SlideShare a Scribd company logo
1 of 15
Download to read offline
Evidence for common machinery utilized by the early and late RNA
localization pathways in Xenopus oocytes
Soheun Chooa,1
, Bianca Heinrichb,1
, J. Nicholas Betleyb
, Zhao Chenb
, James O. Deshlera,b,*
a
Molecular Biology and Biochemistry Program, Boston University, Boston, Massachusetts 02215, USA
b
Department of Biology, Boston University, 5 Cummington Street, Boston, Massachusetts 02215, USA
Received for publication 13 July 2004, revised 25 October 2004, accepted 27 October 2004
Available online 18 November 2004
Abstract
In Xenopus, an early and a late pathway exist for the selective localization of RNAs to the vegetal cortex during oogenesis. Previous work
has suggested that distinct cellular mechanisms mediate localization during these pathways. Here, we provide several independent lines of
evidence supporting the existence of common machinery for RNA localization during the early and late pathways. Data from RNA
microinjection assays show that early and late pathway RNAs compete for common localization factors in vivo, and that the same short RNA
sequence motifs are required for localization during both pathways. In addition, quantitative filter binding assays demonstrate that the late
localization factor Vg RBP/Vera binds specifically to several early pathway RNA localization elements. Finally, confocal imaging shows that
early pathway RNAs associate with a perinuclear microtubule network that connects to the mitochondrial cloud of stage I oocytes suggesting
that motor driven transport plays a role during the early pathway as it does during the late pathway. Taken together, our data indicate that
common machinery functions during the early and late pathways. Thus, RNA localization to the vegetal cortex may be a regulated process
such that differential interactions with basal factors determine when distinct RNAs are localized during oogenesis.
D 2004 Elsevier Inc. All rights reserved.
Keywords: RNA localization; Xenopus; Oocytes; Vg1 RBP; Vera; Vg1; Xcat-2; KH domain; Microtubules
Introduction
The subcellular localization of specific mRNAs to
distinct regions of a cell is an important mechanism for
generating asymmetric gene expression in germ cells and
somatic cells (Kloc et al., 2001; Palacios and St. Johnston,
2001), and the sequences that specify localization generally
reside in the 3V untranslated region (3V UTR) of localized
transcripts. In Xenopus, two temporally defined pathways
exist for localizing specific RNAs to the vegetal cortex
during oogenesis (Forristall et al., 1995; Kloc and Etkin,
1995). RNAs that utilize the early pathway, such as Xcat-2,
are distributed uniformly throughout the cytoplasm of early
stage I oocytes (50–100 Am diameter), but become localized
to the mitochondrial cloud in later stage I oocytes (120–300
Am diameter) and subsequently associate with the vegetal
cortex of stage II oocytes (Zhou and King, 1996a).
Additional examples of RNAs that use the early pathway
include Xpat (Hudson and Woodland, 1998), Xwnt-11 (Ku
and Melton, 1993), DEADSouth (MacArthur et al., 2000)
Xdazl (Houston et al., 1998), and Xlsirts (Kloc et al., 1993).
Treatment of stage I oocytes with microtubule depolymeriz-
ing drugs has little or no effect on the distribution of
endogenous early pathway RNAs associated with the
mitochondrial cloud which has led to a view that the early
pathway functions independently of microtubules (Kloc and
Etkin, 1995).
The late or Vg1 pathway begins when Xcat-2 and other
early RNAs become associated with the vegetal cortex. In
early and late stage I oocytes, Vg1 is distributed throughout
the cytoplasm and little is detected in the mitochondrial
cloud (Betley et al., 2002; Forristall et al., 1995; Kloc and
Etkin, 1995). In stage II oocytes, endogenous Vg1 becomes
0012-1606/$ - see front matter D 2004 Elsevier Inc. All rights reserved.
doi:10.1016/j.ydbio.2004.10.019
* Corresponding author. Department of Biology, Boston University, 5
Cummington Street, Boston, MA 02215. Fax: +1 617 353 8484.
E-mail address: jdeshler@bu.edu (J.O. Deshler).
1
These two authors contributed equally to this work.
Developmental Biology 278 (2005) 103–117
www.elsevier.com/locate/ydbio
associated with a wedge-shaped region of cytoplasm
between the oocyte nucleus and the vegetal pole (Kloc
and Etkin, 1995, 1998). This wedge-shaped region is
enriched in membranes of the endoplasmic reticulum
(ER), and it has been proposed that the ER may play a
role in the localization process (Deshler et al., 1997; Kloc
and Etkin, 1998). Subsequent transport of Vg1 from this
wedge-shaped region to the vegetal cortex of stage II/III
oocytes requires microtubules (Kloc and Etkin, 1995;
Yisraeli et al., 1990) and heterotrimeric kinesin II (Betley
et al., 2004). In stage III/IV oocytes, actin filaments (Kloc
and Etkin, 1995; Yisraeli et al., 1990), cytokeratin filaments
(Alarcon and Elinson, 2001), and localized RNAs them-
selves (Heasman et al., 2001; Kloc and Etkin, 1994) help to
anchor early and late RNAs close to the plasma membrane
in the vegetal cortex. In addition to these cytoskeletal
requirements, three RNA binding proteins have been
implicated in the late localization pathway. HnRNP I binds
the consensus sequence YYUCU (Y = pyrimidine) (Lewis
et al., 2004) that is required for efficient localization of two
late pathway RNAs, Vg1 and VegT (Bubunenko et al., 2002;
Cote et al., 1999), and Vg RBP/Vera (Deshler et al., 1998;
Havin et al., 1998) is an ER associated protein that binds
specifically to the sequence UUCAC which is also essential
for localization during the late pathway (Deshler et al.,
1997, 1998; Kwon et al., 2002). Recently, a double-stranded
RNA binding protein, Staufen, has also been implicated as a
factor required for localization during the late pathway
(Yoon and Mowry, 2004).
The differences in timing and possible cytoskeletal
requirements have led to a prevailing view that distinct
mechanisms are responsible for RNA localization during the
early and late pathway. However, several observations are
not entirely consistent with this model. For example, RNAs
which normally use the early pathway, such as Xcat-2 (Zhou
and King, 1996b), Xpat (Hudson and Woodland, 1998), and
Xlsirts (Allen et al., 2003), will localize to the vegetal cortex
during the late pathway if injected into stage III oocytes.
Moreover, this ectopic localization of early RNAs during
the late pathway requires intact microtubules (Zhou and
King, 1996b) and Kinesin II (Betley et al., 2004) indicating
that early RNAs are able to interact specifically with the late
pathway localization machinery in vivo. In addition, the LEs
specifying early or late localization have similar sequence
characteristics; both have an evolutionarily conserved
abundance of short CAC-containing motifs that are essential
for localization (Betley et al., 2002; Bubunenko et al., 2002;
Deshler et al., 1997, 1998; Kwon et al., 2002). These
observations suggest that the early and late pathway may
share a common set of cellular factors that are required for
RNA localization in stage I as well as in stage II/III oocytes.
Here, we provide four independent lines of evidence each
of which supports the hypothesis that the early and late
pathways utilize common cellular machinery for RNA
localization. First, we show that small motifs, which are
required for localization during the late pathway, are also
required for localization during the early pathway and vice
versa. Second, competition assays designed to detect
whether two different RNAs interact with common or
distinct localization factors in vivo show that early and late
pathway RNAs specifically compete for the same local-
ization factors in oocytes. Third, biochemical assays show
that the late localization factor Vg RBP/Vera binds
specifically to several early pathway LEs, and point
mutations that reduce Vera binding block localization during
the early pathway. Fourth, confocal imaging reveals that
early pathway RNAs associate specifically with a micro-
tubule network that surrounds the nucleus and extends into
the mitochondrial cloud of stage I oocytes suggesting that
directed transport may play a role during the early pathway
as it does during the late pathway. Together, these findings
strongly suggest that early and late pathway RNAs interact
with basal localization machinery that functions throughout
the early and late pathways.
Materials and methods
Plasmid constructs, mutants, and in vitro transcription
reactions
RNAs for microinjection were transcribed from linear-
ized plasmid DNA templates. The VgLE RNA was tran-
scribed from XbaI linearized p19.366 which was
constructed by cloning a blunt ended 366 nucleotide
BsmI–SspI fragment of the Vg1 gene into HincII digested
pTZ19U (Bio-Rad Laboratories). The DaE2 RNA was
transcribed from XbaI digested pDaE2 which is identical
to p19.366 except that the UUCAC motifs were deleted by
site directed mutagenesis as previously described (Deshler
et al., 1997). The Xcat-2 MCLE was transcribed from
HindIII digested pMCLE which was made by cloning a
PCR product containing the MCLE (nucleotides 391–630 of
the Xcat-2 gene) into SalI–HindIII digested pTZ18U
(partial MCS). This vector is a modified version of pTZ18U
(Bio-Rad Laboratories) that has the promoter proximal half
of the polylinker removed by deleting the region between
the EcoRI and XbaI sites using endonuclease digestion, an
end fill-in reaction with the Klenow fragment of DNA
polymerase, and self-ligation. The DUGCAC mutant RNA
was transcribed from pMCLEDUGCAC which is identical
to pMCLE except all six UGCAC motifs are deleted as
described previously (Betley et al., 2002). The localization
elements of DEADSouth (DSLE) and Xdaz1 (XdazLE) were
transcribed from BamHI digested pDSLE and pXdazLE,
respectively, which were constructed by cloning PCR
products containing the localization elements of DEAD-
South (DSLE) (nt 1796–2125) or Xdaz1 (XdazLE) (1335–
1780) (Betley et al., 2002) into EcoRI–BamHI digested
pTZ18U (partial MCS). The XpatLE RNA was transcribed
from SalI digested pXpatLE which was constructed by
cloning a PCR fragment containing the XpatLE (nt 1063–
S. Choo et al. / Developmental Biology 278 (2005) 103–117104
1589 of the Xpat gene) (Betley et al., 2002) into BamHI–
SalI digested pTZ18U. All constructs were verified by DNA
sequencing. The Vg9WYCAY mutant was constructed by
site directed mutagenesis using sequential PCR mediated
mutagenesis with partially overlapping oligonucleotide
amplification primers. The mutant VgLE fragment replaced
wild type fragment of p19.366 which was verified by DNA
sequencing.
Unlabeled competitor RNAs were transcribed in vitro
using 250 Units of T7 RNA polymerase (Epicentre) in 500
Al reactions containing 20 Ag of linearized plasmid DNA
template, 4 mM ATP, CTP, GTP and UTP, 20 mM MgCl2,
10 ACi 32
P-labeled UTP, 0.25 units inorganic pyrophospha-
tase (Sigma), 1 mM DTT, 40 mM Tris, pH 8.0, 2 mM
spermidine, and 200 Units of RNAse inhibitor (Promega).
These reactions were incubated at 378C for 5 h, treated with
DNAse, extracted with phenol and precipitated with 2.5 M
ammonium acetate, 2.5 volumes of ethanol, and 0.5 Ag/Al
glycogen (Roche). Digoxygenin-labeled RNAs were tran-
scribed in 50 Al reactions containing 1 Ag of plasmid
template, 25 units of T7 RNA polymerase, 0.33 mM ATP,
CTP, and GTP, 0.22 mM UTP, 110 AM Digoxigenin-11-
UTP (Roche), 8 mM MgCl2, 1 mM DTT, 10 ACi 32
P-labeled
UTP, 40 units RNAse inhibitor, 40 mM Tris, pH 8.0, and 2
mM spermidine. These reactions proceeded at 378C for 2–3
h followed by precipitation with 2.5 M ammonium acetate,
2.5 volumes of ethanol, and 30 Ag glycogen. RNAs were
quantified either by A260 or by determining the percent of
nucleotides incorporated into full-length RNAs during the
transcription reaction. The quality of all RNAs was verified
by autoradiographic analysis after running the RNAs in 6%
polyacrylamide gels containing 8 M urea.
In vivo localization and competition assays
All oocytes were isolated from Xenopus laevis (Nasco)
and staged by size according to Dumont (Dumont, 1972).
Stage I (50–300 Am diameter) and Stage II (300–450 Am
diameter) oocytes were isolated from 5 to 7 cm females as
previously described (Kloc and Etkin, 1999) except 100
Ag/ml liberase blendzyme III (Roche) was used instead of
type II collagenase. Stage IV (450–600 Am diameter)
oocytes were isolated from 6 to 8 cm albino females
(Nasco) using the same method, but with 200 Ag/ml
liberase blendzyme III. The segment of RNA that targets
Xcat-2 to the mitochondrial cloud has been referred to as the
mitochondrial cloud localization element (MCLE) and
resides in the first approximately 220 nucleotides of the
Xcat-2 3V UTR (Kloc et al., 2000; Zhou and King, 1996a).
To determine if the MCLE RNA lacking the UGCAC motifs
localizes in stages II–IV oocytes, 15 fmol of digoxygenin
labeled MCLE or DUGCAC mutant RNA (Betley et al.,
2002) was injected into each oocyte and the localization
signal was detected as previously described with developing
times of 20 min (Kloc and Etkin, 1999). Localization assays
in stage I oocytes (approximately 200 Am diameter) involved
injecting 0.1 fmol of RNA in 0.1 nl and developing the
colorometric reaction for 45 min. In these oocytes, the
mitochondrial clouds are adjacent to the germinal vesicle and
do not yet associate with the vegetal cortex. For competition
assays in early stage II oocytes (approximately 325 Am
diameter), the amount of injected digoxygenin-labeled VgLE
was titrated to determine the lower limit of reliably detecting
localization to the vegetal wedge-shaped region. This stage
of oogenesis is easily distinguished from stage I because the
mitochondrial cloud begins to contact the vegetal cortex and
forms the vegetal wedge. Localization was detected in 80–
90% of oocytes injected with 0.3 fmol of dig-VgLE, but not
with half this amount (data not shown). Varying amounts of
unlabeled MCLE, VgLE or Xenopus h-globin (XhG) were
then co-injected with dig-VgLE to determine the molar
excess required for specific competition. A 200- to 300-fold
molar excess of unlabeled VgLE RNA was found to be
required to inhibit localization in N90% of injected oocytes.
Competition of unlabeled XhG–VgLE in stage IV oocytes
was done as previously described (Kwon et al., 2002) and
localization of the XhG–VgLE RNA was detected by whole
mount in situ hybridization using digoxygenin labeled
antisense XhG RNA. The time course analysis of Fig. 4
was performed in mid-stage II oocytes (360–380 Am
diameter). Oocytes were injected with 4 fmol of dig-MCLE
or dig-VgLE in 1.5 nl and the colorometric reaction was
developed for 20 min for all oocytes examined.
Recombinant Vera and in vitro RNA binding assays
The full length Vera gene was amplified using the
polymerase chain reaction (PCR) and oligonucleotide
primers containing unique restriction sites. The PCR product
was cloned into EcoRI–BglII digested pQE60 (Qiagen)
resulting in the expression plasmid, pQEVeraHIS which
place a 6Â-His tag at the C-terminus of Vera. The se-
quences of the upstream and downstream PCR primers were
5V-AATTTCACACAGAATTCATTAAAGAGGAG-
AAATTAACCATGAACAAGCTGTATATTGGAAA-
CCTC-3V and 5V-TTTTCCTGAAAGATCTTTTT-
CTTCTCGGTTGGGGTTG-3V, respectively. One hundred
milliliters of XL-1 Blue containing pQEVeraHIS was grown
overnight in rich medium containing ampicillin (100 Ag/ml).
This culture was diluted 10 fold in the same medium and
allowed to grow at 378C to an A600 of 0.8 at which point
Vera expression was induced with 1 mM IPTG for 4 h at
378C. Following centrifugation, the cell pellet was resus-
pended in 15 mL of high salt buffer (HSB) (1 M NaCl, 50
mM sodium phosphate buffer, pH 7.8, 1 % Triton X-100, 1
mM DTT) and the cells were homogenized by running them
through a French press twice. Cell debris was cleared by
centrifugation, and the soluble fraction was stored at À708C
in 300-Al aliquots. On the day of filter binding, one aliquot
of the soluble fraction was mixed with 10 Al of HSB-
equilibrated Ni-NTA resin (Qiagen) and allowed to rotate at
48C for 1 h. The protein-bound resin was washed once each
S. Choo et al. / Developmental Biology 278 (2005) 103–117 105
with 300 Al HSB, PBS containing 1 mM DTT, and PBS
containing 1 mM DTT and 75 mM imidazole, and Vera was
eluted off the beads using 10 Al of PBS containing 1 mM
DTT, 500 mM imidazole, and 4 M urea. Vera protein
concentration was determined using the micro protein assay
(Bio-Rad) with bovine serum albumin as a standard. For
filter binding assays, 50 Al reactions containing 50 Ag/ml
yeast tRNA (Roche) and 5 mg/ml heparin, 104
cpm of 32
P-
labeled RNA probes (b0.1 nM final concentration), and
varying amounts of Vera in 1Â FBA buffer (5% glycerol,
100 mM NaCl, 50 mM Tris, pH 8.0, 1 mM MgCl2, 1 mM
DTT) were incubated at 248C for 30 min. These reactions
were then filtered through a Hybond-C membrane (Amer-
sham Biosciences) which had been equilibrated by filtering
100 Al of 1Â FBA and 50 Al of 1Â FBA containing 0.1 mg/
ml yeast tRNA per well using a dot–blot vacuum apparatus
(Bio-Rad). The membrane was washed with 200 Al of 1Â
FBA per well, and the amount of bound RNA was measured
on a Phosphorimager screen (Molecular Dynamics). Dis-
sociation constants (Kd) were determined by calculating the
amount of recombinant Vera required to bind half of the
RNA probe using Sigma Plot to curve fit the data. All Kd
values represent the average of at least five independent
measurements performed on different days and the standard
deviations are shown. To determine relative binding
affinities for Vera, the Kd for the wild type RNA was
divided by the Kd for mutant RNAs. This calculation
eliminates any error in the wild type RNA measurement
because by definition, its value becomes one. The relative
Vera binding for mutant RNAs was averaged from at least
six measurements and plotted in the histograms shown with
error bars representing the standard deviation of each value.
Since the value for the wild type is always one, there is no
error in its value, and thus no error bars are shown for the
wild type VgLE.
Confocal imaging
The localization element for Vg1 (VgLE), Xcat-2
(MCLE), or the XhG RNA was fluorescently labeled for
injections into stage I oocytes with Alexa-Fluor-546-14-
UTP (Molecular Probes) during in vitro transcription as
recommended by the manufacturer. Oocytes were injected
with the indicated fluorescent RNA and cultured in 0.5 Â
Leibovitz L-15 medium (Sigma) containing 7.5 mM
HEPES, pH 7.6, 10 mg/ml tetracycline (Sigma) and 50
mg/ml gentamycin (Roche). Oocytes were then fixed in
MEMFA (0.1 M MOPS, pH 7.5, 1 mM EGTA, 1 mM
MgSO4, 1.4% formaldehyde) or MEMFA supplemented
with 0.5 AM taxol (Paclitaxel) (Sigma) to preserve micro-
tubules. Immunocytochemistry was performed as previously
described (Betley et al., 2004) using a 1:250 dilution of the
DM1A monoclonal anti-a-tubulin antibody (Sigma) and a
1:300 dilution of fluorescein-conjugated donkey anti-mouse
secondary antibody (Jackson ImmunoResearch). Oocytes
were mounted in Vectashield (Vector Laboratories, Inc.) for
imaging. Confocal images were acquired on a Zeiss LSM
510 confocal laser scanning microscope with an Argon
Laser (488 nm) for detection of fluorescein (green) and a
Helium-Neon laser (543 nm) for detection of Alexa-Fluor
546 (red). No green fluorescence was detected in control
oocytes incubated with secondary but not primary anti-
bodies. No red fluorescence was detected in control oocytes
not injected with fluorescent RNA.
Results
The same functional motifs are required for localization
during the early and late pathway
Previous reports have shown that early pathway RNAs
will localize during the late pathway if injected into stage II/
III oocytes suggesting that early RNAs can interact with late
pathway localization machinery (Allen et al., 2003; Hudson
and Woodland, 1998; Zhou and King, 1996b). Since it has
been reported that late pathway localization elements do not
localize during the early pathway (Kloc and Etkin, 1998;
Zhou and King, 1996a), it has generally been assumed that
distinct machinery is responsible for localization during the
early pathway. However, through the course of our work
with stage I oocytes we discovered that the late pathway
Vg1 localization element (VgLE) does have the ability to
localize robustly and reproducibly to the mitochondrial
cloud using whole mount labeling procedures in stage I
oocytes (Fig. 1A). This localization of the VgLE appears to
be bona fide early localization since confocal images of
stage I oocytes injected with fluorescently labeled VgLE
show the VgLE distributed diffusely throughout the interior
of the mitochondrial cloud (Fig. 1B). These results indicate
that late pathway RNAs can also interact with the early
localization machinery. We attribute our ability to detect
localization of the VgLE during the early pathway to
modified procedures for oocyte isolation which produce
stage I oocytes of very high quality that routinely survive
over 1 week in culture. This ability of the VgLE to localize
during the early pathway is not entirely unexpected given
that a low amount of endogenous Vg1 becomes associated
with the mitochondrial cloud in stage I oocytes (Betley et
al., 2002). While localization of the VgLE to the mitochon-
drial cloud does not indicate whether distinct or common
factors are utilized during the early and late pathway, it does
show for the first time to our knowledge that the VgLE can
interact with early pathway localization machinery.
If distinct cellular machinery were responsible for local-
ization during the early and late pathway, small mutations
which block localization during one pathway might not
necessarily affect localization during the other. In this
regard, interactions between the VgLE and the early
pathway localization machinery, for example, could involve
sequence motifs within the VgLE that are distinct from
motifs required for localization during the late pathway. To
S. Choo et al. / Developmental Biology 278 (2005) 103–117106
investigate this possibility, we analyzed the localization
potential of mutant RNAs lacking previously characterized
functional motifs thought to be involved in only the early or
late pathway. The UUCAC or E2 motif is repeated several
times in late pathway RNAs and extensive mutational
analyses have shown that UUCAC motifs are required for
localization during the late pathway (Deshler et al., 1997,
1998; Kwon et al., 2002; Lewis et al., 2004). To determine if
UUCAC motifs are required for localization during the early
pathway we injected a mutant VgLE lacking all UUCAC
motifs, DaE2, into stage I oocytes and assayed for local-
ization to the mitochondrial cloud. The DaE2 mutant fails to
localize to the mitochondrial cloud of stage I oocytes (Fig.
1) indicating that the UUCAC motif which is known to
specify localization during the late pathway is also required
to signal localization during the early pathway. Similar
results were obtained through the analysis of UGCAC
motifs which are required for localization of the Xcat-2
mitochondrial cloud localization element (MCLE) during
the early pathway (Betley et al., 2002). To test if UGCAC-
dependent interactions play a role during the late pathway,
we injected a mutant MCLE (DUGCAC) that lacks all six
UGCAC motifs (Betley et al., 2002) into stage II, stage III,
and stage IV oocytes and assayed for localization to the
vegetal cortex. Whereas wild type MCLE localizes in stages
II–IV oocytes consistent with previous findings (Zhou and
King, 1996a), localization of the DUGCAC mutant is
completely abolished in stages II–IV oocytes (Fig. 2). Thus,
like the UUCAC motif in the VgLE, the UGCAC motifs of
the MCLE are required for interactions with the early and
late pathway localization machinery. Since these results
appear most consistent with a model proposing the existence
of common localization factors that function during the
early and late pathways, additional experiments were
designed to pursue this hypothesis.
Early and late pathway RNAs compete for common
localization factors in vivo
To test whether Vg1 and early pathway RNAs interact
with common or distinct localization machinery in vivo,
we used a competition assay in which digoxygenin labeled
VgLE (dig-VgLE) was co-injected with a molar excess of
unlabeled RNA into stage II oocytes. If the dig-VgLE and
unlabelled RNA interact with common localization factors,
specific inhibition of localization should be observed.
When the dig-VgLE was injected into stage II oocytes
with no competitor or a 300-fold molar excess of a
nonspecific Xenopus h-globin (XhG) RNA (Fig. 3),
localization to the wedge region was clearly observed in
80–86% of oocytes examined (Table 1). However, local-
ization of dig-VgLE was specifically competed by a molar
excess of unlabeled VgLE RNA (Fig. 3) with the
percentage of oocytes showing localization falling to 8%
(Table 1). These control experiments show that specific
competition for Vg1 or late localization factors occurred in
Fig. 1. Vg1 RNA can localize during the early pathway. (A) Stage I oocytes (approximately 200 Am diameter) were injected with the Xcat-2 mitochondrial
cloud localization element (MCLE), the Vg1 localization element (VgLE) or a VgLE mutant lacking all four UUCAC motifs (DaE2) labeled with digoxygenin
for detection. A fragment of the Xenopus h-globin gene (XhG) which does not localize in oocytes was used as a negative control. As can be seen, the MCLE
and VgLE show robust localization to the mitochondrial cloud (arrows), whereas the XhG and DaE2 RNAs do not. (B) Confocal images of stage I oocytes
injected with fluorescently labeled VgLE or XhG RNA. Localization of the VgLE can be seen as diffusely distributed throughout the interior of the
mitochondrial cloud (arrow), whereas the XhG RNA does not enter the cloud (arrow). This shows that the VgLE can indeed utilize the early pathway. The
germinal vesicle or nucleus (N) is indicated in various oocytes for orientation. Scale bar = 100 Am.
S. Choo et al. / Developmental Biology 278 (2005) 103–117 107
stage II oocytes. Using these same conditions, it can be
seen that localization of the dig-VgLE is also abolished by
a molar excess of early pathway RNAs. When the MCLE
or Xpat localization element (XpatLE) is used to compete
for localization of the dig-VgLE, localization is abolished
(Fig. 3) with the percentage of oocytes showing local-
ization dropping to 0% and 2%, respectively (Table 1).
This shows that the localization elements of Xcat-2 and
Fig. 3. Early pathway RNAs specifically compete for late localization factors in vivo. Early stage II oocytes (top row) were injected with digoxygenin labeled
VgLE (Blue text) and a 300-fold molar excess of the indicated competitor RNA (Black text). Localization to the wedge region (black arrows) is clear in oocytes
injected with no competitor (No Comp) or a nonspecific competitor (XhG). However, specific competition does occur with the VgLE itself, the MCLE and the
XpatLE. Stage IV oocytes (bottom row) were injected with an XhG–VgLE fusion RNA (Blue text) and a 100-fold molar excesses of the indicated competitor
RNAs (black text). Localization of the XhG–VgLE RNA was detected by in situ hybridization using a digoxygenin labeled RNA probe complementary only to
the XhG sequence. In the absence of competitor, localization is clearly observed in the vegetal cortex from the pole to the equatorial region (white arrowhead).
At this concentration, the VgLE reduces the number of oocytes showing localization (see Table 2), and those oocytes showing localization display only a
limited pattern at the vegetal pole which is turned upwards in this photograph for easy visualization. The MCLE and XpatLE competitor RNAs completely
block localization same concentration. Competition by the MCLE requires UGCAC motifs in stage IV oocytes since the DUGCAC RNA mutant fails to
compete for localization at the vegetal cortex (white arrowhead). Scale bar = 50 mm (top row) or 100 Am (bottom row).
Fig. 2. UGCAC motifs are required for localization during the late pathway. Late stage II, stage III, or stage IVoocytes were injected with digoxygenin labeled
MCLE or an MCLE mutant (DUGCAC) lacking all six UGCAC motifs. Localization to the vegetal cortex (arrows) is apparent in all oocytes injected with the
MCLE, but not in oocytes injected with the mutant. Scale bar = 100 AM.
S. Choo et al. / Developmental Biology 278 (2005) 103–117108
Xpat specifically compete for Vg1 localization factors in
stage II oocytes. Furthermore, competition for Vg1 local-
ization factors requires both the UUCAC and UGCAC
motifs since the DaE2 VgLE mutant and DUGCAC MCLE
mutant fail to compete efficiently for localization of the
dig-VgLE (Table 1).
Similar results were obtained in stage IV oocytes.
However, to ensure that the results obtained in stage II
oocytes were not due to experimental procedures such as
labeling the VgLE with digoxygenin, we used a previously
established method that relies on the ability to detect
localization of an unlabeled XhG–VgLE fusion RNA by in
situ hybridization to the XhG sequence. As reported
previously (Kwon et al., 2002), a 100-fold excess of
VgLE competitor RNA caused a large reduction in the
localization of XhG–VgLE fusion RNA (Fig. 3), and
localization frequencies dropped from 84% with no
competitor to 20% with the VgLE competitor (Table 2).
Consistent with our findings in stage II oocytes, the MCLE
and XpatLE also compete specifically for Vg1 localization
factors in stage IV oocytes (Fig. 3). Furthermore, specific
competition for late localization factors appears to be a
general feature of early pathway RNAs since the local-
ization elements of Xdazl (XdazLE) and DEADSouth
(DSLE) also compete efficiently for localization of the
XhG–VgLE fusion RNA in stage IV oocytes (Table 2).
Competition for late localization factors in stage IV oocytes
also requires the UUCAC and UGCAC motifs since the
DaE2 VgLE mutant (Kwon et al., 2002) and DUGCAC
MCLE mutant (Fig. 3, Table 2) fail to compete for
localization of the XhG–VgLE in stage IV oocytes. The
results in stages II and IV oocytes show that early pathway
RNAs specifically interact with late pathway localization
factors throughout the late pathway. Interestingly, early
pathway RNAs appear to compete more efficiently for
blateQ localization factors than Vg1, since all four early
RNAs reduce localization of the XhG–VgLE RNA to less
than the 20% obtained with the VgLE (Table 2). More-
over, unlabeled VgLE does not compete for localization of
digoxygenin labeled MCLE (dig-MCLE) when these
competition experiments are reversed (data not shown).
Together, our competition data indicate that early RNAs
out-compete late RNAs for interactions with the vegetal
localization machinery at all stages of oogenesis.
Early RNAs contain bstrongerQ localization elements than
late pathway RNAs
Since early pathway RNAs appear to compete more
efficiently for late localization factors than Vg1, we
hypothesized that early RNAs may contain bstrongerQ
localization elements than Vg1. To test this, we performed
a time course experiment to monitor localization in mid-
stage II oocytes. Approximately 250 stage II oocytes were
injected with similar amounts of digoxygenin-labeled
MCLE or VgLE. These oocytes were then cultured for
various amounts of time after which 50 oocytes were
placed in fixative until the end of the time course. RNA
localization was then assayed in all oocytes simultane-
ously, and the percentage of oocytes showing localization
at the vegetal cortex was determined. The results from this
time course experiment clearly show that the MCLE
localizes more rapidly than the VgLE (Fig. 4). More than
40% of the oocytes injected with the MCLE showed
localization at the vegetal cortex 3 h after injection,
whereas 12 h were required for a similar level of
localization in oocytes injected with the VgLE. The lower
frequencies of VgLE localization at early time points was
not due to the inability of these oocytes to efficiently
localize the VgLE as equal frequencies of localization
were obtained for the MCLE and VgLE after 48 h of
culture. These results show that the early pathway Xcat-2
RNA contains a bstrongerQ localization element than the
late pathway Vg1 RNA. The greater ability of early
pathway RNAs to compete for late localization factors
(Table 2) and the faster localization of early RNAs (Fig. 4)
during the late pathway suggest that early RNAs interact
more favorably than late RNAs with the late localization
machinery in vivo. This interpretation is consistent with
our previous observations showing that the MCLE recruits
more Kinesin II into the vegetal pathway than does the
VgLE (Betley et al., 2004); Kinesin II is required for
localization of the MCLE and VgLE to the vegetal cortex
during the late pathway (Betley et al., 2004).
Table 2
Competition of Vg1 localization in stage IV oocytes
RNA probe Competitor RNA Molar excess Localizationa
XhG–VgLE None 100Â 84% (87)
XhG–VgLE VgLE 100Â 20% (15)
XhG–VgLE MCLE 100Â 7% (14)
XhG–VgLE XpatLE 100Â 14% (21)
XhG–VgLE DSLE 100Â 3% (66)
XhG–VgLE XdazLE 100Â 2% (80)
XhG–VgLE DUGCAC 100Â 95% (21)
a
The percent localization represents the number of oocytes showing clear
localization with the total number of injected oocytes shown in parentheses.
Table 1
Competition of Vg1 localization in stage II oocytes
RNA probe Competitor RNA Molar excess Localizationa
Dig-VgLE None 300Â 80% (50)
Dig-VgLE XhG 300Â 86% (58)
Dig-VgLE VgLE 300Â 8% (37)
Dig-VgLE MCLE 300Â 0% (36)
Dig-VgLE XpatLE 150Âb
2% (39)
Dig-VgLE DaE2 300Â 70% (43)
Dig-VgLE DUGCAC 300Â 69% (39)
a
The percent localization represents the number of oocytes showing clear
localization with the total number of injected oocytes shown in parentheses.
b
The XpatLE could only be concentrated to roughly half the concentration
of the other competitor RNAs.
S. Choo et al. / Developmental Biology 278 (2005) 103–117 109
The late localization factor Vera binds specifically to early
pathway localization elements
The in vivo competition experiments show clearly that
early and late RNAs interact with common localization
factors during the late pathway. One factor implicated in the
late localization pathway is Vg RBP/Vera which is an RNA
binding protein that specifically binds to late pathway
localization elements of Veg T and Vg1 mRNAs (Bubu-
nenko et al., 2002; Deshler et al., 1997, 1998; Havin et al.,
1998; Kwon et al., 2002). Vera binds preferentially to
UUCAC motifs (Deshler et al., 1998) and point mutations in
UUCAC motifs that reduce Vera binding impair local-
ization, whereas point mutations within UUCAC motifs that
do not reduce Vera binding do not impair localization during
the late pathway (Kwon et al., 2002). Moreover, Vera
associates with late pathway RNAs in vivo (Kress et al.,
2004) and microinjection of anti-Vera antibodies reduces
localization during the late pathway (Kwon et al., 2002).
Since UUCAC motifs are required for localization during
the early pathway (Fig. 1), we decided to explore further a
potential role for Vera during the early pathway. We first
sought to determine if Vera binds specifically to the same
early pathway RNA localization elements that compete for
Vg1 localization in vivo (Table 3). To quantitatively assess
binding of Vera to early pathway RNAs we produced
recombinant Vera protein which, in UV cross-linking
assays, binds with similar sequence specificity to the VgLE
as endogenous Vera protein does in Xenopus oocyte extracts
(data not shown). The affinity of Vera for the VgLE and
other RNAs was determined using filter binding assays with
recombinant Vera and 32
P-labeled RNAs. Binding of Vera to
the VgLE is detectable at a protein concentration of 5 nM
(Fig. 5) resulting in a Kd of 11 nM (Table 3), whereas
binding to similarly sized RNA fragments of the XhG gene,
Vg1 open reading frame (ORF), Xcat-2 ORF, or the
transcription vector used in these experiments does not
occur below 200–400 nM (Fig. 5). The Kd for nonspecific
RNAs is greater than 1.4 AM (Table 3) indicating that
recombinant Vera has at least 100-fold higher affinity for the
VgLE than it does for nonspecific RNAs. Vera also binds
specifically to the same four early localization elements
MCLE, XpatLE, DSLE, and XdazLE that compete for Vg1
localization factors in vivo (Table 3). The Kd for early
pathway RNAs lies in a range from 17 to 47 nM (Fig. 5,
Table 3) which is similar to the Kd for Vg1, but no
correlation between affinities for purified Vera in vitro and
pathway selectivity in vivo was apparent. Since Vera binds
specifically to the localization elements of four early
pathway RNAs (Fig. 5, Table 3), and deletion of UUCAC
motifs blocks localization of the VgLE during the early
pathway (Fig. 1), it appears that Vera may play a role in
localization during the early pathway.
A putative consensus binding site for Vera has previously
been determined to be WYCAY (W = A or U, Y =
pyrimidine) (Fig. 6A) (Deshler et al., 1998). This consensus
site was determined using a short RNA probe containing
four tandem copies of the UUCAC motif. Since this short
RNA has a much lower affinity for Vera than the full length
MCLE or VgLE (data not shown), high-affinity binding of
Vera to RNA localization elements is likely to involve other
interactions in addition to those mediated by WYCAY
motifs. Nevertheless, we sought to test further the involve-
ment of Vera during the early pathway by creating point
mutations in putative Vera binding sites. To do this, we
converted all nine WYCAY motifs in the VgLE to WACAY.
This particular mutant design was chosen because we
wanted to determine if Vera binding could be diminished
without affecting localization. Such a finding would negate
Table 3
Vera specifically binds early pathway localization elements
RNA Kd (nM)
VgLE 11 F 2
MCLE 47 F 5
XpatLE 24 F 3
DSLE 40 F 9
XdazLE 17 F 2
XhG )
N1400
Vg 1 ORF
Vector
Xcat-2 ORF
Kd’s for the indicated RNA were determined using filter binding assays.
The standard error of the mean is shown.
Fig. 4. Early pathway RNAs have bstronger Q localization elements than late
pathway RNAs. Oocytes were injected with either the Vg1 (VgLE) or Xcat-2
(MCLE) localization element, and a group of oocytes were placed in fixative
at the indicated time points. The percentage of oocytes showing localization
was determined and plotted as percent (%) localization. Note that the MCLE
localizes faster than the VgLE and that after 48 h, the localization frequencies
of each RNA are indistinguishable from each other.
S. Choo et al. / Developmental Biology 278 (2005) 103–117110
Fig. 5. Recombinant Vera binds specifically and with high affinity to early pathway RNAs. (A) Filter binding experiment showing increasing
concentrations of Vera at the bottom and several 32
P-labeled RNAs. XhG is the Xenopus h Globin fragment commonly used as a negative control for
RNA localization. In addition, fragments of the transcription vector and of the Xcat-2 and Vg1 open reading frames (ORFs) were used as negative controls
for binding. (B) Binding curves for the VgLE, MCLE, XpatLE, DEADSouth LE (DSLE), and the XdazLE. The error bars represent the standard deviations
in fraction bound at each protein concentration from at least five different experiments for each RNA. The Kd was determined for each RNA and is shown
in Table 3.
S. Choo et al. / Developmental Biology 278 (2005) 103–117 111
an essential role for Vera in the early pathway. Since ACAY
motifs are common in the localization elements of vegetal
RNAs (Betley et al., 2002), we believed converting the
second position to A in the WYCAY motifs would be the
most likely point mutation to reveal Vera-independent
localization if such localization were possible. This muta-
tion, however, abolished localization of the VgLE during
the early pathway since the Vg9WACAY mutant failed to
localize to the mitochondrial cloud of stage I oocytes (Fig.
6B). Identical results were obtained in stage II/III oocytes
consistent with a role for WYCAY motifs for the late
pathway (data not shown). It was striking to observe the
severe impairment of localization created by these point
mutations given that larger 15–20 nucleotide deletion
mutations along the entire Vg1 localization element have
little or no detectable effect on localization during the late
pathway (Deshler et al., 1997; Gautreau et al., 1997). When
the relative Vera binding was measured for the Vg9WACAY
mutant, a 40% reduction was observed when compared to
the wild type VgLE (Fig. 6C). This is a smaller decrease
than seen with the DaE2 mutant VgLE which has four of the
nine WYCAY motifs precisely deleted, shows a 60%
reduction in Vera binding (Fig. 6C) and also blocks
localization during the early pathway (Fig. 1). We have
analyzed a number of other point mutations in the VgLE
and MCLE for Vera binding and localization during the
early and late pathways. While we have not yet determined
whether these mutations affect Vera binding by eliminating
specific binding sites for Vera or by altering the secondary
structure of these RNAs, we observe a correlation such that
point mutations which incrementally reduce Vera binding
show a corresponding decrease in localization during the
Fig. 6. Point mutations in WYCAY motifs reduce Vera binding and RNA localization during the early pathway. (A) The RNA sequence of the 366 nucleotide
VgLE is shown with WYCAY motifs underlined, and the UUCAC version of WYCAY motifs over and underlined. (B) Stage I oocytes were injected with the
wild type VgLE, the DaE2 mutant VgLE which has all UUCAC motifs deleted, or the Vg9WACAY mutant which has all nine WYCAYs converted to the
corresponding WACAYs. Localization of the VgLE to the mitochondrial cloud is severely impaired by mutations in these putative Vera binding sites. Scale
bar = 50 Am. (C) Relative binding affinity of Vera for the VgLE and the Vg9WACAY mutant which shows a 40 percent decrease in relative Vera binding. The
DaE2 mutant, which has been used to implicate Vera during the late pathway, is shown for comparison and demonstrates a 60% reduction in Vera binding in
these assays. Error bar reflects the standard error of mean values generated in each binding assay.
S. Choo et al. / Developmental Biology 278 (2005) 103–117112
early and late pathways (data not shown). While these data
do not prove a requirement for Vera during the localization
process, they strongly suggest that Vera, a factor previously
implicated only in the late pathway, is also involved in
localization during the early pathway. This analysis also
lends further support to our conclusion that the same cis
elements are required for localization during the early and
late pathways since point mutations that disrupt localization
Fig. 7. Early pathway RNAs specifically associate with a perinuclear microtubule network that extends to the mitochondrial cloud of stage I oocytes. (A)
Schematic diagram of a stage I oocyte showing the germinal vesicle or nucleus (N) and mitochondrial cloud (M). An optical section tangential to the nucleus
and mitochondrial cloud is depicted as a parallelogram and is similar to the ones shown in (B) and (C, top row only). (An) and (Vg) signify the future animal–
vegetal axis. (B) Visualization of the perinuclear microtubule (MT) network (green) in a stage I oocyte. Perinuclear microtubule bundles (arrowheads) radially
oriented with respect to the mitochondrial cloud and perinuclear microtubule enriched foci (arrows) are indicated. (C) Double labeling of microtubules (green)
and fluorescently MCLE labeled RNA (red) 1 h (top row) or 6 h (middle row) after injection of the RNA. Note that after 1 h little or no MCLE RNA is detected
in the mitochondrial cloud (M), but many RNA particles can be seen associated with perinuclear bundles of microtubules (arrowheads) extending toward the
mitochondrial cloud. After 6 h, several MCLE particles are associated with perinuclear microtubule foci (arrow) and extensive localization of the MCLE to the
mitochondrial cloud is apparent (large arrowhead). Within the mitochondrial cloud, however, the RNA becomes dispersed and the large particles are not
apparent. At this same time point, the XhG RNA (bottom row) shows no accumulation of particles on perinuclear foci (arrows) or within the mitochondrial
cloud (large arrowhead). At the 6-h time points, optical sections are oriented along the future animal–vegetal axis and run through the mitochondrial cloud and
nucleus. Scale bar = 10 Am.
S. Choo et al. / Developmental Biology 278 (2005) 103–117 113
during the early pathway also impair localization during the
late pathway.
Xcat-2 RNA associates with a perinuclear microtubule
network that extends to the mitochondrial cloud of stage I
oocytes
The data obtained in our competition, biochemical, and
mutational analyses uniformly point to the existence of
common localization machinery participating in the early
and late pathway. However, previous reports employing the
use of agents that disrupt microtubules have suggested that
microtubules are required for the late, but not for the early
pathway (Kloc and Etkin, 1995; Kloc et al., 1996; Yisraeli et
al., 1990). This has contributed to an alternative view
implicating distinct mechanisms for RNA localization
during the early and late pathways. Because stage I oocytes
possess a population of acetylated and presumably stable
microtubules (Gard, 1999) and the early pathway local-
ization element of Xcat-2 can recruit a microtubule-depend-
ent motor in stage II oocytes (Betley et al., 2004), we
decided to reexamine a potential role for microtubules
during the early pathway. To do this, we assessed the spatial
relationship between microtubules, the mitochondrial cloud,
and fluorescently labeled MCLE RNA injected into stage I
oocytes.
Confocal imaging of stage I oocytes labeled with an anti-
a-tubulin antibody showed that microtubules are distributed
throughout the cytoplasm and are enriched in the cortical
region as previously reported (Gard, 1999). However,
optical sections acquired at the periphery of the nucleus
(Fig. 7A) revealed a network consisting of microtubule
bundles that are wrapped around the nucleus and intersect at
dense patches or foci (Fig. 7B). These foci are easily
distinguished from the surrounding network due to their
association with numerous radially oriented microtubule
bundles and their enhanced staining. Interestingly, this
perinuclear network of microtubule fibers extends into the
mitochondrial cloud which itself is surrounded by a ring of
microtubules (Fig. 7B).
To characterize the distribution of microtubules and
RNAs utilizing the early pathway, we injected fluorescently
labeled MCLE or XhG RNA into stage I oocytes and
visualized the microtubules using immunocytochemistry.
Oocytes were fixed 1 or 6 h after injection of RNA.
Remarkably, at both time points, we observed large particles
with the MCLE that associated with bundles and foci of the
microtubule network and seemed to follow the network into
the mitochondrial cloud. The XhG RNA did not show any
apparent association with this network (Fig. 7C). After 1 h
much of the MCLE RNA remains dispersed throughout the
cytoplasm, and the mitochondrial cloud does not yet show
labeling by localized RNA. However, at this time, we
observe large MCLE RNA particles or granules that
colocalize with perinuclear foci and microtubule fibers
extending into the mitochondrial cloud (Fig. 7C, top row).
At the later time point (6 h), the MCLE has accumulated in
the mitochondrial cloud (Fig. 7C, middle row), whereas the
XhG RNA does not (Fig. 7C, bottom row). Interestingly,
once the MCLE RNA enters the mitochondrial cloud, it
becomes diffusely distributed, and large particles are no
longer visible. This suggests that the MCLE RNA granules
dissociate into smaller complexes once they reach their final
destination in stage I oocytes as has been observed with the
myelin basic protein mRNA in oligodendrocytes (Ainger et
al., 1993). In addition to association with microtubule
bundles of the perinuclear network and localization in the
mitochondrial cloud, we also observe an accumulation of
MCLE particles on perinuclear foci that interconnect
microtubule bundles of the network. One striking example
can be seen in Fig. 7C (middle row), where a number of
MCLE particles are clustered on one of these foci. No
accumulation of granules on perinuclear foci was detected in
oocytes injected with XhG RNA (Fig. 7C, bottom row).
This organization of the microtubule network and specific
association of the MCLE RNA with it in stage I oocytes
suggest that microtubule dependent transport may play a
role in RNA localization during the early pathway as it does
during the late pathway (Betley et al., 2002; Kloc and Etkin,
1995; Yisraeli et al., 1990).
Discussion
In Xenopus, the spatial distributions of different
mRNAs in growing oocytes have indicated that at least
two temporally defined pathways exist for RNA local-
ization to the vegetal cortex (Forristall et al., 1995; Kloc
and Etkin, 1995). Here, we have designed experiments to
explore whether distinct or common cellular machinery
recognizes and directs RNA localization during the early
and late pathway. We have shown that a late pathway
RNA, Vg1, can localize during the early pathway and that
the same short (5 nt) cis motifs required for late
localization are also required for early localization and
vice versa. Competition experiments reveal that the local-
ization elements of four different early pathway RNAs
including Xcat-2, Xpat, Xdazl, and DEADSouth specifi-
cally compete for late localization factors in vivo. In
addition, these same early pathway RNAs are bound
specifically by the late localization factor Vg RBP/Vera,
and point mutations that reduce Vera binding impair
localization during the early and late pathways. We also
show that the Xcat-2 RNA localization element forms
particles which associate with a microtubule network that
surrounds the nucleus and extends into the mitochondrial
cloud of stage I oocytes suggesting that directed transport
may play a role in the early pathway as it does during the
late pathway. Taken together, these results support a model
in which a common set of cellular factors participates in
the localization of vegetal RNAs during the early and late
pathways. While this model in no way excludes the
S. Choo et al. / Developmental Biology 278 (2005) 103–117114
existence of pathway-specific factors, it suggests that the
early and late pathways may have evolved as a conse-
quence of distinct RNAs acquiring the ability to interact
differentially with common or basal RNA localization
machinery that remains active throughout stages I–IV of
oogenesis.
The role of short cis motifs in RNA localization to the
vegetal cortex
RNA localization elements that specify localization to
the vegetal cortex contain a variety of short repeated motifs
that are important for signaling interactions with cellular
machinery required for the localization process (Betley et
al., 2002; Bubunenko et al., 2002; Deshler et al., 1997,
1998; Kwon et al., 2002). These motifs may provide
binding sites for RNA binding proteins or they may
participate in intramolecular or intermolecular RNA–RNA
interactions required for association with the localization
machinery. Moreover, it is possible that individual motifs
participate in a combination of these functions during
various aspects of the localization process. In these studies,
we show that WYCAY motifs are required for localization
during the early as well as the late pathway. Since WYCAY
is a consensus binding site for Vg RBP/Vera (Deshler et al.,
1997) and point mutations that disrupt Vera binding impair
localization, our findings suggest that at least one function
of WYCAY motifs is to bind Vera which is known to be
expressed during the early and late pathways (Deshler et al.,
1997).
We have also shown that the UGCAC motif is required
for localization during both pathways (Fig. 2) (Betley et al.,
2002), but this motif is not bound specifically by Vera in
vitro as previously reported (Betley et al., 2002; Deshler et
al., 1998). In fact, the DUGCAC MCLE mutant binds Vera
with an affinity that is indistinguishable from the wild type
MCLE (Kd = 46 and 47 nM, respectively). Therefore, the
UGCAC motif may interact with another protein component
of the localization machinery. Alternatively, or possibly in
addition, UGCAC motifs may mediate RNA–RNA base
pairing interactions required for localization. This is
plausible because the first four nucleotides of UGCAC are
palindromic giving them the ability to base pair with each
other. Furthermore, RNA–RNA base pairing is thought to
be important for the microtubule dependent localization of
bicoid RNA in Drosophila (Ferrandon et al., 1997; Wagner
et al., 2001). Interestingly, the double-stranded RNA bind-
ing protein, Staufen, mediates localization of bicoid mRNA
in Drosophila oocytes (St. Johnston et al., 1991), and
Xenopus Staufen1 has recently been implicated in the
localization of RNAs in Xenopus oocytes (Yoon and
Mowry, 2004). Thus, it is possible that RNA–RNA
interactions also play a role in localizing RNAs in Xenopus
oocytes. Hence, more work is required to determine whether
UGCAC motifs mediate RNA–protein or RNA–RNA
interactions required for the localization process.
What factors limit the RNA localization process in vivo?
Our in vivo competition results may provide a clue about
the localization factors that limit the localization process in
oocytes. The concentration of RNA required to compete the
localization machinery in stage I/II oocytes (300 Am
diameter) is approximately 6 AM. However, the concen-
tration of Vera in these oocytes is roughly 4.5 times higher
(27 F 4 AM) as determined by Western blot analysis (data
not shown) consistent with amounts reported to be present
in stage IVoocytes (Kwon et al., 2002). Such a high level of
endogenous Vera may explain why antibody inhibition
experiments designed to bknock outQ Vera function only
partially disrupt the localization process in oocytes (Kwon et
al., 2002). Therefore, Vera may not be limiting for local-
ization in vivo. Another factor implicated in vegetal RNA
localization in Staufen1 (Yoon and Mowry, 2004). The
endogenous levels of Xenopus Staufen may be lower than
that of Vera. We have recently cloned Staufen2 in Xenopus
and shown that its concentration is 10 F 1 AM in stage I/II
oocytes (Z. Chen, manuscript in preparation) which is
similar to the levels required to saturate the localization
machinery in vivo. Thus, it is possible that XStau2 may be
limiting for localization in oocytes as it appears to be for the
localization of bulk mRNA into the dendrites of neurons
(Tang et al., 2001). Identifying the concentration-limited
components of the localization machinery will be an
important first step for understanding the regulation of the
RNA localization process.
Previous work has suggested that the Vg1 localization
element forms an RNA–protein complex consisting of at
least six RNA binding proteins that may be required to
initiate the RNA localization process in oocytes (Mowry,
1996). Since motifs that bind Vera (UUCAC) (Kwon et
al., 2002) and motifs that do not bind Vera (UGCAC)
(Fig. 3 and Table 1) are both required for competing
localization in vivo, it is possible that RNA localization
elements must maintain their capacity to form a bpre-
localization complexQ in order to compete for the local-
ization machinery in vivo. The formation of such a
putative multi-protein complex could involve cooperative
interactions such that mutating specific binding sites for
Vera or other critical components would be sufficient to
render the mutant RNA incapable of assembling the
complex and competing for the localization machinery in
vivo. Evidence for the existence of Vera and Staufen
containing RNP complexes in vivo has recently been
provided by co-immunoprecipitation assays of Vera,
XStau1 and the Vg1 RNA in stage III oocytes (Kress et
al., 2004; Yoon and Mowry, 2004). The formation of this
bpre-localizationQ complex would then lead to the recruit-
ment of molecular motors such as heterotrimeric Kinesin
II for transport to the cortex (Betley et al., 2004). In this
scenario, mutations that disrupt formation of the pre-
localization complex would affect all phases of this
presumably multi-step localization process.
S. Choo et al. / Developmental Biology 278 (2005) 103–117 115
Directed versus diffusion-mediated RNA transport
The arrangement of the microtubule network with
respect to the mitochondrial cloud and the specific
association of Xcat-2 RNA granules with it (Fig. 7)
indicate that a directed transport mechanism may function
during the early pathway. However, treatment of stage I
oocytes with nocodazole has little or no effect on the
distribution of early pathway RNAs (Kloc and Etkin, 1995;
Kloc et al., 1996) leading some to propose that directed
RNA transport does not play a role during the early
pathway (Chang et al., 2004). We believe that two factors
may contribute to this paradox. First, it has been shown that
acetylated microtubules exist in the perinuclear region of
stage I oocytes (Gard, 1999), and while nocodazole
disrupts the organization of microtubules in stage I oocytes
(data not shown), the detection of a subpopulation of
microtubules could be obscured in nocodazole treated
oocytes. Interestingly, a stable and acetylated population
of microtubules has been found to play a role in oocyte
specification in Drosophila (Roper and Brown, 2004).
Secondly, diffusion must also play an important role in the
localization process. In fact, interactions between Kinesin II
and vegetal RNAs appear to be quite dynamic in vivo
(Betley et al., 2004). Thus, due to the small size of stage I
oocytes, diffusion may be sufficient for the localization of
injected RNAs to the mitochondrial cloud. However,
diffusion would not be expected to be sufficient for the
localization of vegetal RNAs in larger stage III/IV during
the late pathway. The consideration of stable microtubules
and/or the possibility of diffusion-mediated localization of
injected RNAs in small stage I oocytes may help explain
why previous work has invoked two distinct mechanisms
for RNA localization during the early and late pathways;
these studies relied exclusively on detecting the sensitivity
of RNA localization to microtubule depolymerizing drugs
in stage I (approximately 200 Am diameter) and stage IV
(approximately 800 Am diameter) oocytes. In any case,
given that the MCLE forms RNA granules that associate
with microtubules entering the mitochondrial cloud (Fig.
7), we now believe that diffusion and directed transport
work together in stage I as well as stage III/IV oocytes to
mediate the selective localization of vegetal RNAs during
the early and late pathways, respectively, in Xenopus
oocytes.
What determines when distinct RNAs become localized
during oogenesis?
We have shown that the Vg1 localization element is
capable of localizing to the mitochondrial cloud if injected
into stage I oocytes. What then prevents endogenous Vg1
from localizing to the mitochondrial cloud of stage I
oocytes? One possibility is that negatively acting cis-
regulatory elements in either the VgLE or other regions of
Vg1 serve to attenuate or inhibit localization until stage II of
oogenesis. Another possibility is based on our observed
kinetic differences in the localization of and early and late
pathway localization elements. It is clear that the MCLE
localizes faster than the VgLE during the late pathway in
stage II oocytes (Fig. 4). It is also known that the early
pathway RNA, Xlsirt, is first to localize to the cortex when
co-injected with the VgLE into stage II oocytes (Kloc et al.,
1996). Taken together, these observations show that early
RNAs contain stronger localization elements than late
pathway RNAs. Therefore, at endogenous concentrations,
Vg1 may not sufficiently compete for the necessary local-
ization machinery until early pathway RNAs become
localized in the mitochondrial cloud or anchored at the
vegetal cortex. Interestingly, in very early stage I oocytes
(approximately 50 Am diameter), endogenous Xcat-2 is
distributed throughout the cytoplasm (Zhou and King,
1996a) as is endogenous Vg1 mRNA in later stage I
oocytes (Forristall et al., 1995; Kloc and Etkin, 1995). This
combined with results presented in this work suggests that a
similar pathway may be followed for both endogenous
RNAs, but that each follows this pathway at slightly
different times during oogenesis.
While our work does not exclude the possibility of
pathway-specific RNA localization factors, we have pro-
vided evidence for the existence of common or basal
localization machinery that is involved in both the early and
late pathways. This may provide an economical mean by
which oocytes generate a high number of temporally
defined pathways without evolving novel machinery for
each one. In fact, it has been shown by double-labeling in
situ hybridization that different early pathway RNAs are
actually localized to distinct domains within the mitochon-
drial cloud of stage I oocytes suggesting that different early
RNAs utilize distinct temporally defined pathways (Kloc
and Etkin, 1995). Moreover, we have identified an RNA,
Xlerk, which clearly uses a pathway that is intermediate
between early and late RNAs (Betley et al., 2002). In this
model, relative differences in the affinity of specific RNAs
for basal components of the localization machinery could
determine when a transcript is localized during oogenesis.
Interestingly, two recent reports have shown that early and
late RNAs interact specifically with a number of RNA
binding proteins that are distinct from Vera (Chang et al.,
2004; Claussen et al., 2004) supporting the notion that
common machinery is utilized during the early and late
pathways. The identification and biochemical character-
ization of additional factors involved in the localization
process will be required to fully understand how such a
mechanism is regulated to establish polarized distributions
of distinct factors in the egg.
Acknowledgments
This work was supported by the National Institutes of
Health Grant R01 GM59682 to J.O.D.
S. Choo et al. / Developmental Biology 278 (2005) 103–117116
References
Ainger, K., Avossa, D., Morgan, F., Hill, S.J., Barry, C., Barbarese, E.,
Carson, J.H., 1993. Transport and localization of exogenous myelin
basic protein mRNA microinjected into oligodendrocytes. J. Cell Biol.
123, 431–441.
Alarcon, V.B., Elinson, R.P., 2001. RNA anchoring in the vegetal cortex of
the Xenopus oocyte. J. Cell Sci. 114, 1731–1741.
Allen, L., Kloc, M., Etkin, L.D., 2003. Identification and characterization of the
Xlsirt cis-acting RNA localization element. Differentiation 71, 311–321.
Betley, J.N., Frith, M.C., Graber, J.H., Choo, S., Deshler, J.O., 2002. A
ubiquitous and conserved signal for RNA localization in chordates.
Curr. Biol. 12, 1756–1761.
Betley, J.N., Heinrich, B., Vernos, I., Sardet, C., Prodon, F., Deshler, J.O.,
2004. Kinesin II meditates Vg1 mRNA transport in Xenopus oocytes.
Curr. Biol. 14, 219–224.
Bubunenko, M., Kress, T.L., Vempati, U.D., Mowry, K.L., King, M.L.,
2002. A consensus RNA signal that directs germ layer determinants to
the vegetal cortex of Xenopus oocytes. Dev. Biol. 248, 82–92.
Chang, P., Torres, J., Lewis, R.A., Mowry, K.L., Houliston, E., King, M.L.,
2004. Localization of RNAs to the mitochondrial cloud in Xenopus
oocytes through entrapment and association with endoplasmic retic-
ulum. Mol. Biol. Cell 15, 4669–4681.
Claussen, M., Horvay, K., Pieler, T., 2004. Evidence for overlapping, but
not identical, protein machineries operating in vegetal RNA localization
along early and late pathways in Xenopus oocytes. Development 131,
4263–4273.
Cote, C.A., Gautreau, D., Denegre, J.M., Kress, T.L., Terry, N.A., Mowry,
K.L., 1999. A Xenopus protein related to hnRNP I has a role in
cytoplasmic RNA localization. Mol. Cell 4, 431–437.
Deshler, J.O., Highett, M.I., Schnapp, B.J., 1997. Localization of Xenopus
Vg1 mRNA by Vera protein and the endoplasmic reticulum. Science
276, 1128–1131.
Deshler, J.O., Highett, M.I., Abramson, T., Schnapp, B.J., 1998. A highly
conserved RNA-binding protein for cytoplasmic mRNA localization in
vertebrates. Curr. Biol. 8, 489–496.
Dumont, J.N., 1972. Oogenesis in Xenopus laevis (Daudin). I. Stages of
oocyte development in laboratory maintained animals. J. Morphol. 136,
153–179.
Ferrandon, D., Koch, I., Westhof, E., Nusslein-Volhard, C., 1997. RNA–
RNA interaction is required for the formation of specific bicoid mRNA 3V
UTR-STAUFEN ribonucleoprotein particles. EMBO J. 16, 1751–1758.
Forristall, C., Pondel, M., Chen, L., King, M.L., 1995. Patterns of
localization and cytoskeletal association of two vegetally localized
RNAs, Vg1 and Xcat-2. Development 121, 201–208.
Gard, D.L., 1999. Confocal microscopy and 3-D reconstruction of the
cytoskeleton of Xenopus oocytes. Microsc. Res. Tech. 44, 388–414.
Gautreau, D., Cote, C.A., Mowry, K.L., 1997. Two copies of a subelement
from the Vg1 RNA localization sequence are sufficient to direct vegetal
localization in Xenopus oocytes. Development 124, 5013–5020.
Havin, L., Git, A., Elisha, Z., Oberman, F., Yaniv, K., Schwartz, S.P.,
Standart, N., Yisraeli, J.K., 1998. RNA-binding protein conserved in
both microtubule- and microfilament-based RNA localization. Genes
Dev. 12, 1593–1598.
Heasman, J., Wessely, O., Langland, R., Craig, E.J., Kessler, D.S., 2001.
Vegetal localization of maternal mRNAs is disrupted by VegT
depletion. Dev. Biol. 240, 377–386.
Houston, D.W., Zhang, J., Maines, J.Z., Wasserman, S.A., King, M.L.,
1998. A Xenopus DAZ-like gene encodes an RNA component of germ
plasm and is a functional homologue of Drosophila boule. Develop-
ment 125, 171–180.
Hudson, C., Woodland, H.R., 1998. Xpat, a gene expressed specifically in
germ plasm and primordial germ cells of Xenopus laevis. Mech. Dev.
73, 159–168.
Kloc, M., Etkin, L.D., 1994. Delocalization of Vg1 mRNA from the vegetal
cortex in Xenopus oocytes after destruction of Xlsirt RNA. Science 265,
1101–1103.
Kloc, M., Etkin, L.D., 1995. Two distinct pathways for the localization of
RNAs at the vegetal cortex in Xenopus oocytes. Development 121,
287–297.
Kloc, M., Etkin, L.D., 1998. Apparent continuity between the messenger
transport organizer and late RNA localization pathways during
oogenesis in Xenopus. Mech. Dev. 73, 95–106.
Kloc, M., Etkin, L.D., 1999. Analysis of localized RNAs in Xenopus
Oocytes. In: Richter, J.D. (Ed.), A Comparative Methods Approach to
the Study of Oocytes and Embryos. Oxford Univ. Press, Inc., New
York, pp. 256–278.
Kloc, M., Spohr, G., Etkin, L.D., 1993. Translocation of repetitive RNA
sequences with the germ plasm in Xenopus oocytes. Science 262,
1712–1714.
Kloc, M., Larabell, C., Etkin, L.D., 1996. Elaboration of the messenger
transport organizer pathway for localization of RNA to the vegetal
cortex of Xenopus oocytes. Dev. Biol. 180, 119–130.
Kloc, M., Bilinski, S., Pui-Yee Chan, A., Etkin, L.D., 2000. The targeting
of Xcat2 mRNA to the germinal granules depends on a cis-acting
germinal granule localization element within the 3VUTR. Dev. Biol.
217, 221–229.
Kloc, M., Bilinski, S., Chan, A.P., Allen, L.H., Zearfoss, N.R., Etkin, L.D.,
2001. RNA localization and germ cell determination in Xenopus. Int.
Rev. Cytol. 203, 63–91.
Kress, T.L., Yoon, Y.J., Mowry, K.L., 2004. Nuclear RNP complex assembly
initiates cytoplasmic RNA localization. J. Cell Biol. 165, 203–211.
Ku, M., Melton, D.A., 1993. Xwnt-11: a maternally expressed Xenopus wnt
gene. Development 119, 1161–1173.
Kwon, S., Abramson, T., Munro, T.P., John, C.M., Kohrmann, M.,
Schnapp, B.J., 2002. UUCAC and Vera dependent localization of
VegT RNA in Xenopus oocytes. Curr. Biol. 12, 558–564.
Lewis, R.A., Kress, T.L., Cote, C.A., Gautreau, D., Rokop, M.E., Mowry,
K.L., 2004. Conserved and clustered RNA recognition sequences are a
critical feature of signals directing RNA localization in Xenopus
oocytes. Mech. Dev. 121, 101–109.
MacArthur, H., Houston, D.W., Bubunenko, M., Mosquera, L., King, M.L.,
2000. DEADSouth is a germ plasm specific DEAD-box RNA helicase
in Xenopus related to eIF4A. Mech. Dev. 95, 291–295.
Mowry, K.L., 1996. Complex formation between stage-specific oocyte
factors and a Xenopus mRNA localization element. Proc. Natl. Acad.
Sci. U. S. A. 93, 14608–14613.
Palacios, I.M., St. Johnston, D., 2001. Getting the message across: the
intracellular localization of mRNAs in higher eukaryotes. Annu. Rev.
Cell Dev. Biol. 17, 569–614.
Roper, K., Brown, N.H., 2004. A spectraplakin is enriched on the fusome
and organizes microtubules during oocyte specification in Drosophila.
Curr. Biol. 14, 99–110.
St. Johnston, D., Beuchle, D., Nusslein-Volhard, C., 1991. Staufen, a gene
required to localize maternal RNAs in the Drosophila egg. Cell 66,
51–63.
Tang, S.J., Meulemans, D., Vazquez, L., Colaco, N., Schuman, E., 2001. A
role for a rat homolog of Staufen in the transport of RNA to neuronal
dendrites. Neuron 32, 463–475.
Wagner, C., Palacios, I., Jaeger, L., St. Johnston, D., Ehresmann, B.,
Ehresmann, C., Brunel, C., 2001. Dimerization of the 3VUTR of bicoid
mRNA involves a two-step mechanism. J. Mol. Biol. 313, 511–524.
Yisraeli, J.K., Sokol, S., Melton, D.A., 1990. A two-step model for the
localization of maternal mRNA in Xenopus oocytes: involvement of
microtubules and microfilaments in the translocation and anchoring of
Vg1 mRNA. Development 108, 289–298.
Yoon, Y.J., Mowry, K.L., 2004. Xenopus Staufen is a component of a
ribonucleoprotein complex containing Vg1 RNA and Kinesin. Develop-
ment 131, 3035–3045.
Zhou, Y., King, M.L., 1996a. Localization of Xcat-2 RNA, a putative germ
plasm component, to the mitochondrial cloud in Xenopus stage I
oocytes. Development 122, 2947–2953.
Zhou, Y., King, M.L., 1996b. RNA transport to the vegetal cortex of
Xenopus oocytes. Dev. Biol. 179, 173–183.
S. Choo et al. / Developmental Biology 278 (2005) 103–117 117

More Related Content

What's hot

Gutell 022.mbp.1992.52.0075
Gutell 022.mbp.1992.52.0075Gutell 022.mbp.1992.52.0075
Gutell 022.mbp.1992.52.0075Robin Gutell
 
Comparative transcriptomics
Comparative transcriptomicsComparative transcriptomics
Comparative transcriptomicsSayak Ghosh
 
Genetic organization of eukaryotes and prokaryotes
Genetic organization of eukaryotes and prokaryotes Genetic organization of eukaryotes and prokaryotes
Genetic organization of eukaryotes and prokaryotes Theabhi.in
 
RFLP - Restriction Fragment Length Polymorphism
RFLP - Restriction Fragment Length PolymorphismRFLP - Restriction Fragment Length Polymorphism
RFLP - Restriction Fragment Length PolymorphismDeepa Arumugam
 
nucleo-cytoplasmic intraction,Anterograde and Retrograde singnaling
nucleo-cytoplasmic intraction,Anterograde and Retrograde singnalingnucleo-cytoplasmic intraction,Anterograde and Retrograde singnaling
nucleo-cytoplasmic intraction,Anterograde and Retrograde singnalingBalaji Rathod
 
Genetic Engineering presentation
Genetic Engineering presentationGenetic Engineering presentation
Genetic Engineering presentationBALAJI SANTHAKUMAR
 
Organellar genome and its composition
Organellar genome and its compositionOrganellar genome and its composition
Organellar genome and its compositionShilpa C
 
The language of life (all the subtitles)first ppt 2 bimester
The language of life (all the subtitles)first ppt 2 bimesterThe language of life (all the subtitles)first ppt 2 bimester
The language of life (all the subtitles)first ppt 2 bimesterSofia Paz
 
Concept of gene
Concept of geneConcept of gene
Concept of geneAman Ullah
 
Mitochondrial DNA in Taxonomy and Phylogeny
Mitochondrial DNA in Taxonomy and PhylogenyMitochondrial DNA in Taxonomy and Phylogeny
Mitochondrial DNA in Taxonomy and PhylogenyRachel Jacob
 
485 lec4 the_genome
485 lec4 the_genome485 lec4 the_genome
485 lec4 the_genomehhalhaddad
 

What's hot (20)

Gutell 022.mbp.1992.52.0075
Gutell 022.mbp.1992.52.0075Gutell 022.mbp.1992.52.0075
Gutell 022.mbp.1992.52.0075
 
Final Draft
Final DraftFinal Draft
Final Draft
 
Comparative transcriptomics
Comparative transcriptomicsComparative transcriptomics
Comparative transcriptomics
 
PIIS089662731300319X
PIIS089662731300319XPIIS089662731300319X
PIIS089662731300319X
 
Genetic organization of eukaryotes and prokaryotes
Genetic organization of eukaryotes and prokaryotes Genetic organization of eukaryotes and prokaryotes
Genetic organization of eukaryotes and prokaryotes
 
Chloroplast dna
Chloroplast dnaChloroplast dna
Chloroplast dna
 
Organelle DNA
Organelle DNAOrganelle DNA
Organelle DNA
 
RFLP - Restriction Fragment Length Polymorphism
RFLP - Restriction Fragment Length PolymorphismRFLP - Restriction Fragment Length Polymorphism
RFLP - Restriction Fragment Length Polymorphism
 
nucleo-cytoplasmic intraction,Anterograde and Retrograde singnaling
nucleo-cytoplasmic intraction,Anterograde and Retrograde singnalingnucleo-cytoplasmic intraction,Anterograde and Retrograde singnaling
nucleo-cytoplasmic intraction,Anterograde and Retrograde singnaling
 
Genetic Engineering presentation
Genetic Engineering presentationGenetic Engineering presentation
Genetic Engineering presentation
 
Regulatory RNA at epigenetic level
Regulatory RNA at epigenetic level Regulatory RNA at epigenetic level
Regulatory RNA at epigenetic level
 
MITOCHONDRIAL GENETICS
MITOCHONDRIAL GENETICSMITOCHONDRIAL GENETICS
MITOCHONDRIAL GENETICS
 
Organellar genome and its composition
Organellar genome and its compositionOrganellar genome and its composition
Organellar genome and its composition
 
Mitochondrial and chloroplast DNA
Mitochondrial and chloroplast DNAMitochondrial and chloroplast DNA
Mitochondrial and chloroplast DNA
 
Genomics
GenomicsGenomics
Genomics
 
Mitochondrial dna
Mitochondrial dnaMitochondrial dna
Mitochondrial dna
 
The language of life (all the subtitles)first ppt 2 bimester
The language of life (all the subtitles)first ppt 2 bimesterThe language of life (all the subtitles)first ppt 2 bimester
The language of life (all the subtitles)first ppt 2 bimester
 
Concept of gene
Concept of geneConcept of gene
Concept of gene
 
Mitochondrial DNA in Taxonomy and Phylogeny
Mitochondrial DNA in Taxonomy and PhylogenyMitochondrial DNA in Taxonomy and Phylogeny
Mitochondrial DNA in Taxonomy and Phylogeny
 
485 lec4 the_genome
485 lec4 the_genome485 lec4 the_genome
485 lec4 the_genome
 

Viewers also liked

Semantic enrichment of places with vgi sources a knowledge based approach
Semantic enrichment of places with vgi sources a knowledge based approachSemantic enrichment of places with vgi sources a knowledge based approach
Semantic enrichment of places with vgi sources a knowledge based approachCamille Tardy
 
Noe viaje extraordinario_mt
Noe viaje extraordinario_mtNoe viaje extraordinario_mt
Noe viaje extraordinario_mtNoemi Basarte
 
RECORRIDO POR EL MUSEO MITRE
RECORRIDO POR EL MUSEO MITRERECORRIDO POR EL MUSEO MITRE
RECORRIDO POR EL MUSEO MITREmajocastilla
 
Central Neuroxial blockage ( Spinal and Epidural block ) By Dr Sachin Gaikwad
Central Neuroxial blockage ( Spinal and Epidural block ) By Dr Sachin GaikwadCentral Neuroxial blockage ( Spinal and Epidural block ) By Dr Sachin Gaikwad
Central Neuroxial blockage ( Spinal and Epidural block ) By Dr Sachin GaikwadSachin Gaikwad
 
Identificando el factor mínimo de una iglesia local
Identificando el factor mínimo de una iglesia localIdentificando el factor mínimo de una iglesia local
Identificando el factor mínimo de una iglesia localRecursos Cristianos. Org
 
Atelier numérique : Google analytics :comprendre les statistiques de mon site...
Atelier numérique : Google analytics :comprendre les statistiques de mon site...Atelier numérique : Google analytics :comprendre les statistiques de mon site...
Atelier numérique : Google analytics :comprendre les statistiques de mon site...Destination Brocéliande
 

Viewers also liked (7)

Semantic enrichment of places with vgi sources a knowledge based approach
Semantic enrichment of places with vgi sources a knowledge based approachSemantic enrichment of places with vgi sources a knowledge based approach
Semantic enrichment of places with vgi sources a knowledge based approach
 
Noe viaje extraordinario_mt
Noe viaje extraordinario_mtNoe viaje extraordinario_mt
Noe viaje extraordinario_mt
 
RECORRIDO POR EL MUSEO MITRE
RECORRIDO POR EL MUSEO MITRERECORRIDO POR EL MUSEO MITRE
RECORRIDO POR EL MUSEO MITRE
 
Prueba
PruebaPrueba
Prueba
 
Central Neuroxial blockage ( Spinal and Epidural block ) By Dr Sachin Gaikwad
Central Neuroxial blockage ( Spinal and Epidural block ) By Dr Sachin GaikwadCentral Neuroxial blockage ( Spinal and Epidural block ) By Dr Sachin Gaikwad
Central Neuroxial blockage ( Spinal and Epidural block ) By Dr Sachin Gaikwad
 
Identificando el factor mínimo de una iglesia local
Identificando el factor mínimo de una iglesia localIdentificando el factor mínimo de una iglesia local
Identificando el factor mínimo de una iglesia local
 
Atelier numérique : Google analytics :comprendre les statistiques de mon site...
Atelier numérique : Google analytics :comprendre les statistiques de mon site...Atelier numérique : Google analytics :comprendre les statistiques de mon site...
Atelier numérique : Google analytics :comprendre les statistiques de mon site...
 

Similar to Choo et al., 2004

DevOpsとcloudで達成する再現性のあるDNAシーケンス解析とスーパーコンピューティング
DevOpsとcloudで達成する再現性のあるDNAシーケンス解析とスーパーコンピューティングDevOpsとcloudで達成する再現性のあるDNAシーケンス解析とスーパーコンピューティング
DevOpsとcloudで達成する再現性のあるDNAシーケンス解析とスーパーコンピューティングItoshi Nikaido
 
Writing assignment 4 molecular cell biology
Writing assignment 4   molecular cell biologyWriting assignment 4   molecular cell biology
Writing assignment 4 molecular cell biologycorv629
 
Ceraseetal2021complete.pdf
Ceraseetal2021complete.pdfCeraseetal2021complete.pdf
Ceraseetal2021complete.pdfAndreaCerase3
 
Ptacin_et_al-2013-Cellular_Microbiology (review)
Ptacin_et_al-2013-Cellular_Microbiology (review)Ptacin_et_al-2013-Cellular_Microbiology (review)
Ptacin_et_al-2013-Cellular_Microbiology (review)Jerod Ptacin
 
Undergraduate Research Grant
Undergraduate Research GrantUndergraduate Research Grant
Undergraduate Research GrantKaitlin Zoccola
 
Perales and Bentley MolCell 2009
Perales and Bentley MolCell 2009Perales and Bentley MolCell 2009
Perales and Bentley MolCell 2009Roberto Perales PhD
 
Identification and Characterization of Saccharomyces cerevisiae Cdc6 DNA-bind...
Identification and Characterization of Saccharomyces cerevisiae Cdc6 DNA-bind...Identification and Characterization of Saccharomyces cerevisiae Cdc6 DNA-bind...
Identification and Characterization of Saccharomyces cerevisiae Cdc6 DNA-bind...gan-navi
 
Akhtar et al. Stem Cell Reports (2015)
Akhtar et al. Stem Cell Reports (2015)Akhtar et al. Stem Cell Reports (2015)
Akhtar et al. Stem Cell Reports (2015)Aslam Akhtar, MS
 
Plant Cell-2005-Mlotshwa-2873-85
Plant Cell-2005-Mlotshwa-2873-85Plant Cell-2005-Mlotshwa-2873-85
Plant Cell-2005-Mlotshwa-2873-85Braden Roth
 
The Thiazide-sensitive NaCl
The Thiazide-sensitive NaClThe Thiazide-sensitive NaCl
The Thiazide-sensitive NaClAvin Snyder
 

Similar to Choo et al., 2004 (20)

NSMB_2008
NSMB_2008NSMB_2008
NSMB_2008
 
DevOpsとcloudで達成する再現性のあるDNAシーケンス解析とスーパーコンピューティング
DevOpsとcloudで達成する再現性のあるDNAシーケンス解析とスーパーコンピューティングDevOpsとcloudで達成する再現性のあるDNAシーケンス解析とスーパーコンピューティング
DevOpsとcloudで達成する再現性のあるDNAシーケンス解析とスーパーコンピューティング
 
Dawaliby et al 2010
Dawaliby et al 2010Dawaliby et al 2010
Dawaliby et al 2010
 
Dawaliby et al 2010
Dawaliby et al 2010Dawaliby et al 2010
Dawaliby et al 2010
 
replicación 2010.pdf
replicación 2010.pdfreplicación 2010.pdf
replicación 2010.pdf
 
Writing assignment 4 molecular cell biology
Writing assignment 4   molecular cell biologyWriting assignment 4   molecular cell biology
Writing assignment 4 molecular cell biology
 
Ceraseetal2021complete.pdf
Ceraseetal2021complete.pdfCeraseetal2021complete.pdf
Ceraseetal2021complete.pdf
 
Summary of Doctoral Research
Summary of Doctoral ResearchSummary of Doctoral Research
Summary of Doctoral Research
 
Ptacin_et_al-2013-Cellular_Microbiology (review)
Ptacin_et_al-2013-Cellular_Microbiology (review)Ptacin_et_al-2013-Cellular_Microbiology (review)
Ptacin_et_al-2013-Cellular_Microbiology (review)
 
Undergraduate Research Grant
Undergraduate Research GrantUndergraduate Research Grant
Undergraduate Research Grant
 
Perales and Bentley MolCell 2009
Perales and Bentley MolCell 2009Perales and Bentley MolCell 2009
Perales and Bentley MolCell 2009
 
Presentation final
Presentation finalPresentation final
Presentation final
 
zmk2820
zmk2820zmk2820
zmk2820
 
Identification and Characterization of Saccharomyces cerevisiae Cdc6 DNA-bind...
Identification and Characterization of Saccharomyces cerevisiae Cdc6 DNA-bind...Identification and Characterization of Saccharomyces cerevisiae Cdc6 DNA-bind...
Identification and Characterization of Saccharomyces cerevisiae Cdc6 DNA-bind...
 
Akhtar et al. Stem Cell Reports (2015)
Akhtar et al. Stem Cell Reports (2015)Akhtar et al. Stem Cell Reports (2015)
Akhtar et al. Stem Cell Reports (2015)
 
PNAS_2005
PNAS_2005PNAS_2005
PNAS_2005
 
Mundo viral
Mundo viralMundo viral
Mundo viral
 
replicación ADN.pdf
replicación ADN.pdfreplicación ADN.pdf
replicación ADN.pdf
 
Plant Cell-2005-Mlotshwa-2873-85
Plant Cell-2005-Mlotshwa-2873-85Plant Cell-2005-Mlotshwa-2873-85
Plant Cell-2005-Mlotshwa-2873-85
 
The Thiazide-sensitive NaCl
The Thiazide-sensitive NaClThe Thiazide-sensitive NaCl
The Thiazide-sensitive NaCl
 

Choo et al., 2004

  • 1. Evidence for common machinery utilized by the early and late RNA localization pathways in Xenopus oocytes Soheun Chooa,1 , Bianca Heinrichb,1 , J. Nicholas Betleyb , Zhao Chenb , James O. Deshlera,b,* a Molecular Biology and Biochemistry Program, Boston University, Boston, Massachusetts 02215, USA b Department of Biology, Boston University, 5 Cummington Street, Boston, Massachusetts 02215, USA Received for publication 13 July 2004, revised 25 October 2004, accepted 27 October 2004 Available online 18 November 2004 Abstract In Xenopus, an early and a late pathway exist for the selective localization of RNAs to the vegetal cortex during oogenesis. Previous work has suggested that distinct cellular mechanisms mediate localization during these pathways. Here, we provide several independent lines of evidence supporting the existence of common machinery for RNA localization during the early and late pathways. Data from RNA microinjection assays show that early and late pathway RNAs compete for common localization factors in vivo, and that the same short RNA sequence motifs are required for localization during both pathways. In addition, quantitative filter binding assays demonstrate that the late localization factor Vg RBP/Vera binds specifically to several early pathway RNA localization elements. Finally, confocal imaging shows that early pathway RNAs associate with a perinuclear microtubule network that connects to the mitochondrial cloud of stage I oocytes suggesting that motor driven transport plays a role during the early pathway as it does during the late pathway. Taken together, our data indicate that common machinery functions during the early and late pathways. Thus, RNA localization to the vegetal cortex may be a regulated process such that differential interactions with basal factors determine when distinct RNAs are localized during oogenesis. D 2004 Elsevier Inc. All rights reserved. Keywords: RNA localization; Xenopus; Oocytes; Vg1 RBP; Vera; Vg1; Xcat-2; KH domain; Microtubules Introduction The subcellular localization of specific mRNAs to distinct regions of a cell is an important mechanism for generating asymmetric gene expression in germ cells and somatic cells (Kloc et al., 2001; Palacios and St. Johnston, 2001), and the sequences that specify localization generally reside in the 3V untranslated region (3V UTR) of localized transcripts. In Xenopus, two temporally defined pathways exist for localizing specific RNAs to the vegetal cortex during oogenesis (Forristall et al., 1995; Kloc and Etkin, 1995). RNAs that utilize the early pathway, such as Xcat-2, are distributed uniformly throughout the cytoplasm of early stage I oocytes (50–100 Am diameter), but become localized to the mitochondrial cloud in later stage I oocytes (120–300 Am diameter) and subsequently associate with the vegetal cortex of stage II oocytes (Zhou and King, 1996a). Additional examples of RNAs that use the early pathway include Xpat (Hudson and Woodland, 1998), Xwnt-11 (Ku and Melton, 1993), DEADSouth (MacArthur et al., 2000) Xdazl (Houston et al., 1998), and Xlsirts (Kloc et al., 1993). Treatment of stage I oocytes with microtubule depolymeriz- ing drugs has little or no effect on the distribution of endogenous early pathway RNAs associated with the mitochondrial cloud which has led to a view that the early pathway functions independently of microtubules (Kloc and Etkin, 1995). The late or Vg1 pathway begins when Xcat-2 and other early RNAs become associated with the vegetal cortex. In early and late stage I oocytes, Vg1 is distributed throughout the cytoplasm and little is detected in the mitochondrial cloud (Betley et al., 2002; Forristall et al., 1995; Kloc and Etkin, 1995). In stage II oocytes, endogenous Vg1 becomes 0012-1606/$ - see front matter D 2004 Elsevier Inc. All rights reserved. doi:10.1016/j.ydbio.2004.10.019 * Corresponding author. Department of Biology, Boston University, 5 Cummington Street, Boston, MA 02215. Fax: +1 617 353 8484. E-mail address: jdeshler@bu.edu (J.O. Deshler). 1 These two authors contributed equally to this work. Developmental Biology 278 (2005) 103–117 www.elsevier.com/locate/ydbio
  • 2. associated with a wedge-shaped region of cytoplasm between the oocyte nucleus and the vegetal pole (Kloc and Etkin, 1995, 1998). This wedge-shaped region is enriched in membranes of the endoplasmic reticulum (ER), and it has been proposed that the ER may play a role in the localization process (Deshler et al., 1997; Kloc and Etkin, 1998). Subsequent transport of Vg1 from this wedge-shaped region to the vegetal cortex of stage II/III oocytes requires microtubules (Kloc and Etkin, 1995; Yisraeli et al., 1990) and heterotrimeric kinesin II (Betley et al., 2004). In stage III/IV oocytes, actin filaments (Kloc and Etkin, 1995; Yisraeli et al., 1990), cytokeratin filaments (Alarcon and Elinson, 2001), and localized RNAs them- selves (Heasman et al., 2001; Kloc and Etkin, 1994) help to anchor early and late RNAs close to the plasma membrane in the vegetal cortex. In addition to these cytoskeletal requirements, three RNA binding proteins have been implicated in the late localization pathway. HnRNP I binds the consensus sequence YYUCU (Y = pyrimidine) (Lewis et al., 2004) that is required for efficient localization of two late pathway RNAs, Vg1 and VegT (Bubunenko et al., 2002; Cote et al., 1999), and Vg RBP/Vera (Deshler et al., 1998; Havin et al., 1998) is an ER associated protein that binds specifically to the sequence UUCAC which is also essential for localization during the late pathway (Deshler et al., 1997, 1998; Kwon et al., 2002). Recently, a double-stranded RNA binding protein, Staufen, has also been implicated as a factor required for localization during the late pathway (Yoon and Mowry, 2004). The differences in timing and possible cytoskeletal requirements have led to a prevailing view that distinct mechanisms are responsible for RNA localization during the early and late pathway. However, several observations are not entirely consistent with this model. For example, RNAs which normally use the early pathway, such as Xcat-2 (Zhou and King, 1996b), Xpat (Hudson and Woodland, 1998), and Xlsirts (Allen et al., 2003), will localize to the vegetal cortex during the late pathway if injected into stage III oocytes. Moreover, this ectopic localization of early RNAs during the late pathway requires intact microtubules (Zhou and King, 1996b) and Kinesin II (Betley et al., 2004) indicating that early RNAs are able to interact specifically with the late pathway localization machinery in vivo. In addition, the LEs specifying early or late localization have similar sequence characteristics; both have an evolutionarily conserved abundance of short CAC-containing motifs that are essential for localization (Betley et al., 2002; Bubunenko et al., 2002; Deshler et al., 1997, 1998; Kwon et al., 2002). These observations suggest that the early and late pathway may share a common set of cellular factors that are required for RNA localization in stage I as well as in stage II/III oocytes. Here, we provide four independent lines of evidence each of which supports the hypothesis that the early and late pathways utilize common cellular machinery for RNA localization. First, we show that small motifs, which are required for localization during the late pathway, are also required for localization during the early pathway and vice versa. Second, competition assays designed to detect whether two different RNAs interact with common or distinct localization factors in vivo show that early and late pathway RNAs specifically compete for the same local- ization factors in oocytes. Third, biochemical assays show that the late localization factor Vg RBP/Vera binds specifically to several early pathway LEs, and point mutations that reduce Vera binding block localization during the early pathway. Fourth, confocal imaging reveals that early pathway RNAs associate specifically with a micro- tubule network that surrounds the nucleus and extends into the mitochondrial cloud of stage I oocytes suggesting that directed transport may play a role during the early pathway as it does during the late pathway. Together, these findings strongly suggest that early and late pathway RNAs interact with basal localization machinery that functions throughout the early and late pathways. Materials and methods Plasmid constructs, mutants, and in vitro transcription reactions RNAs for microinjection were transcribed from linear- ized plasmid DNA templates. The VgLE RNA was tran- scribed from XbaI linearized p19.366 which was constructed by cloning a blunt ended 366 nucleotide BsmI–SspI fragment of the Vg1 gene into HincII digested pTZ19U (Bio-Rad Laboratories). The DaE2 RNA was transcribed from XbaI digested pDaE2 which is identical to p19.366 except that the UUCAC motifs were deleted by site directed mutagenesis as previously described (Deshler et al., 1997). The Xcat-2 MCLE was transcribed from HindIII digested pMCLE which was made by cloning a PCR product containing the MCLE (nucleotides 391–630 of the Xcat-2 gene) into SalI–HindIII digested pTZ18U (partial MCS). This vector is a modified version of pTZ18U (Bio-Rad Laboratories) that has the promoter proximal half of the polylinker removed by deleting the region between the EcoRI and XbaI sites using endonuclease digestion, an end fill-in reaction with the Klenow fragment of DNA polymerase, and self-ligation. The DUGCAC mutant RNA was transcribed from pMCLEDUGCAC which is identical to pMCLE except all six UGCAC motifs are deleted as described previously (Betley et al., 2002). The localization elements of DEADSouth (DSLE) and Xdaz1 (XdazLE) were transcribed from BamHI digested pDSLE and pXdazLE, respectively, which were constructed by cloning PCR products containing the localization elements of DEAD- South (DSLE) (nt 1796–2125) or Xdaz1 (XdazLE) (1335– 1780) (Betley et al., 2002) into EcoRI–BamHI digested pTZ18U (partial MCS). The XpatLE RNA was transcribed from SalI digested pXpatLE which was constructed by cloning a PCR fragment containing the XpatLE (nt 1063– S. Choo et al. / Developmental Biology 278 (2005) 103–117104
  • 3. 1589 of the Xpat gene) (Betley et al., 2002) into BamHI– SalI digested pTZ18U. All constructs were verified by DNA sequencing. The Vg9WYCAY mutant was constructed by site directed mutagenesis using sequential PCR mediated mutagenesis with partially overlapping oligonucleotide amplification primers. The mutant VgLE fragment replaced wild type fragment of p19.366 which was verified by DNA sequencing. Unlabeled competitor RNAs were transcribed in vitro using 250 Units of T7 RNA polymerase (Epicentre) in 500 Al reactions containing 20 Ag of linearized plasmid DNA template, 4 mM ATP, CTP, GTP and UTP, 20 mM MgCl2, 10 ACi 32 P-labeled UTP, 0.25 units inorganic pyrophospha- tase (Sigma), 1 mM DTT, 40 mM Tris, pH 8.0, 2 mM spermidine, and 200 Units of RNAse inhibitor (Promega). These reactions were incubated at 378C for 5 h, treated with DNAse, extracted with phenol and precipitated with 2.5 M ammonium acetate, 2.5 volumes of ethanol, and 0.5 Ag/Al glycogen (Roche). Digoxygenin-labeled RNAs were tran- scribed in 50 Al reactions containing 1 Ag of plasmid template, 25 units of T7 RNA polymerase, 0.33 mM ATP, CTP, and GTP, 0.22 mM UTP, 110 AM Digoxigenin-11- UTP (Roche), 8 mM MgCl2, 1 mM DTT, 10 ACi 32 P-labeled UTP, 40 units RNAse inhibitor, 40 mM Tris, pH 8.0, and 2 mM spermidine. These reactions proceeded at 378C for 2–3 h followed by precipitation with 2.5 M ammonium acetate, 2.5 volumes of ethanol, and 30 Ag glycogen. RNAs were quantified either by A260 or by determining the percent of nucleotides incorporated into full-length RNAs during the transcription reaction. The quality of all RNAs was verified by autoradiographic analysis after running the RNAs in 6% polyacrylamide gels containing 8 M urea. In vivo localization and competition assays All oocytes were isolated from Xenopus laevis (Nasco) and staged by size according to Dumont (Dumont, 1972). Stage I (50–300 Am diameter) and Stage II (300–450 Am diameter) oocytes were isolated from 5 to 7 cm females as previously described (Kloc and Etkin, 1999) except 100 Ag/ml liberase blendzyme III (Roche) was used instead of type II collagenase. Stage IV (450–600 Am diameter) oocytes were isolated from 6 to 8 cm albino females (Nasco) using the same method, but with 200 Ag/ml liberase blendzyme III. The segment of RNA that targets Xcat-2 to the mitochondrial cloud has been referred to as the mitochondrial cloud localization element (MCLE) and resides in the first approximately 220 nucleotides of the Xcat-2 3V UTR (Kloc et al., 2000; Zhou and King, 1996a). To determine if the MCLE RNA lacking the UGCAC motifs localizes in stages II–IV oocytes, 15 fmol of digoxygenin labeled MCLE or DUGCAC mutant RNA (Betley et al., 2002) was injected into each oocyte and the localization signal was detected as previously described with developing times of 20 min (Kloc and Etkin, 1999). Localization assays in stage I oocytes (approximately 200 Am diameter) involved injecting 0.1 fmol of RNA in 0.1 nl and developing the colorometric reaction for 45 min. In these oocytes, the mitochondrial clouds are adjacent to the germinal vesicle and do not yet associate with the vegetal cortex. For competition assays in early stage II oocytes (approximately 325 Am diameter), the amount of injected digoxygenin-labeled VgLE was titrated to determine the lower limit of reliably detecting localization to the vegetal wedge-shaped region. This stage of oogenesis is easily distinguished from stage I because the mitochondrial cloud begins to contact the vegetal cortex and forms the vegetal wedge. Localization was detected in 80– 90% of oocytes injected with 0.3 fmol of dig-VgLE, but not with half this amount (data not shown). Varying amounts of unlabeled MCLE, VgLE or Xenopus h-globin (XhG) were then co-injected with dig-VgLE to determine the molar excess required for specific competition. A 200- to 300-fold molar excess of unlabeled VgLE RNA was found to be required to inhibit localization in N90% of injected oocytes. Competition of unlabeled XhG–VgLE in stage IV oocytes was done as previously described (Kwon et al., 2002) and localization of the XhG–VgLE RNA was detected by whole mount in situ hybridization using digoxygenin labeled antisense XhG RNA. The time course analysis of Fig. 4 was performed in mid-stage II oocytes (360–380 Am diameter). Oocytes were injected with 4 fmol of dig-MCLE or dig-VgLE in 1.5 nl and the colorometric reaction was developed for 20 min for all oocytes examined. Recombinant Vera and in vitro RNA binding assays The full length Vera gene was amplified using the polymerase chain reaction (PCR) and oligonucleotide primers containing unique restriction sites. The PCR product was cloned into EcoRI–BglII digested pQE60 (Qiagen) resulting in the expression plasmid, pQEVeraHIS which place a 6Â-His tag at the C-terminus of Vera. The se- quences of the upstream and downstream PCR primers were 5V-AATTTCACACAGAATTCATTAAAGAGGAG- AAATTAACCATGAACAAGCTGTATATTGGAAA- CCTC-3V and 5V-TTTTCCTGAAAGATCTTTTT- CTTCTCGGTTGGGGTTG-3V, respectively. One hundred milliliters of XL-1 Blue containing pQEVeraHIS was grown overnight in rich medium containing ampicillin (100 Ag/ml). This culture was diluted 10 fold in the same medium and allowed to grow at 378C to an A600 of 0.8 at which point Vera expression was induced with 1 mM IPTG for 4 h at 378C. Following centrifugation, the cell pellet was resus- pended in 15 mL of high salt buffer (HSB) (1 M NaCl, 50 mM sodium phosphate buffer, pH 7.8, 1 % Triton X-100, 1 mM DTT) and the cells were homogenized by running them through a French press twice. Cell debris was cleared by centrifugation, and the soluble fraction was stored at À708C in 300-Al aliquots. On the day of filter binding, one aliquot of the soluble fraction was mixed with 10 Al of HSB- equilibrated Ni-NTA resin (Qiagen) and allowed to rotate at 48C for 1 h. The protein-bound resin was washed once each S. Choo et al. / Developmental Biology 278 (2005) 103–117 105
  • 4. with 300 Al HSB, PBS containing 1 mM DTT, and PBS containing 1 mM DTT and 75 mM imidazole, and Vera was eluted off the beads using 10 Al of PBS containing 1 mM DTT, 500 mM imidazole, and 4 M urea. Vera protein concentration was determined using the micro protein assay (Bio-Rad) with bovine serum albumin as a standard. For filter binding assays, 50 Al reactions containing 50 Ag/ml yeast tRNA (Roche) and 5 mg/ml heparin, 104 cpm of 32 P- labeled RNA probes (b0.1 nM final concentration), and varying amounts of Vera in 1Â FBA buffer (5% glycerol, 100 mM NaCl, 50 mM Tris, pH 8.0, 1 mM MgCl2, 1 mM DTT) were incubated at 248C for 30 min. These reactions were then filtered through a Hybond-C membrane (Amer- sham Biosciences) which had been equilibrated by filtering 100 Al of 1Â FBA and 50 Al of 1Â FBA containing 0.1 mg/ ml yeast tRNA per well using a dot–blot vacuum apparatus (Bio-Rad). The membrane was washed with 200 Al of 1Â FBA per well, and the amount of bound RNA was measured on a Phosphorimager screen (Molecular Dynamics). Dis- sociation constants (Kd) were determined by calculating the amount of recombinant Vera required to bind half of the RNA probe using Sigma Plot to curve fit the data. All Kd values represent the average of at least five independent measurements performed on different days and the standard deviations are shown. To determine relative binding affinities for Vera, the Kd for the wild type RNA was divided by the Kd for mutant RNAs. This calculation eliminates any error in the wild type RNA measurement because by definition, its value becomes one. The relative Vera binding for mutant RNAs was averaged from at least six measurements and plotted in the histograms shown with error bars representing the standard deviation of each value. Since the value for the wild type is always one, there is no error in its value, and thus no error bars are shown for the wild type VgLE. Confocal imaging The localization element for Vg1 (VgLE), Xcat-2 (MCLE), or the XhG RNA was fluorescently labeled for injections into stage I oocytes with Alexa-Fluor-546-14- UTP (Molecular Probes) during in vitro transcription as recommended by the manufacturer. Oocytes were injected with the indicated fluorescent RNA and cultured in 0.5 Â Leibovitz L-15 medium (Sigma) containing 7.5 mM HEPES, pH 7.6, 10 mg/ml tetracycline (Sigma) and 50 mg/ml gentamycin (Roche). Oocytes were then fixed in MEMFA (0.1 M MOPS, pH 7.5, 1 mM EGTA, 1 mM MgSO4, 1.4% formaldehyde) or MEMFA supplemented with 0.5 AM taxol (Paclitaxel) (Sigma) to preserve micro- tubules. Immunocytochemistry was performed as previously described (Betley et al., 2004) using a 1:250 dilution of the DM1A monoclonal anti-a-tubulin antibody (Sigma) and a 1:300 dilution of fluorescein-conjugated donkey anti-mouse secondary antibody (Jackson ImmunoResearch). Oocytes were mounted in Vectashield (Vector Laboratories, Inc.) for imaging. Confocal images were acquired on a Zeiss LSM 510 confocal laser scanning microscope with an Argon Laser (488 nm) for detection of fluorescein (green) and a Helium-Neon laser (543 nm) for detection of Alexa-Fluor 546 (red). No green fluorescence was detected in control oocytes incubated with secondary but not primary anti- bodies. No red fluorescence was detected in control oocytes not injected with fluorescent RNA. Results The same functional motifs are required for localization during the early and late pathway Previous reports have shown that early pathway RNAs will localize during the late pathway if injected into stage II/ III oocytes suggesting that early RNAs can interact with late pathway localization machinery (Allen et al., 2003; Hudson and Woodland, 1998; Zhou and King, 1996b). Since it has been reported that late pathway localization elements do not localize during the early pathway (Kloc and Etkin, 1998; Zhou and King, 1996a), it has generally been assumed that distinct machinery is responsible for localization during the early pathway. However, through the course of our work with stage I oocytes we discovered that the late pathway Vg1 localization element (VgLE) does have the ability to localize robustly and reproducibly to the mitochondrial cloud using whole mount labeling procedures in stage I oocytes (Fig. 1A). This localization of the VgLE appears to be bona fide early localization since confocal images of stage I oocytes injected with fluorescently labeled VgLE show the VgLE distributed diffusely throughout the interior of the mitochondrial cloud (Fig. 1B). These results indicate that late pathway RNAs can also interact with the early localization machinery. We attribute our ability to detect localization of the VgLE during the early pathway to modified procedures for oocyte isolation which produce stage I oocytes of very high quality that routinely survive over 1 week in culture. This ability of the VgLE to localize during the early pathway is not entirely unexpected given that a low amount of endogenous Vg1 becomes associated with the mitochondrial cloud in stage I oocytes (Betley et al., 2002). While localization of the VgLE to the mitochon- drial cloud does not indicate whether distinct or common factors are utilized during the early and late pathway, it does show for the first time to our knowledge that the VgLE can interact with early pathway localization machinery. If distinct cellular machinery were responsible for local- ization during the early and late pathway, small mutations which block localization during one pathway might not necessarily affect localization during the other. In this regard, interactions between the VgLE and the early pathway localization machinery, for example, could involve sequence motifs within the VgLE that are distinct from motifs required for localization during the late pathway. To S. Choo et al. / Developmental Biology 278 (2005) 103–117106
  • 5. investigate this possibility, we analyzed the localization potential of mutant RNAs lacking previously characterized functional motifs thought to be involved in only the early or late pathway. The UUCAC or E2 motif is repeated several times in late pathway RNAs and extensive mutational analyses have shown that UUCAC motifs are required for localization during the late pathway (Deshler et al., 1997, 1998; Kwon et al., 2002; Lewis et al., 2004). To determine if UUCAC motifs are required for localization during the early pathway we injected a mutant VgLE lacking all UUCAC motifs, DaE2, into stage I oocytes and assayed for local- ization to the mitochondrial cloud. The DaE2 mutant fails to localize to the mitochondrial cloud of stage I oocytes (Fig. 1) indicating that the UUCAC motif which is known to specify localization during the late pathway is also required to signal localization during the early pathway. Similar results were obtained through the analysis of UGCAC motifs which are required for localization of the Xcat-2 mitochondrial cloud localization element (MCLE) during the early pathway (Betley et al., 2002). To test if UGCAC- dependent interactions play a role during the late pathway, we injected a mutant MCLE (DUGCAC) that lacks all six UGCAC motifs (Betley et al., 2002) into stage II, stage III, and stage IV oocytes and assayed for localization to the vegetal cortex. Whereas wild type MCLE localizes in stages II–IV oocytes consistent with previous findings (Zhou and King, 1996a), localization of the DUGCAC mutant is completely abolished in stages II–IV oocytes (Fig. 2). Thus, like the UUCAC motif in the VgLE, the UGCAC motifs of the MCLE are required for interactions with the early and late pathway localization machinery. Since these results appear most consistent with a model proposing the existence of common localization factors that function during the early and late pathways, additional experiments were designed to pursue this hypothesis. Early and late pathway RNAs compete for common localization factors in vivo To test whether Vg1 and early pathway RNAs interact with common or distinct localization machinery in vivo, we used a competition assay in which digoxygenin labeled VgLE (dig-VgLE) was co-injected with a molar excess of unlabeled RNA into stage II oocytes. If the dig-VgLE and unlabelled RNA interact with common localization factors, specific inhibition of localization should be observed. When the dig-VgLE was injected into stage II oocytes with no competitor or a 300-fold molar excess of a nonspecific Xenopus h-globin (XhG) RNA (Fig. 3), localization to the wedge region was clearly observed in 80–86% of oocytes examined (Table 1). However, local- ization of dig-VgLE was specifically competed by a molar excess of unlabeled VgLE RNA (Fig. 3) with the percentage of oocytes showing localization falling to 8% (Table 1). These control experiments show that specific competition for Vg1 or late localization factors occurred in Fig. 1. Vg1 RNA can localize during the early pathway. (A) Stage I oocytes (approximately 200 Am diameter) were injected with the Xcat-2 mitochondrial cloud localization element (MCLE), the Vg1 localization element (VgLE) or a VgLE mutant lacking all four UUCAC motifs (DaE2) labeled with digoxygenin for detection. A fragment of the Xenopus h-globin gene (XhG) which does not localize in oocytes was used as a negative control. As can be seen, the MCLE and VgLE show robust localization to the mitochondrial cloud (arrows), whereas the XhG and DaE2 RNAs do not. (B) Confocal images of stage I oocytes injected with fluorescently labeled VgLE or XhG RNA. Localization of the VgLE can be seen as diffusely distributed throughout the interior of the mitochondrial cloud (arrow), whereas the XhG RNA does not enter the cloud (arrow). This shows that the VgLE can indeed utilize the early pathway. The germinal vesicle or nucleus (N) is indicated in various oocytes for orientation. Scale bar = 100 Am. S. Choo et al. / Developmental Biology 278 (2005) 103–117 107
  • 6. stage II oocytes. Using these same conditions, it can be seen that localization of the dig-VgLE is also abolished by a molar excess of early pathway RNAs. When the MCLE or Xpat localization element (XpatLE) is used to compete for localization of the dig-VgLE, localization is abolished (Fig. 3) with the percentage of oocytes showing local- ization dropping to 0% and 2%, respectively (Table 1). This shows that the localization elements of Xcat-2 and Fig. 3. Early pathway RNAs specifically compete for late localization factors in vivo. Early stage II oocytes (top row) were injected with digoxygenin labeled VgLE (Blue text) and a 300-fold molar excess of the indicated competitor RNA (Black text). Localization to the wedge region (black arrows) is clear in oocytes injected with no competitor (No Comp) or a nonspecific competitor (XhG). However, specific competition does occur with the VgLE itself, the MCLE and the XpatLE. Stage IV oocytes (bottom row) were injected with an XhG–VgLE fusion RNA (Blue text) and a 100-fold molar excesses of the indicated competitor RNAs (black text). Localization of the XhG–VgLE RNA was detected by in situ hybridization using a digoxygenin labeled RNA probe complementary only to the XhG sequence. In the absence of competitor, localization is clearly observed in the vegetal cortex from the pole to the equatorial region (white arrowhead). At this concentration, the VgLE reduces the number of oocytes showing localization (see Table 2), and those oocytes showing localization display only a limited pattern at the vegetal pole which is turned upwards in this photograph for easy visualization. The MCLE and XpatLE competitor RNAs completely block localization same concentration. Competition by the MCLE requires UGCAC motifs in stage IV oocytes since the DUGCAC RNA mutant fails to compete for localization at the vegetal cortex (white arrowhead). Scale bar = 50 mm (top row) or 100 Am (bottom row). Fig. 2. UGCAC motifs are required for localization during the late pathway. Late stage II, stage III, or stage IVoocytes were injected with digoxygenin labeled MCLE or an MCLE mutant (DUGCAC) lacking all six UGCAC motifs. Localization to the vegetal cortex (arrows) is apparent in all oocytes injected with the MCLE, but not in oocytes injected with the mutant. Scale bar = 100 AM. S. Choo et al. / Developmental Biology 278 (2005) 103–117108
  • 7. Xpat specifically compete for Vg1 localization factors in stage II oocytes. Furthermore, competition for Vg1 local- ization factors requires both the UUCAC and UGCAC motifs since the DaE2 VgLE mutant and DUGCAC MCLE mutant fail to compete efficiently for localization of the dig-VgLE (Table 1). Similar results were obtained in stage IV oocytes. However, to ensure that the results obtained in stage II oocytes were not due to experimental procedures such as labeling the VgLE with digoxygenin, we used a previously established method that relies on the ability to detect localization of an unlabeled XhG–VgLE fusion RNA by in situ hybridization to the XhG sequence. As reported previously (Kwon et al., 2002), a 100-fold excess of VgLE competitor RNA caused a large reduction in the localization of XhG–VgLE fusion RNA (Fig. 3), and localization frequencies dropped from 84% with no competitor to 20% with the VgLE competitor (Table 2). Consistent with our findings in stage II oocytes, the MCLE and XpatLE also compete specifically for Vg1 localization factors in stage IV oocytes (Fig. 3). Furthermore, specific competition for late localization factors appears to be a general feature of early pathway RNAs since the local- ization elements of Xdazl (XdazLE) and DEADSouth (DSLE) also compete efficiently for localization of the XhG–VgLE fusion RNA in stage IV oocytes (Table 2). Competition for late localization factors in stage IV oocytes also requires the UUCAC and UGCAC motifs since the DaE2 VgLE mutant (Kwon et al., 2002) and DUGCAC MCLE mutant (Fig. 3, Table 2) fail to compete for localization of the XhG–VgLE in stage IV oocytes. The results in stages II and IV oocytes show that early pathway RNAs specifically interact with late pathway localization factors throughout the late pathway. Interestingly, early pathway RNAs appear to compete more efficiently for blateQ localization factors than Vg1, since all four early RNAs reduce localization of the XhG–VgLE RNA to less than the 20% obtained with the VgLE (Table 2). More- over, unlabeled VgLE does not compete for localization of digoxygenin labeled MCLE (dig-MCLE) when these competition experiments are reversed (data not shown). Together, our competition data indicate that early RNAs out-compete late RNAs for interactions with the vegetal localization machinery at all stages of oogenesis. Early RNAs contain bstrongerQ localization elements than late pathway RNAs Since early pathway RNAs appear to compete more efficiently for late localization factors than Vg1, we hypothesized that early RNAs may contain bstrongerQ localization elements than Vg1. To test this, we performed a time course experiment to monitor localization in mid- stage II oocytes. Approximately 250 stage II oocytes were injected with similar amounts of digoxygenin-labeled MCLE or VgLE. These oocytes were then cultured for various amounts of time after which 50 oocytes were placed in fixative until the end of the time course. RNA localization was then assayed in all oocytes simultane- ously, and the percentage of oocytes showing localization at the vegetal cortex was determined. The results from this time course experiment clearly show that the MCLE localizes more rapidly than the VgLE (Fig. 4). More than 40% of the oocytes injected with the MCLE showed localization at the vegetal cortex 3 h after injection, whereas 12 h were required for a similar level of localization in oocytes injected with the VgLE. The lower frequencies of VgLE localization at early time points was not due to the inability of these oocytes to efficiently localize the VgLE as equal frequencies of localization were obtained for the MCLE and VgLE after 48 h of culture. These results show that the early pathway Xcat-2 RNA contains a bstrongerQ localization element than the late pathway Vg1 RNA. The greater ability of early pathway RNAs to compete for late localization factors (Table 2) and the faster localization of early RNAs (Fig. 4) during the late pathway suggest that early RNAs interact more favorably than late RNAs with the late localization machinery in vivo. This interpretation is consistent with our previous observations showing that the MCLE recruits more Kinesin II into the vegetal pathway than does the VgLE (Betley et al., 2004); Kinesin II is required for localization of the MCLE and VgLE to the vegetal cortex during the late pathway (Betley et al., 2004). Table 2 Competition of Vg1 localization in stage IV oocytes RNA probe Competitor RNA Molar excess Localizationa XhG–VgLE None 100Â 84% (87) XhG–VgLE VgLE 100Â 20% (15) XhG–VgLE MCLE 100Â 7% (14) XhG–VgLE XpatLE 100Â 14% (21) XhG–VgLE DSLE 100Â 3% (66) XhG–VgLE XdazLE 100Â 2% (80) XhG–VgLE DUGCAC 100Â 95% (21) a The percent localization represents the number of oocytes showing clear localization with the total number of injected oocytes shown in parentheses. Table 1 Competition of Vg1 localization in stage II oocytes RNA probe Competitor RNA Molar excess Localizationa Dig-VgLE None 300Â 80% (50) Dig-VgLE XhG 300Â 86% (58) Dig-VgLE VgLE 300Â 8% (37) Dig-VgLE MCLE 300Â 0% (36) Dig-VgLE XpatLE 150Âb 2% (39) Dig-VgLE DaE2 300Â 70% (43) Dig-VgLE DUGCAC 300Â 69% (39) a The percent localization represents the number of oocytes showing clear localization with the total number of injected oocytes shown in parentheses. b The XpatLE could only be concentrated to roughly half the concentration of the other competitor RNAs. S. Choo et al. / Developmental Biology 278 (2005) 103–117 109
  • 8. The late localization factor Vera binds specifically to early pathway localization elements The in vivo competition experiments show clearly that early and late RNAs interact with common localization factors during the late pathway. One factor implicated in the late localization pathway is Vg RBP/Vera which is an RNA binding protein that specifically binds to late pathway localization elements of Veg T and Vg1 mRNAs (Bubu- nenko et al., 2002; Deshler et al., 1997, 1998; Havin et al., 1998; Kwon et al., 2002). Vera binds preferentially to UUCAC motifs (Deshler et al., 1998) and point mutations in UUCAC motifs that reduce Vera binding impair local- ization, whereas point mutations within UUCAC motifs that do not reduce Vera binding do not impair localization during the late pathway (Kwon et al., 2002). Moreover, Vera associates with late pathway RNAs in vivo (Kress et al., 2004) and microinjection of anti-Vera antibodies reduces localization during the late pathway (Kwon et al., 2002). Since UUCAC motifs are required for localization during the early pathway (Fig. 1), we decided to explore further a potential role for Vera during the early pathway. We first sought to determine if Vera binds specifically to the same early pathway RNA localization elements that compete for Vg1 localization in vivo (Table 3). To quantitatively assess binding of Vera to early pathway RNAs we produced recombinant Vera protein which, in UV cross-linking assays, binds with similar sequence specificity to the VgLE as endogenous Vera protein does in Xenopus oocyte extracts (data not shown). The affinity of Vera for the VgLE and other RNAs was determined using filter binding assays with recombinant Vera and 32 P-labeled RNAs. Binding of Vera to the VgLE is detectable at a protein concentration of 5 nM (Fig. 5) resulting in a Kd of 11 nM (Table 3), whereas binding to similarly sized RNA fragments of the XhG gene, Vg1 open reading frame (ORF), Xcat-2 ORF, or the transcription vector used in these experiments does not occur below 200–400 nM (Fig. 5). The Kd for nonspecific RNAs is greater than 1.4 AM (Table 3) indicating that recombinant Vera has at least 100-fold higher affinity for the VgLE than it does for nonspecific RNAs. Vera also binds specifically to the same four early localization elements MCLE, XpatLE, DSLE, and XdazLE that compete for Vg1 localization factors in vivo (Table 3). The Kd for early pathway RNAs lies in a range from 17 to 47 nM (Fig. 5, Table 3) which is similar to the Kd for Vg1, but no correlation between affinities for purified Vera in vitro and pathway selectivity in vivo was apparent. Since Vera binds specifically to the localization elements of four early pathway RNAs (Fig. 5, Table 3), and deletion of UUCAC motifs blocks localization of the VgLE during the early pathway (Fig. 1), it appears that Vera may play a role in localization during the early pathway. A putative consensus binding site for Vera has previously been determined to be WYCAY (W = A or U, Y = pyrimidine) (Fig. 6A) (Deshler et al., 1998). This consensus site was determined using a short RNA probe containing four tandem copies of the UUCAC motif. Since this short RNA has a much lower affinity for Vera than the full length MCLE or VgLE (data not shown), high-affinity binding of Vera to RNA localization elements is likely to involve other interactions in addition to those mediated by WYCAY motifs. Nevertheless, we sought to test further the involve- ment of Vera during the early pathway by creating point mutations in putative Vera binding sites. To do this, we converted all nine WYCAY motifs in the VgLE to WACAY. This particular mutant design was chosen because we wanted to determine if Vera binding could be diminished without affecting localization. Such a finding would negate Table 3 Vera specifically binds early pathway localization elements RNA Kd (nM) VgLE 11 F 2 MCLE 47 F 5 XpatLE 24 F 3 DSLE 40 F 9 XdazLE 17 F 2 XhG ) N1400 Vg 1 ORF Vector Xcat-2 ORF Kd’s for the indicated RNA were determined using filter binding assays. The standard error of the mean is shown. Fig. 4. Early pathway RNAs have bstronger Q localization elements than late pathway RNAs. Oocytes were injected with either the Vg1 (VgLE) or Xcat-2 (MCLE) localization element, and a group of oocytes were placed in fixative at the indicated time points. The percentage of oocytes showing localization was determined and plotted as percent (%) localization. Note that the MCLE localizes faster than the VgLE and that after 48 h, the localization frequencies of each RNA are indistinguishable from each other. S. Choo et al. / Developmental Biology 278 (2005) 103–117110
  • 9. Fig. 5. Recombinant Vera binds specifically and with high affinity to early pathway RNAs. (A) Filter binding experiment showing increasing concentrations of Vera at the bottom and several 32 P-labeled RNAs. XhG is the Xenopus h Globin fragment commonly used as a negative control for RNA localization. In addition, fragments of the transcription vector and of the Xcat-2 and Vg1 open reading frames (ORFs) were used as negative controls for binding. (B) Binding curves for the VgLE, MCLE, XpatLE, DEADSouth LE (DSLE), and the XdazLE. The error bars represent the standard deviations in fraction bound at each protein concentration from at least five different experiments for each RNA. The Kd was determined for each RNA and is shown in Table 3. S. Choo et al. / Developmental Biology 278 (2005) 103–117 111
  • 10. an essential role for Vera in the early pathway. Since ACAY motifs are common in the localization elements of vegetal RNAs (Betley et al., 2002), we believed converting the second position to A in the WYCAY motifs would be the most likely point mutation to reveal Vera-independent localization if such localization were possible. This muta- tion, however, abolished localization of the VgLE during the early pathway since the Vg9WACAY mutant failed to localize to the mitochondrial cloud of stage I oocytes (Fig. 6B). Identical results were obtained in stage II/III oocytes consistent with a role for WYCAY motifs for the late pathway (data not shown). It was striking to observe the severe impairment of localization created by these point mutations given that larger 15–20 nucleotide deletion mutations along the entire Vg1 localization element have little or no detectable effect on localization during the late pathway (Deshler et al., 1997; Gautreau et al., 1997). When the relative Vera binding was measured for the Vg9WACAY mutant, a 40% reduction was observed when compared to the wild type VgLE (Fig. 6C). This is a smaller decrease than seen with the DaE2 mutant VgLE which has four of the nine WYCAY motifs precisely deleted, shows a 60% reduction in Vera binding (Fig. 6C) and also blocks localization during the early pathway (Fig. 1). We have analyzed a number of other point mutations in the VgLE and MCLE for Vera binding and localization during the early and late pathways. While we have not yet determined whether these mutations affect Vera binding by eliminating specific binding sites for Vera or by altering the secondary structure of these RNAs, we observe a correlation such that point mutations which incrementally reduce Vera binding show a corresponding decrease in localization during the Fig. 6. Point mutations in WYCAY motifs reduce Vera binding and RNA localization during the early pathway. (A) The RNA sequence of the 366 nucleotide VgLE is shown with WYCAY motifs underlined, and the UUCAC version of WYCAY motifs over and underlined. (B) Stage I oocytes were injected with the wild type VgLE, the DaE2 mutant VgLE which has all UUCAC motifs deleted, or the Vg9WACAY mutant which has all nine WYCAYs converted to the corresponding WACAYs. Localization of the VgLE to the mitochondrial cloud is severely impaired by mutations in these putative Vera binding sites. Scale bar = 50 Am. (C) Relative binding affinity of Vera for the VgLE and the Vg9WACAY mutant which shows a 40 percent decrease in relative Vera binding. The DaE2 mutant, which has been used to implicate Vera during the late pathway, is shown for comparison and demonstrates a 60% reduction in Vera binding in these assays. Error bar reflects the standard error of mean values generated in each binding assay. S. Choo et al. / Developmental Biology 278 (2005) 103–117112
  • 11. early and late pathways (data not shown). While these data do not prove a requirement for Vera during the localization process, they strongly suggest that Vera, a factor previously implicated only in the late pathway, is also involved in localization during the early pathway. This analysis also lends further support to our conclusion that the same cis elements are required for localization during the early and late pathways since point mutations that disrupt localization Fig. 7. Early pathway RNAs specifically associate with a perinuclear microtubule network that extends to the mitochondrial cloud of stage I oocytes. (A) Schematic diagram of a stage I oocyte showing the germinal vesicle or nucleus (N) and mitochondrial cloud (M). An optical section tangential to the nucleus and mitochondrial cloud is depicted as a parallelogram and is similar to the ones shown in (B) and (C, top row only). (An) and (Vg) signify the future animal– vegetal axis. (B) Visualization of the perinuclear microtubule (MT) network (green) in a stage I oocyte. Perinuclear microtubule bundles (arrowheads) radially oriented with respect to the mitochondrial cloud and perinuclear microtubule enriched foci (arrows) are indicated. (C) Double labeling of microtubules (green) and fluorescently MCLE labeled RNA (red) 1 h (top row) or 6 h (middle row) after injection of the RNA. Note that after 1 h little or no MCLE RNA is detected in the mitochondrial cloud (M), but many RNA particles can be seen associated with perinuclear bundles of microtubules (arrowheads) extending toward the mitochondrial cloud. After 6 h, several MCLE particles are associated with perinuclear microtubule foci (arrow) and extensive localization of the MCLE to the mitochondrial cloud is apparent (large arrowhead). Within the mitochondrial cloud, however, the RNA becomes dispersed and the large particles are not apparent. At this same time point, the XhG RNA (bottom row) shows no accumulation of particles on perinuclear foci (arrows) or within the mitochondrial cloud (large arrowhead). At the 6-h time points, optical sections are oriented along the future animal–vegetal axis and run through the mitochondrial cloud and nucleus. Scale bar = 10 Am. S. Choo et al. / Developmental Biology 278 (2005) 103–117 113
  • 12. during the early pathway also impair localization during the late pathway. Xcat-2 RNA associates with a perinuclear microtubule network that extends to the mitochondrial cloud of stage I oocytes The data obtained in our competition, biochemical, and mutational analyses uniformly point to the existence of common localization machinery participating in the early and late pathway. However, previous reports employing the use of agents that disrupt microtubules have suggested that microtubules are required for the late, but not for the early pathway (Kloc and Etkin, 1995; Kloc et al., 1996; Yisraeli et al., 1990). This has contributed to an alternative view implicating distinct mechanisms for RNA localization during the early and late pathways. Because stage I oocytes possess a population of acetylated and presumably stable microtubules (Gard, 1999) and the early pathway local- ization element of Xcat-2 can recruit a microtubule-depend- ent motor in stage II oocytes (Betley et al., 2004), we decided to reexamine a potential role for microtubules during the early pathway. To do this, we assessed the spatial relationship between microtubules, the mitochondrial cloud, and fluorescently labeled MCLE RNA injected into stage I oocytes. Confocal imaging of stage I oocytes labeled with an anti- a-tubulin antibody showed that microtubules are distributed throughout the cytoplasm and are enriched in the cortical region as previously reported (Gard, 1999). However, optical sections acquired at the periphery of the nucleus (Fig. 7A) revealed a network consisting of microtubule bundles that are wrapped around the nucleus and intersect at dense patches or foci (Fig. 7B). These foci are easily distinguished from the surrounding network due to their association with numerous radially oriented microtubule bundles and their enhanced staining. Interestingly, this perinuclear network of microtubule fibers extends into the mitochondrial cloud which itself is surrounded by a ring of microtubules (Fig. 7B). To characterize the distribution of microtubules and RNAs utilizing the early pathway, we injected fluorescently labeled MCLE or XhG RNA into stage I oocytes and visualized the microtubules using immunocytochemistry. Oocytes were fixed 1 or 6 h after injection of RNA. Remarkably, at both time points, we observed large particles with the MCLE that associated with bundles and foci of the microtubule network and seemed to follow the network into the mitochondrial cloud. The XhG RNA did not show any apparent association with this network (Fig. 7C). After 1 h much of the MCLE RNA remains dispersed throughout the cytoplasm, and the mitochondrial cloud does not yet show labeling by localized RNA. However, at this time, we observe large MCLE RNA particles or granules that colocalize with perinuclear foci and microtubule fibers extending into the mitochondrial cloud (Fig. 7C, top row). At the later time point (6 h), the MCLE has accumulated in the mitochondrial cloud (Fig. 7C, middle row), whereas the XhG RNA does not (Fig. 7C, bottom row). Interestingly, once the MCLE RNA enters the mitochondrial cloud, it becomes diffusely distributed, and large particles are no longer visible. This suggests that the MCLE RNA granules dissociate into smaller complexes once they reach their final destination in stage I oocytes as has been observed with the myelin basic protein mRNA in oligodendrocytes (Ainger et al., 1993). In addition to association with microtubule bundles of the perinuclear network and localization in the mitochondrial cloud, we also observe an accumulation of MCLE particles on perinuclear foci that interconnect microtubule bundles of the network. One striking example can be seen in Fig. 7C (middle row), where a number of MCLE particles are clustered on one of these foci. No accumulation of granules on perinuclear foci was detected in oocytes injected with XhG RNA (Fig. 7C, bottom row). This organization of the microtubule network and specific association of the MCLE RNA with it in stage I oocytes suggest that microtubule dependent transport may play a role in RNA localization during the early pathway as it does during the late pathway (Betley et al., 2002; Kloc and Etkin, 1995; Yisraeli et al., 1990). Discussion In Xenopus, the spatial distributions of different mRNAs in growing oocytes have indicated that at least two temporally defined pathways exist for RNA local- ization to the vegetal cortex (Forristall et al., 1995; Kloc and Etkin, 1995). Here, we have designed experiments to explore whether distinct or common cellular machinery recognizes and directs RNA localization during the early and late pathway. We have shown that a late pathway RNA, Vg1, can localize during the early pathway and that the same short (5 nt) cis motifs required for late localization are also required for early localization and vice versa. Competition experiments reveal that the local- ization elements of four different early pathway RNAs including Xcat-2, Xpat, Xdazl, and DEADSouth specifi- cally compete for late localization factors in vivo. In addition, these same early pathway RNAs are bound specifically by the late localization factor Vg RBP/Vera, and point mutations that reduce Vera binding impair localization during the early and late pathways. We also show that the Xcat-2 RNA localization element forms particles which associate with a microtubule network that surrounds the nucleus and extends into the mitochondrial cloud of stage I oocytes suggesting that directed transport may play a role in the early pathway as it does during the late pathway. Taken together, these results support a model in which a common set of cellular factors participates in the localization of vegetal RNAs during the early and late pathways. While this model in no way excludes the S. Choo et al. / Developmental Biology 278 (2005) 103–117114
  • 13. existence of pathway-specific factors, it suggests that the early and late pathways may have evolved as a conse- quence of distinct RNAs acquiring the ability to interact differentially with common or basal RNA localization machinery that remains active throughout stages I–IV of oogenesis. The role of short cis motifs in RNA localization to the vegetal cortex RNA localization elements that specify localization to the vegetal cortex contain a variety of short repeated motifs that are important for signaling interactions with cellular machinery required for the localization process (Betley et al., 2002; Bubunenko et al., 2002; Deshler et al., 1997, 1998; Kwon et al., 2002). These motifs may provide binding sites for RNA binding proteins or they may participate in intramolecular or intermolecular RNA–RNA interactions required for association with the localization machinery. Moreover, it is possible that individual motifs participate in a combination of these functions during various aspects of the localization process. In these studies, we show that WYCAY motifs are required for localization during the early as well as the late pathway. Since WYCAY is a consensus binding site for Vg RBP/Vera (Deshler et al., 1997) and point mutations that disrupt Vera binding impair localization, our findings suggest that at least one function of WYCAY motifs is to bind Vera which is known to be expressed during the early and late pathways (Deshler et al., 1997). We have also shown that the UGCAC motif is required for localization during both pathways (Fig. 2) (Betley et al., 2002), but this motif is not bound specifically by Vera in vitro as previously reported (Betley et al., 2002; Deshler et al., 1998). In fact, the DUGCAC MCLE mutant binds Vera with an affinity that is indistinguishable from the wild type MCLE (Kd = 46 and 47 nM, respectively). Therefore, the UGCAC motif may interact with another protein component of the localization machinery. Alternatively, or possibly in addition, UGCAC motifs may mediate RNA–RNA base pairing interactions required for localization. This is plausible because the first four nucleotides of UGCAC are palindromic giving them the ability to base pair with each other. Furthermore, RNA–RNA base pairing is thought to be important for the microtubule dependent localization of bicoid RNA in Drosophila (Ferrandon et al., 1997; Wagner et al., 2001). Interestingly, the double-stranded RNA bind- ing protein, Staufen, mediates localization of bicoid mRNA in Drosophila oocytes (St. Johnston et al., 1991), and Xenopus Staufen1 has recently been implicated in the localization of RNAs in Xenopus oocytes (Yoon and Mowry, 2004). Thus, it is possible that RNA–RNA interactions also play a role in localizing RNAs in Xenopus oocytes. Hence, more work is required to determine whether UGCAC motifs mediate RNA–protein or RNA–RNA interactions required for the localization process. What factors limit the RNA localization process in vivo? Our in vivo competition results may provide a clue about the localization factors that limit the localization process in oocytes. The concentration of RNA required to compete the localization machinery in stage I/II oocytes (300 Am diameter) is approximately 6 AM. However, the concen- tration of Vera in these oocytes is roughly 4.5 times higher (27 F 4 AM) as determined by Western blot analysis (data not shown) consistent with amounts reported to be present in stage IVoocytes (Kwon et al., 2002). Such a high level of endogenous Vera may explain why antibody inhibition experiments designed to bknock outQ Vera function only partially disrupt the localization process in oocytes (Kwon et al., 2002). Therefore, Vera may not be limiting for local- ization in vivo. Another factor implicated in vegetal RNA localization in Staufen1 (Yoon and Mowry, 2004). The endogenous levels of Xenopus Staufen may be lower than that of Vera. We have recently cloned Staufen2 in Xenopus and shown that its concentration is 10 F 1 AM in stage I/II oocytes (Z. Chen, manuscript in preparation) which is similar to the levels required to saturate the localization machinery in vivo. Thus, it is possible that XStau2 may be limiting for localization in oocytes as it appears to be for the localization of bulk mRNA into the dendrites of neurons (Tang et al., 2001). Identifying the concentration-limited components of the localization machinery will be an important first step for understanding the regulation of the RNA localization process. Previous work has suggested that the Vg1 localization element forms an RNA–protein complex consisting of at least six RNA binding proteins that may be required to initiate the RNA localization process in oocytes (Mowry, 1996). Since motifs that bind Vera (UUCAC) (Kwon et al., 2002) and motifs that do not bind Vera (UGCAC) (Fig. 3 and Table 1) are both required for competing localization in vivo, it is possible that RNA localization elements must maintain their capacity to form a bpre- localization complexQ in order to compete for the local- ization machinery in vivo. The formation of such a putative multi-protein complex could involve cooperative interactions such that mutating specific binding sites for Vera or other critical components would be sufficient to render the mutant RNA incapable of assembling the complex and competing for the localization machinery in vivo. Evidence for the existence of Vera and Staufen containing RNP complexes in vivo has recently been provided by co-immunoprecipitation assays of Vera, XStau1 and the Vg1 RNA in stage III oocytes (Kress et al., 2004; Yoon and Mowry, 2004). The formation of this bpre-localizationQ complex would then lead to the recruit- ment of molecular motors such as heterotrimeric Kinesin II for transport to the cortex (Betley et al., 2004). In this scenario, mutations that disrupt formation of the pre- localization complex would affect all phases of this presumably multi-step localization process. S. Choo et al. / Developmental Biology 278 (2005) 103–117 115
  • 14. Directed versus diffusion-mediated RNA transport The arrangement of the microtubule network with respect to the mitochondrial cloud and the specific association of Xcat-2 RNA granules with it (Fig. 7) indicate that a directed transport mechanism may function during the early pathway. However, treatment of stage I oocytes with nocodazole has little or no effect on the distribution of early pathway RNAs (Kloc and Etkin, 1995; Kloc et al., 1996) leading some to propose that directed RNA transport does not play a role during the early pathway (Chang et al., 2004). We believe that two factors may contribute to this paradox. First, it has been shown that acetylated microtubules exist in the perinuclear region of stage I oocytes (Gard, 1999), and while nocodazole disrupts the organization of microtubules in stage I oocytes (data not shown), the detection of a subpopulation of microtubules could be obscured in nocodazole treated oocytes. Interestingly, a stable and acetylated population of microtubules has been found to play a role in oocyte specification in Drosophila (Roper and Brown, 2004). Secondly, diffusion must also play an important role in the localization process. In fact, interactions between Kinesin II and vegetal RNAs appear to be quite dynamic in vivo (Betley et al., 2004). Thus, due to the small size of stage I oocytes, diffusion may be sufficient for the localization of injected RNAs to the mitochondrial cloud. However, diffusion would not be expected to be sufficient for the localization of vegetal RNAs in larger stage III/IV during the late pathway. The consideration of stable microtubules and/or the possibility of diffusion-mediated localization of injected RNAs in small stage I oocytes may help explain why previous work has invoked two distinct mechanisms for RNA localization during the early and late pathways; these studies relied exclusively on detecting the sensitivity of RNA localization to microtubule depolymerizing drugs in stage I (approximately 200 Am diameter) and stage IV (approximately 800 Am diameter) oocytes. In any case, given that the MCLE forms RNA granules that associate with microtubules entering the mitochondrial cloud (Fig. 7), we now believe that diffusion and directed transport work together in stage I as well as stage III/IV oocytes to mediate the selective localization of vegetal RNAs during the early and late pathways, respectively, in Xenopus oocytes. What determines when distinct RNAs become localized during oogenesis? We have shown that the Vg1 localization element is capable of localizing to the mitochondrial cloud if injected into stage I oocytes. What then prevents endogenous Vg1 from localizing to the mitochondrial cloud of stage I oocytes? One possibility is that negatively acting cis- regulatory elements in either the VgLE or other regions of Vg1 serve to attenuate or inhibit localization until stage II of oogenesis. Another possibility is based on our observed kinetic differences in the localization of and early and late pathway localization elements. It is clear that the MCLE localizes faster than the VgLE during the late pathway in stage II oocytes (Fig. 4). It is also known that the early pathway RNA, Xlsirt, is first to localize to the cortex when co-injected with the VgLE into stage II oocytes (Kloc et al., 1996). Taken together, these observations show that early RNAs contain stronger localization elements than late pathway RNAs. Therefore, at endogenous concentrations, Vg1 may not sufficiently compete for the necessary local- ization machinery until early pathway RNAs become localized in the mitochondrial cloud or anchored at the vegetal cortex. Interestingly, in very early stage I oocytes (approximately 50 Am diameter), endogenous Xcat-2 is distributed throughout the cytoplasm (Zhou and King, 1996a) as is endogenous Vg1 mRNA in later stage I oocytes (Forristall et al., 1995; Kloc and Etkin, 1995). This combined with results presented in this work suggests that a similar pathway may be followed for both endogenous RNAs, but that each follows this pathway at slightly different times during oogenesis. While our work does not exclude the possibility of pathway-specific RNA localization factors, we have pro- vided evidence for the existence of common or basal localization machinery that is involved in both the early and late pathways. This may provide an economical mean by which oocytes generate a high number of temporally defined pathways without evolving novel machinery for each one. In fact, it has been shown by double-labeling in situ hybridization that different early pathway RNAs are actually localized to distinct domains within the mitochon- drial cloud of stage I oocytes suggesting that different early RNAs utilize distinct temporally defined pathways (Kloc and Etkin, 1995). Moreover, we have identified an RNA, Xlerk, which clearly uses a pathway that is intermediate between early and late RNAs (Betley et al., 2002). In this model, relative differences in the affinity of specific RNAs for basal components of the localization machinery could determine when a transcript is localized during oogenesis. Interestingly, two recent reports have shown that early and late RNAs interact specifically with a number of RNA binding proteins that are distinct from Vera (Chang et al., 2004; Claussen et al., 2004) supporting the notion that common machinery is utilized during the early and late pathways. The identification and biochemical character- ization of additional factors involved in the localization process will be required to fully understand how such a mechanism is regulated to establish polarized distributions of distinct factors in the egg. Acknowledgments This work was supported by the National Institutes of Health Grant R01 GM59682 to J.O.D. S. Choo et al. / Developmental Biology 278 (2005) 103–117116
  • 15. References Ainger, K., Avossa, D., Morgan, F., Hill, S.J., Barry, C., Barbarese, E., Carson, J.H., 1993. Transport and localization of exogenous myelin basic protein mRNA microinjected into oligodendrocytes. J. Cell Biol. 123, 431–441. Alarcon, V.B., Elinson, R.P., 2001. RNA anchoring in the vegetal cortex of the Xenopus oocyte. J. Cell Sci. 114, 1731–1741. Allen, L., Kloc, M., Etkin, L.D., 2003. Identification and characterization of the Xlsirt cis-acting RNA localization element. Differentiation 71, 311–321. Betley, J.N., Frith, M.C., Graber, J.H., Choo, S., Deshler, J.O., 2002. A ubiquitous and conserved signal for RNA localization in chordates. Curr. Biol. 12, 1756–1761. Betley, J.N., Heinrich, B., Vernos, I., Sardet, C., Prodon, F., Deshler, J.O., 2004. Kinesin II meditates Vg1 mRNA transport in Xenopus oocytes. Curr. Biol. 14, 219–224. Bubunenko, M., Kress, T.L., Vempati, U.D., Mowry, K.L., King, M.L., 2002. A consensus RNA signal that directs germ layer determinants to the vegetal cortex of Xenopus oocytes. Dev. Biol. 248, 82–92. Chang, P., Torres, J., Lewis, R.A., Mowry, K.L., Houliston, E., King, M.L., 2004. Localization of RNAs to the mitochondrial cloud in Xenopus oocytes through entrapment and association with endoplasmic retic- ulum. Mol. Biol. Cell 15, 4669–4681. Claussen, M., Horvay, K., Pieler, T., 2004. Evidence for overlapping, but not identical, protein machineries operating in vegetal RNA localization along early and late pathways in Xenopus oocytes. Development 131, 4263–4273. Cote, C.A., Gautreau, D., Denegre, J.M., Kress, T.L., Terry, N.A., Mowry, K.L., 1999. A Xenopus protein related to hnRNP I has a role in cytoplasmic RNA localization. Mol. Cell 4, 431–437. Deshler, J.O., Highett, M.I., Schnapp, B.J., 1997. Localization of Xenopus Vg1 mRNA by Vera protein and the endoplasmic reticulum. Science 276, 1128–1131. Deshler, J.O., Highett, M.I., Abramson, T., Schnapp, B.J., 1998. A highly conserved RNA-binding protein for cytoplasmic mRNA localization in vertebrates. Curr. Biol. 8, 489–496. Dumont, J.N., 1972. Oogenesis in Xenopus laevis (Daudin). I. Stages of oocyte development in laboratory maintained animals. J. Morphol. 136, 153–179. Ferrandon, D., Koch, I., Westhof, E., Nusslein-Volhard, C., 1997. RNA– RNA interaction is required for the formation of specific bicoid mRNA 3V UTR-STAUFEN ribonucleoprotein particles. EMBO J. 16, 1751–1758. Forristall, C., Pondel, M., Chen, L., King, M.L., 1995. Patterns of localization and cytoskeletal association of two vegetally localized RNAs, Vg1 and Xcat-2. Development 121, 201–208. Gard, D.L., 1999. Confocal microscopy and 3-D reconstruction of the cytoskeleton of Xenopus oocytes. Microsc. Res. Tech. 44, 388–414. Gautreau, D., Cote, C.A., Mowry, K.L., 1997. Two copies of a subelement from the Vg1 RNA localization sequence are sufficient to direct vegetal localization in Xenopus oocytes. Development 124, 5013–5020. Havin, L., Git, A., Elisha, Z., Oberman, F., Yaniv, K., Schwartz, S.P., Standart, N., Yisraeli, J.K., 1998. RNA-binding protein conserved in both microtubule- and microfilament-based RNA localization. Genes Dev. 12, 1593–1598. Heasman, J., Wessely, O., Langland, R., Craig, E.J., Kessler, D.S., 2001. Vegetal localization of maternal mRNAs is disrupted by VegT depletion. Dev. Biol. 240, 377–386. Houston, D.W., Zhang, J., Maines, J.Z., Wasserman, S.A., King, M.L., 1998. A Xenopus DAZ-like gene encodes an RNA component of germ plasm and is a functional homologue of Drosophila boule. Develop- ment 125, 171–180. Hudson, C., Woodland, H.R., 1998. Xpat, a gene expressed specifically in germ plasm and primordial germ cells of Xenopus laevis. Mech. Dev. 73, 159–168. Kloc, M., Etkin, L.D., 1994. Delocalization of Vg1 mRNA from the vegetal cortex in Xenopus oocytes after destruction of Xlsirt RNA. Science 265, 1101–1103. Kloc, M., Etkin, L.D., 1995. Two distinct pathways for the localization of RNAs at the vegetal cortex in Xenopus oocytes. Development 121, 287–297. Kloc, M., Etkin, L.D., 1998. Apparent continuity between the messenger transport organizer and late RNA localization pathways during oogenesis in Xenopus. Mech. Dev. 73, 95–106. Kloc, M., Etkin, L.D., 1999. Analysis of localized RNAs in Xenopus Oocytes. In: Richter, J.D. (Ed.), A Comparative Methods Approach to the Study of Oocytes and Embryos. Oxford Univ. Press, Inc., New York, pp. 256–278. Kloc, M., Spohr, G., Etkin, L.D., 1993. Translocation of repetitive RNA sequences with the germ plasm in Xenopus oocytes. Science 262, 1712–1714. Kloc, M., Larabell, C., Etkin, L.D., 1996. Elaboration of the messenger transport organizer pathway for localization of RNA to the vegetal cortex of Xenopus oocytes. Dev. Biol. 180, 119–130. Kloc, M., Bilinski, S., Pui-Yee Chan, A., Etkin, L.D., 2000. The targeting of Xcat2 mRNA to the germinal granules depends on a cis-acting germinal granule localization element within the 3VUTR. Dev. Biol. 217, 221–229. Kloc, M., Bilinski, S., Chan, A.P., Allen, L.H., Zearfoss, N.R., Etkin, L.D., 2001. RNA localization and germ cell determination in Xenopus. Int. Rev. Cytol. 203, 63–91. Kress, T.L., Yoon, Y.J., Mowry, K.L., 2004. Nuclear RNP complex assembly initiates cytoplasmic RNA localization. J. Cell Biol. 165, 203–211. Ku, M., Melton, D.A., 1993. Xwnt-11: a maternally expressed Xenopus wnt gene. Development 119, 1161–1173. Kwon, S., Abramson, T., Munro, T.P., John, C.M., Kohrmann, M., Schnapp, B.J., 2002. UUCAC and Vera dependent localization of VegT RNA in Xenopus oocytes. Curr. Biol. 12, 558–564. Lewis, R.A., Kress, T.L., Cote, C.A., Gautreau, D., Rokop, M.E., Mowry, K.L., 2004. Conserved and clustered RNA recognition sequences are a critical feature of signals directing RNA localization in Xenopus oocytes. Mech. Dev. 121, 101–109. MacArthur, H., Houston, D.W., Bubunenko, M., Mosquera, L., King, M.L., 2000. DEADSouth is a germ plasm specific DEAD-box RNA helicase in Xenopus related to eIF4A. Mech. Dev. 95, 291–295. Mowry, K.L., 1996. Complex formation between stage-specific oocyte factors and a Xenopus mRNA localization element. Proc. Natl. Acad. Sci. U. S. A. 93, 14608–14613. Palacios, I.M., St. Johnston, D., 2001. Getting the message across: the intracellular localization of mRNAs in higher eukaryotes. Annu. Rev. Cell Dev. Biol. 17, 569–614. Roper, K., Brown, N.H., 2004. A spectraplakin is enriched on the fusome and organizes microtubules during oocyte specification in Drosophila. Curr. Biol. 14, 99–110. St. Johnston, D., Beuchle, D., Nusslein-Volhard, C., 1991. Staufen, a gene required to localize maternal RNAs in the Drosophila egg. Cell 66, 51–63. Tang, S.J., Meulemans, D., Vazquez, L., Colaco, N., Schuman, E., 2001. A role for a rat homolog of Staufen in the transport of RNA to neuronal dendrites. Neuron 32, 463–475. Wagner, C., Palacios, I., Jaeger, L., St. Johnston, D., Ehresmann, B., Ehresmann, C., Brunel, C., 2001. Dimerization of the 3VUTR of bicoid mRNA involves a two-step mechanism. J. Mol. Biol. 313, 511–524. Yisraeli, J.K., Sokol, S., Melton, D.A., 1990. A two-step model for the localization of maternal mRNA in Xenopus oocytes: involvement of microtubules and microfilaments in the translocation and anchoring of Vg1 mRNA. Development 108, 289–298. Yoon, Y.J., Mowry, K.L., 2004. Xenopus Staufen is a component of a ribonucleoprotein complex containing Vg1 RNA and Kinesin. Develop- ment 131, 3035–3045. Zhou, Y., King, M.L., 1996a. Localization of Xcat-2 RNA, a putative germ plasm component, to the mitochondrial cloud in Xenopus stage I oocytes. Development 122, 2947–2953. Zhou, Y., King, M.L., 1996b. RNA transport to the vegetal cortex of Xenopus oocytes. Dev. Biol. 179, 173–183. S. Choo et al. / Developmental Biology 278 (2005) 103–117 117