoff, Li, F., Lu, S., Vida, L., Thomson, J.A., and Honig,
V., Kennedy, M., Woo, S., Fehling, H.J., and G.R. 2001. Bone morphogenetic protein 4 in-
Keller, G. 2004. Development of definitive en- duces efficient hematopoietic differentiation of
doderm from embryonic stem cells in culture. rhesus monkey embryonic stem cells in vitro.
Development 131:1651-1662. Blood 98:335-342.
LaBonne, C., Burke, B., and Whitman, M. 1995. Lindsley, R.C., Gill, J.G., Kyba, M., Murphy, T.L.,
Role of MAP kinase in mesoderm induction and and Murphy, K.M. 2006. Canonical Wnt sig-
axial patterning during Xenopus development. naling is required for development of embry-
Development 121:1475-1486. onic stem cell-derived mesoderm. Development
133:3787-3796.
Ladher, R., Mohun, T.J., Smith, J.C., and Snape,
A.M. 1996. Xom: A Xenopus homeobox gene Linker, C. and Stern, C.D. 2004. Neural induction
that mediates the early effects of BMP-4. requires BMP inhibition only as a late step, and
Development 122:2385-2394. involves signals other than FGF and Wnt antag-
onists. Development 131:5671-5681.
Latinkic, B.V. and Smith, J.C. 1999. Goosecoid
and mix.1 repress Brachyury expression and Liu, P., Wakamiya, M., Shea, M.J., Albrecht, U.,
are required for head formation in Xenopus. Behringer, R.R., and Bradley, A. 1999. Require-
Development 126:1769-1779. ment for Wnt3 in vertebrate axis formation. Nat.
Genet. 22:361-365.
Latinkic, B.V., Umbhauer, M., Neal, K., Lerchner,
W., Smith, J.C., and Cunliffe, V. 1997. The Loose, M. and Patient, R. 2004. A genetic regula-
Xenopus Brachyury promoter is activated by tory network for Xenopus mesendoderm forma-
FGF and low concentrations of activin and sup- tion. Dev. Biol. 271:467-478.
pressed by high concentrations of activin and by Lowe, L.A., Yamada, S., and Kuehn, M.R. 2001.
paired-type homeodomain proteins. Genes Dev. Genetic dissection of nodal function in pattern-
11:3265-3276. ing the mouse embryo. Development 128:1831-
Launay, C., Fromentoux, V., Shi, D.L., and Bou- 1843.
caut, J.C. 1996. A truncated FGF receptor Lustig, K.D., Kroll, K.L., Sun, E.E., and Kirschner,
blocks neural induction by endogenous Xenopus M.W. 1996. Expression cloning of a Xenopus
inducers. Development 122:869-880. T-related gene (Xombi) involved in mesodermal
Lawson, A. and Schoenwolf, G.C. 2003. Epiblast patterning and blastopore lip formation. Devel-
and primitive-streak origins of the endoderm opment 122:4001-4012.
in the gastrulating chick embryo. Development Mangan, S., Zaslaver, A., and Alon, U. 2003. The
130:3491-3501. coherent feedforward loop serves as a sign-
Lee, T.I., Rinaldi, N.J., Robert, F., Odom, D.T., sensitive delay element in transcription net-
Bar-Joseph, Z., Gerber, G.K., Hannett, N.M., works. J. Mol. Biol. 334:197-204.
Harbison, C.T., Thompson, C.M., Simon, I., Martin, G.R. 1981. Isolation of a pluripotent cell
Zeitlinger, J., Jennings, E.G., Murry, H.L., line from early mouse embryos cultured in
Gordon, D.B., Ren, B., Wyrick, J.J., Tagne, J.B., medium conditioned by teratocarcinoma stem
Volkert, T.L., Fraenkel, E., Gifford, D.K., and cells. Proc. Natl. Acad. Sci. U.S.A. 78:7634-
Young, R.A. 2002. Transcriptional regulatory 7638.
networks in Saccharomyces cerevisiae. Science
298:799-804. McMahon, A.P. and Moon, R.T. 1989. Ectopic ex-
pression of the proto-oncogene int-1 in Xenopus
Lemaire, P., Garrett, N., and Gurdon, J.B. 1995. embryos leads to duplication of the embryonic
Expression cloning of Siamois, a Xenopus axis. Cell 58:1075-1084.
homeobox gene expressed in dorsal vegetal cells
of blastulae and able to induce a complete sec- Melby, A.E., Beach, C., Mullins, M., and
ondary axis. Cell 81:85-94. Kimelman, D. 2000. Patterning the early ze-
brafish by the opposing actions of bozozok and
Lemaire, P., Darras, S., Caillol, D., and vox/vent. Dev. Biol. 224:275-285.
Kodjabachian, L. 1998. A role for the vegetally
expressed Xenopus gene Mix.1 in endoderm Messenger, N.J., Kabitschke, C., Andrews, R.,
formation and in the restriction of mesoderm Grimmer, D., Miguel, R.N., Blundell, T.L.,
to the marginal zone. Development 125:2371- Smith, J.C., and Wardle, F.C. 2005. Functional
2380. specificity of the Xenopus T-domain protein
brachyury is conferred by its ability to interact
Lerchner, W., Latinkic, B.V., Remacle, J.E., with smad1. Dev. Cell 8:599-610.
Huylebroeck, D., and Smith, J.C. 2000. Region-
specific activation of the Xenopus Brachyury Meyer, J.S., Katz, M.L., Maruniak, J.A., and Kirk,
promoter involves active repression in ectoderm M.D. 2004. Neural differentiation of mouse em-
and endoderm: A study using transgenic frog bryonic stem cells in vitro and after transplanta-
embryos. Development 127:2729-2739. tion into eyes of mutant mice with rapid retinal
degeneration. Brain Res. 1014:131-144.
Leyns, L., Bouwmeester, T., Kim, S.H., Piccolo,
S., and De Robertis, E.M. 1997. Frzb-1 is Milo, R., Shen-Orr, S., Itzkovitz, S., Kashtan, N.,
a secreted antagonist of Wnt signaling ex- Chklovskii, D., and Alon, U. 2002. Network
pressed in the Spemann organizer. Cell 88:747- motifs: Simple building blocks of complex net-
756. works. Science 298:824-827.
Germ Layer
Induction in ESC

1D.1.18
Supplement 1 Current Protocols in Stem Cell Biology
Mishina, Y., Suzuki, A., Ueno, N., and Behringer, Papin, C. and Smith, J.C. 2000. Gradual refinement
R.R. 1995. Bmpr encodes a type I bone morpho- of activin-induced thresholds requires protein
genetic protein receptor that is essential for gas- synthesis. Dev. Biol. 217:166-172.
trulation during mouse embryogenesis. Genes Park, C., Afrikanova, I., Chung, Y.S., Zhang, W.J.,
Dev. 9:3027-3037. Arentson, E., Fong Gh, G., Rosendahl, A., and
Nagy, A., Rossant, J., Nagy, R., Abramow- Choi, K. 2004. A hierarchical order of factors
Newerly, W., and Roder, J.C. 1993. Derivation in the generation of FLK1- and SCL-expressing
of completely cell culture-derived mice from hematopoietic and endothelial progenitors from
early-passage embryonic stem cells. Proc. Natl. embryonic stem cells. Development 131:2749-
Acad. Sci. U.S.A. 90:8424-8428. 2762.
Neave, B., Holder, N., and Patient, R. 1997. A Pera, M.F., Andrade, J., Houssami, S., Reubinoff,
graded response to BMP-4 spatially coordinates B., Trounson, A., Stanley, E.G., Ward-van Oost-
patterning of the mesoderm and ectoderm in the waard, D., and Mummery, C. 2004. Regulation
zebrafish. Mech. Dev. 62:183-195. of human embryonic stem cell differentiation
Nerlov, C., Querfurth, E., Kulessa, H., and Graf, by BMP-2 and its antagonist noggin. J. Cell Sci.
T. 2000. GATA-1 interacts with the myeloid 117:1269-1280.
PU.1 transcription factor and represses PU.1- Piccolo, S., Agius, E., Leyns, L., Bhattacharyya, S.,
dependent transcription. Blood 95:2543-2551. Grunz, H., Bouwmeester, T., and De Robertis,
Ng, E.S., Azzola, L., Sourris, K., Robb, L., Stanley, E.M. 1999. The head inducer Cerberus is a mul-
E.G., and Elefanty, A.G. 2005. The primi- tifunctional antagonist of Nodal, BMP and Wnt
tive streak gene Mixl1 is required for effi- signals. Nature 397:707-710.
cient haematopoiesis and BMP4-induced ven- Piepenburg, O., Grimmer, D., Williams, P.H., and
tral mesoderm patterning in differentiating ES Smith, J.C. 2004. Activin redux: Specifica-
cells. Development 132:873-884. tion of mesodermal pattern in Xenopus by
Nieuwkoop, P.D. 1969. The formation of mesoderm graded concentrations of endogenous activin B.
in Urodelean amphibians. I. Induction by the en- Development 131:4977-4986.
doderm. Wilhelm Roux’s Arch. EntwMech. Org. Rao, Y. 1994. Conversion of a mesodermalizing
162:341-373. molecule, the Xenopus Brachyury gene, into a
Norton, W.H., Mangoli, M., Lele, Z., Pogoda, H.M., neuralizing factor. Genes Dev. 8:939-947.
Diamond, B., Mercurio, S., Russell, C., Teraoka, Reiter, J.F., Alexander, J., Rodaway, A., Yelon, D.,
H., Stickney, H.L., Rauch, G.J., Heisenberg, Patient, R., Holder, N., and Stainier, D.Y. 1999.
C.P., Houart, C., Schilling, T.F., Frohnhoefer, Gata5 is required for the development of the
H.G., Rastegar, S., Neumann, C.J., Gardiner, heart and endoderm in zebrafish. Genes Dev.
R.M., Strahle, U., Geisler, R., Rees, M., Talbot, 13:2983-2995.
W.S., and Wilson, S.W. 2005. Monorail/Foxa2 Reiter, J.F., Kikuchi, Y., and Stainier, D.Y. 2001.
regulates floorplate differentiation and specifi- Multiple roles for Gata5 in zebrafish endoderm
cation of oligodendrocytes, serotonergic raphe formation. Development 128:125-135.
neurones and cranial motoneurones. Develop-
ment 132:645-658. Rhodes, J., Hagen, A., Hsu, K., Deng, M., Liu,
T.X., Look, A.T., and Kanki, J.P. 2005. Interplay
Oh, S.W., Mukhopadhyay, A., Dixit, B.L., Raha, T., of pu.1 and gata1 determines myelo-erythroid
Green, M.R., and Tissenbaum, H.A. 2006. Iden- progenitor cell fate in zebrafish. Dev. Cell 8:97-
tification of direct DAF-16 targets controlling 108.
longevity, metabolism and diapause by chro-
matin immunoprecipitation. Nat. Genet. 38:251- Rodaway, A. and Patient, R. 2001. Mesendoderm:
257. An ancient germ layer? Cell 105:169-172.
Okada, Y., Shimazaki, T., Sobue, G., and Rodaway, A., Takeda, H., Koshida, S., Broadbent,
Okano, H. 2004. Retinoic-acid-concentration- J., Price, B., Smith, J.C., Patient, R., and Holder,
dependent acquisition of neural cell identity dur- N. 1999. Induction of the mesendoderm in the
ing in vitro differentiation of mouse embryonic zebrafish germ ring by yolk cell-derived TGF-β
stem cells. Dev. Biol. 275:124-142. family signals and discrimination of mesoderm
and endoderm by FGF. Development 126:3067-
Onichtchouk, D., Gawantka, V., Dosch, R., Delius, 3078.
H., Hirschfeld, K., Blumenstock, C., and Niehrs,
C. 1996. The Xvent-2 homeobox gene is part of Saka, Y., Tada, M., and Smith, J.C. 2000. A screen
the BMP-4 signalling pathway controlling [cor- for targets of the Xenopus T-box gene Xbra.
rection of controling] dorsoventral patterning Mech. Dev. 93:27-39.
of Xenopus mesoderm. Development 122:3045- Sasai, Y., Lu, B., Steinbeisser, H., Geissert, D.,
3053. Gont, L.K., and De Robertis, E.M. 1994.
Osada, S.I., Saijoh, Y., Frisch, A., Yeo, C.Y., Xenopus chordin: A novel dorsalizing factor ac-
Adachi, H., Watanabe, M., Whitman, M., tivated by organizer-specific homeobox genes.
Hamada, H., and Wright, C.V. 2000. Activin/ Cell 79:779-790.
nodal responsiveness and asymmetric expres- Sasai, Y., Lu, B., Piccolo, S., and De Robertis, E.M. Germ Layer
sion of a Xenopus nodal-related gene con- 1996. Endoderm induction by the organizer- Induction/
verge on a FAST-regulated module in intron 1. secreted factors chordin and noggin in Xenopus Differentiation
Development 127:2503-2514. animal caps. EMBO J. 15:4547-4555. of Embryonic
Stem Cells

1D.1.19
Current Protocols in Stem Cell Biology Supplement 1
Scharf, S.R. and Gerhart, J.C. 1980. Determination head development in mice. Dev. Biol. 213:157-
of the dorsal-ventral axis in eggs of Xenopus 169.
laevis: Complete rescue of uv-impaired eggs by
Spemann, H. 1938. Embryonic Development
oblique orientation before first cleavage. Dev.
and Induction, 1988 (reprinted) ed: Garland
Biol. 79:181-198.
Publishing, Inc. New York and London.
Schier, A.F. 2003. Nodal signaling in vertebrate de-
Stennard, F., Carnac, G., and Gurdon, J.B. 1996.
velopment. Annu. Rev. Cell Dev. Biol. 19:589-
The Xenopus T-box gene, Antipodean, en-
621.
codes a vegetally localised maternal mRNA and
Schier, A.F., Neuhauss, S.C., Helde, K.A., Talbot, can trigger mesoderm formation. Development
W.S., and Driever, W. 1997. The one-eyed pin- 122:4179-4188.
head gene functions in mesoderm and endoderm
Su, H.L., Muguruma, K., Matsuo-Takasaki, M.,
formation in zebrafish and interacts with no tail.
Kengaku, M., Watanabe, K., and Sasai, Y. 2006.
Development 124:327-342.
Generation of cerebellar neuron precursors from
Schmidt, J.E., von Dassow, G., and Kimelman, D. embryonic stem cells. Dev. Biol. 290:287-296.
1996. Regulation of dorsal-ventral patterning:
Sudarwati, S. and Nieuwkoop, P.D. 1971. Meso-
the ventralizing effects of the novel Xenopus
derm formation in the Anuran Xenopus laevis
homeobox gene Vox. Development 122:1711-
(Daudin). Roux Arch. Dev. Biol. 166:189-204.
1721.
Sinner, D., Rankin, S., Lee, M., and Zorn, A.M. Sun, B.I., Bush, S.M., Collins-Racie, L.A.,
2004. Sox17 and beta-catenin cooperate to reg- LaVallie, E.R., DiBlasio-Smith, E.A., Wolfman,
ulate the transcription of endodermal genes. N.M., McCoy, J.M., and Sive, H.L. 1999a.
Development 131:3069-3080. derrière: a TGF-beta family member re-
quired for posterior development in Xenopus.
Sinner, D., Kirilenko, P., Rankin, S., Wei, Development 126:1467-1482.
E., Howard, L., Kofron, M., Heasman, J.,
Woodland, H.R., and Zorn, A.M. 2006. Sun, X., Meyers, E.N., Lewandoski, M., and Martin,
Global analysis of the transcriptional net- G.R. 1999b. Targeted disruption of Fgf8 causes
work controlling Xenopus endoderm formation. failure of cell migration in the gastrulating
Development 133:1955-1966. mouse embryo. Genes Dev. 13:1834-1846.

Slack, J.M., Darlington, B.G., Heath, J.K., and God- Suri, C., Haremaki, T., and Weinstein, D.C. 2004.
save, S.F. 1987. Mesoderm induction in early Inhibition of mesodermal fate by Xenopus
Xenopus embryos by heparin-binding growth HNF3beta/FoxA2. Dev. Biol. 265:90-104.
factors. Nature 326:197-200. Swiers, G., Patient, R., and Loose, M. 2006. Genetic
Smith, J.C. 2001. Making mesoderm?upstream and regulatory networks programming hematopoi-
downstream of Xbra. Int. J. Dev. Biol. 45:219- etic stem cells and erythroid lineage specifica-
224. tion. Dev. Biol. 294:525-540.

Smith, J.C. and Slack, J.M. 1983. Dorsalization Tada, M. and Smith, J.C. 2000. Xwnt11 is a target
and neural induction: Properties of the organizer of Xenopus Brachyury: Regulation of gas-
in Xenopus laevis. J. Embryol. Exp. Morphol. trulation movements via Dishevelled, but not
78:299-317. through the canonical Wnt pathway. Develop-
ment 127:2227-2238.
Smith, W.C. and Harland, R.M. 1991. Injected
Xwnt-8 RNA acts early in Xenopus embryos Tada, M., Casey, E.S., Fairclough, L., and Smith,
to promote formation of a vegetal dorsalizing J.C. 1998. Bix1, a direct target of Xenopus T-box
center. Cell 67:753-765. genes, causes formation of ventral mesoderm
and endoderm. Development 125:3997-4006.
Smith, W.C. and Harland, R.M. 1992. Expression
cloning of noggin, a new dorsalizing factor lo- Tada, S., Era, T., Furusawa, C., Sakurai, H.,
calized to the Spemann organizer in Xenopus Nishikawa, S., Kinoshita, M., Nakao, K., and
embryos. Cell 70:829-840. Chiba, T. 2005. Characterization of mesendo-
derm: A diverging point of the definitive en-
Smith, J.C., Price, B.M., Van Nimmen, K., and doderm and mesoderm in embryonic stem cell
Huylebroeck, D. 1990. Identification of a potent differentiation culture. Development 132:4363-
Xenopus mesoderm-inducing factor as a homo- 4374.
logue of activin A. Nature 345:729-731.
Takahashi, S., Yokota, C., Takano, K.,
Smukler, S.R., Runciman, S.B., Xu, S., and van der Tanegashima, K., Onuma, Y., Goto, J.,
Kooy, D. 2006. Embryonic stem cells assume a and Asashima, M. 2000. Two novel nodal-
primitive neural stem cell fate in the absence of related genes initiate early inductive events
extrinsic influences. J. Cell Biol. 172:79-90. in Xenopus Nieuwkoop center. Development
Sokol, S., Christian, J.L., Moon, R.T., and Melton, 127:5319-5329.
D.A. 1991. Injected wnt RNA induces a com- Tam, P. P., Kanai-Azuma, M., and Kanai, Y.
plete body axis in Xenopus embryos. Cell 2003. Early endoderm development in ver-
67:741-752. tebrates: Lineage differentiation and morpho-
Song, J., Oh, S.P., Schrewe, H., Nomura, M., Lei, genetic function. Curr. Opin. Genet. Dev.
H., Okano, M., Gridley, T., and Li, E. 1999. 13:393-400.
Germ Layer The type II activin receptors are essential for Tam, P.P., Loebel, D.A., and Tanaka, S.S.
Induction in ESC egg cylinder growth, gastrulation, and rostral 2006. Building the mouse gastrula: Signals,

1D.1.20
Supplement 1 Current Protocols in Stem Cell Biology
asymmetry and lineages. Curr. Opin. Genet. Vincent, J.P. and Gerhart, J.C. 1987. Subcortical
Dev. 16:419-425. rotation in Xenopus eggs: An early step in em-
Tao, Q., Yokota, C., Puck, H., Kofron, M., Birsoy, bryonic axis specification. Dev. Biol. 123:526-
B., Yan, D., Asashima, M., Wylie, C.C., Lin, X., 539.
and Heasman, J. 2005. Maternal wnt11 activates Wardle, F.C. and Smith, J.C. 2004. Refinement of
the canonical wnt signaling pathway required gene expression patterns in the early Xenopus
for axis formation in Xenopus embryos. Cell embryo. Development 131:4687-4696.
120:857-871. Wardle, F.C., Odom, D.T., Bell, G.W., Yuan, B.,
Taverner, N.V., Smith, J.C., and Wardle, F.C. 2004. Danford, T.W., Wiellette, E.L., Herbolsheimer,
Identifying transcriptional targets. Genome Biol. E., Sive, H.L., Young, R.A., and Smith, J.C.
5:210. 2006. Zebrafish promoter microarrays identify
Taverner, N.V., Kofron, M., Shin, Y., Kabitschke, actively transcribed embryonic genes. Genome
C., Gilchrist, M.J., Wylie, C., Cho, K.W., Biol. 7:R71.
Heasman, J., and Smith, J.C. 2005. Microarray- Watabe, T., Kim, S., Candia, A., Rothbacher, U.,
based identification of VegT targets in Xenopus. Hashimoto, C., Inoue, K., and Cho, K.W. 1995.
Mech. Dev. 122:333-354. Molecular mechanisms of Spemann’s organizer
Thomsen, G.H. and Melton, D.A. 1993. Processed formation: conserved growth factor synergy be-
Vg1 protein is an axial mesoderm inducer in tween Xenopus and mouse. Genes Dev. 9:3038-
Xenopus. Cell 74:433-441. 3050.
Watanabe, K., Kamiya, D., Nishiyama, A.,
Thomsen, G., Woolf, T., Whitman, M., Sokol, S.,
Katayama, T., Nozaki, S., Kawasaki, H.,
Vaughan, J., Vale, W., and Melton, D.A. 1990.
Watanabe, Y., Mizuseki, K., and Sasai, Y. 2005.
Activins are expressed early in Xenopus em-
Directed differentiation of telencephalic precur-
bryogenesis and can induce axial mesoderm and
sors from embryonic stem cells. Nat. Neurosci.
anterior structures. Cell 63:485-493.
8:288-296.
Toyoizumi, R., Ogasawara, T., Takeuchi, S., and
Weber, H., Symes, C.E., Walmsley, M.E., Rod-
Mogi, K. 2005. Xenopus nodal related-1 is in-
away, A.R., and Patient, R.K. 2000. A role
dispensable only for left-right axis determina-
for GATA5 in Xenopus endoderm specification.
tion. Int. J. Dev. Biol. 49:923-938.
Development 127:4345-4360.
Trindade, M., Tada, M., and Smith, J.C. 1999. DNA-
Weeks, D.L. and Melton, D.A. 1987. A maternal
binding specificity and embryological function
mRNA localized to the vegetal hemisphere in
of Xom (Xvent-2). Dev. Biol. 216:442-456.
Xenopus eggs codes for a growth factor related
Tropepe, V., Hitoshi, S., Sirard, C., Mak, T.W., to TGF-beta. Cell 51:861-867.
Rossant, J., and van der Kooy, D. 2001. Direct Wei, C.L., Wu, Q., Vega, V.B., Chiu, K.P., Ng,
neural fate specification from embryonic stem P., Zhang, T., Shahab, A., Yong, H.C., Fu, Y.,
cells: A primitive mammalian neural stem cell Weng, Z., Liu, Z., Zhao, X.D., Chew, J.L., Lee,
stage acquired through a default mechanism. Y.L., Kuznetsov, V.A., Sung, W.K., Miller, L.D.,
Neuron 30:65-78. Lim B, Liu, E.T., Yu, Q., Ng, H.H., and Ruan,
Umbhauer, M., Marshall, C.J., Mason, C.S., Old, Y. 2006. A global map of p53 transcription-
R.W., and Smith, J.C. 1995. Mesoderm induc- factor binding sites in the human genome. Cell
tion in Xenopus caused by activation of MAP 124:207-219.
kinase. Nature 376:58-62. Weinmann, A. S., Bartley, S. M., Zhang, T., Zhang,
van Steensel, B., Delrow, J., and Henikoff, S. M. Q., and Farnham, P. J. 2001. Use of chromatin
2001. Chromatin profiling using targeted DNA immunoprecipitation to clone novel E2F target
adenine methyltransferase. Nat. Genet. 27:304- promoters. Mol. Cell Biol. 21:6820-6832.
308. Wild, W., Pogge von Strandmann, E., Nastos,
Varlet, I., Collignon, J., and Robertson, E.J. 1997. A., Senkel, S., Lingott-Frieg, A., Bulman, M.,
Nodal expression in the primitive endoderm Bingham, C., Ellard, S., Hattersley, A.T., and
is required for specification of the anterior Ryffel, G.U. 2000. The mutated human gene
axis during mouse gastrulation. Development encoding hepatocyte nuclear factor 1β inhibits
124:1033-1044. kidney formation in developing Xenopus em-
Verani, R., Cappuccio, I., Spinsanti, P., Gradini, bryos. Proc. Natl. Acad. Sci. U.S.A. 97:4695-
R., Caruso, A., Magnotti, M.C., Motolese, M., 4700.
Nicoletti, F., and Melchiorri, D. 2006. Expres- Wiles, M.V. and Johansson, B.M. 1997. Analysis
sion of the Wnt inhibitor Dickkopf-1 is required of factors controlling primary germ layer for-
for the induction of neural markers in mouse mation and early hematopoiesis using embry-
embryonic stem cells differentiating in re- onic stem cell in vitro differentiation. Leukemia
sponse to retinoic acid. J. Neurochem. 100:242- 11:454-456.
250.
Wiles, M.V. and Johansson, B. M. 1999. Embryonic
Vignali, R., Poggi, L., Madeddu, F., and Barsacchi, stem cell development in a chemically defined
G. 2000. HNF1(beta) is required for mesoderm medium. Exp. Cell Res. 247:241-248. Germ Layer
induction in the Xenopus embryo. Development Induction/
127:1455-1465. Williams, P.H., Hagemann, A., Gonzalez-Gaitan, Differentiation
M., and Smith, J.C. 2004. Visualizing of Embryonic
Stem Cells

1D.1.21
Current Protocols in Stem Cell Biology Supplement 1
long-range movement of the morphogen Xnr2 differentiation of mouse ES cells. Nat. Biotech-
in the Xenopus embryo. Curr. Biol. 14:1916- nol. 23:1542-1550.
1923.
Ying, Q.L., Stavridis, M., Griffiths, D., Li, M.,
Winnier, G., Blessing, M., Labosky, P.A., and and Smith, A. 2003. Conversion of embryonic
Hogan, B.L. 1995. Bone morphogenetic protein- stem cells into neuroectodermal precursors in
4 is required for mesoderm formation and adherent monoculture. Nat. Biotechnol. 21:183-
patterning in the mouse. Genes Dev. 9:2105- 186.
2116.
Zernicka-Goetz, M. 2006. The first cell-fate deci-
Xanthos, J.B., Kofron, M., Wylie, C., and sions in the mouse embryo: Destiny is a matter
Heasman, J. 2001. Maternal VegT is the initiator of both chance and choice. Curr. Opin. Genet.
of a molecular network specifying endo- Dev. 16:406-412.
derm in Xenopus laevis. Development 128:167-
Zhang, J., Houston, D.W., King, M.L., Payne, C.,
180.
Wylie, C., and Heasman, J. 1998. The role of
Yamaguchi, T.P., Harpal, K., Henkemeyer, M., and maternal VegT in establishing the primary germ
Rossant, J. 1994. fgfr-1 is required for embry- layers in Xenopus embryos. Cell 94:515-524.
onic growth and mesodermal patterning dur-
Zhang, J. and King, M.L. 1996. Xenopus VegT
ing mouse gastrulation. Genes Dev. 8:3032-
RNA is localized to the vegetal cortex during
3044.
oogenesis and encodes a novel T-box transcrip-
Yang, J., Tan, C., Darken, R.S., Wilson, P.A., and tion factor involved in mesodermal patterning.
Klein, P.S. 2002. Beta-catenin/Tcf-regulated Development 122:4119-4129.
transcription prior to the midblastula transition. Zhang, P., Zhang, X., Iwama, A., Yu, C., Smith,
Development 129:5743-5752. K.A., Mueller, B.U., Narravula, S., Torbett,
Yasunaga, M., Tada, S., Torikai-Nishikawa, S., B.E., Orkin, S.H., and Tenen, D.G. 2000. PU.1
Nakano, Y., Okada, M., Jakt, L.M., Nishikawa, inhibits GATA-1 function and erythroid differ-
S., Chiba, T., and Era, T. 2005. Induction and entiation by blocking GATA-1 DNA binding.
monitoring of definitive and visceral endoderm Blood 96:2641-2648.

Germ Layer
Induction in ESC

1D.1.22
Supplement 1 Current Protocols in Stem Cell Biology

Das könnte Ihnen auch gefallen