ArticlePDF Available

Xenopus oocyte prophase I meiotic arrest is released independently from a decrease in cAMP levels or PKA activity

Authors:
  • Weill Cornell Medicine in Qatar

Abstract and Figures

Vertebrate oocytes arrest at prophase of meiosis I due to high levels of cAMP and PKA activity. In Xenopus progesterone is believed to release meiotic arrest by inhibiting adenylate cyclase, lowering cAMP levels, and repressing protein kinase A (PKA). However the exact timing and extent of cAMP decrease is unclear with conflicting reports in the literature. Using various in vivo reporters for cAMP and PKA at the single cell level in real time, we fail to detect any significant changes in cAMP or PKA in response to progesterone. More interestingly, there was no correlation between the levels of PKA inhibition and the release of meiotic arrest. Furthermore, we devised condition where meiotic arrest could be released in the presence of sustained high levels of cAMP. Consistently, lowering endogenous cAMP levels by over 65% for prolonged time periods failed to induce spontaneous maturation. These results argue that the release of oocyte meiotic arrest in Xenopus occurs independently from lowering either cAMP levels or PKA activity, but rather through a parallel cAMP-PKA-independent pathway.
Content may be subject to copyright.
© 2016. Published by The Company of Biologists Ltd.
Xenopus oocyte prophase I meiotic arrest is released independently from a decrease in
cAMP levels or PKA activity
Nancy Nader, Raphael Courjaret, Maya Dib, Rashmi P. Kulkarni and Khaled Machaca*
Department of Physiology and Biophysics, Weill Cornell Medicine Qatar, Education City –
Qatar Foundation, Doha, Qatar
Keywords: progesterone, oocyte maturation, cAMP, Xenopus, PKA
* Correspondence:
Khaled Machaca
Weill Cornell Medicine Qatar
Education City - Qatar Foundation– Doha, Qatar
E-mail: khm2002@qatar-med.cornell.edu
Tel: +1-646-797-3222
Fax: +974-4492-8422
Summary Statement: Progesterone releases Xenopus oocyte meiotic arrest without any reduction in
cAMP levels or PKA activity.
Development • Advance article
http://dev.biologists.org/lookup/doi/10.1242/dev.136168Access the most recent version at
Development Advance Online Articles. First posted online on 27 April 2016 as 10.1242/dev.136168
ABSTRACT
Vertebrate oocytes arrest at prophase of meiosis I due to high levels of cAMP and PKA
activity. In Xenopus progesterone is believed to release meiotic arrest by inhibiting adenylate
cyclase, lowering cAMP levels, and repressing protein kinase A (PKA). However the exact
timing and extent of cAMP decrease is unclear with conflicting reports in the literature.
Using various in vivo reporters for cAMP and PKA at the single cell level in real time, we fail
to detect any significant changes in cAMP or PKA in response to progesterone. More
interestingly, there was no correlation between the levels of PKA inhibition and the release of
meiotic arrest. Furthermore, we devised condition where meiotic arrest could be released in
the presence of sustained high levels of cAMP. Consistently, lowering endogenous cAMP
levels by over 65% for prolonged time periods failed to induce spontaneous maturation.
These results argue that the release of oocyte meiotic arrest in Xenopus occurs independently
from lowering either cAMP levels or PKA activity, but rather through a parallel cAMP-PKA-
independent pathway.
Development • Advance article
INTRODUCTION
Full-grown vertebrate oocytes arrest in a G2-like state at prophase of meiosis I for prolonged
periods of time during which the oocyte grows and stores macromolecular components
needed for future development (Smith, 1989; Voronina and Wessel, 2003). Before ovulation,
oocytes resume meiosis and arrest in metaphase of meiosis II, in a process termed ‘oocyte
maturation’. Oocyte maturation encompasses drastic cellular remodeling, in line with meiosis
progression, which prepares the egg for fertilization. This is characterized by the dissolution
of the nuclear envelope (referred to as germinal vesicle breakdown (GVBD), extrusion of the
first polar body and chromosome condensation (Bement and Capco, 1990; Nader et al., 2013;
Sadler and Maller, 1985; Smith, 1989; Voronina and Wessel, 2003).
Progesterone (P4) is the classical steroid used to mature Xenopus oocytes in vitro. Though a
role for the classical nuclear steroid receptor cannot be ruled out (Bayaa et al., 2000; Tian et
al., 2000), P4-induced maturation is thought to be achieved primarily through a non-classical
plasma membrane P4 receptor (mPR) (Josefsberg Ben-Yehoshua et al., 2007). It is
generally accepted that, upon hormone binding, mPR inhibits adenylate cyclase (AC)
causing a transient dip in cAMP levels, which is believed to act as the trigger to release
oocyte meiotic arrest by lowering protein kinase A (PKA) activity (Cicirelli and Smith, 1985;
Maller et al., 1979).
There is broad consensus that meiotic arrest in vertebrate oocytes is maintained by high levels
of cAMP-PKA, with the most compelling evidence coming from studies of the constitutively
active orphan GPCR, GRP3, in mouse oocytes (Freudzon et al., 2005; Mehlmann et al.,
2004). Knockout of GPR3 leads to premature oocyte maturation in antral follicles and
sterility. In Xenopus oocytes injection of the catalytic subunit of PKA (PKAc) inhibits P4-
induced oocyte maturation (Daar et al., 1993; Eyers et al., 2005; Matten et al., 1994).
Interestingly however even injection of a catalytically dead PKA mutant inhibits maturation
(Eyers et al., 2005; Schmitt and Nebreda, 2002), although it has been argued that this is due
to residual PKA activity (Eyers et al., 2005). Furthermore, phosphodiesterase (PDE)
inhibitors (Andersen et al., 1998; Han et al., 2006), and treatment with cholera toxin that
activates AC, repress maturation (Huchon et al., 1981; Mulner et al., 1979; Schorderet-
Slatkine et al., 1978). Inhibition of AC accelerated P4-induced oocyte maturation (Sadler et
al., 1986). Consistently, injection of PDE, the PKA regulatory subunit (PKAr), or the PKA
pseudo-substrate inhibitor (PKI), all release meiotic arrest (Andersen et al., 1998; Daar et al.,
1993; Eyers et al., 2005). However, others found that inhibition of PKA activity using H-89
Development • Advance article
or PKI injection failed to stimulate oocyte maturation (Noh and Han, 1998). Furthermore,
induction of a drop in cAMP levels through stimulation of exogenously expressed Gi-coupled
serotonin receptor did not stimulate oocyte maturation (Noh and Han, 1998).
Whereas some studies detect a 10-60% drop in cAMP levels 15 seconds up to 6 hours after
P4 addition (Bravo et al., 1978; Cicirelli and Smith, 1985; Gelerstein et al., 1988; Maller et
al., 1979; Mulner et al., 1979; Nader et al., 2014; Sadler and Maller, 1981; Schorderet-
Slatkine et al., 1982), others failed to detect any changes in cAMP levels in response to P4
(Noh and Han, 1998; O'Connor and Smith, 1976; Schutter et al., 1975; Thibier et al., 1982).
Furthermore, some studies found that a decrease in cAMP levels was not sufficient to induce
spontaneous maturation, and that an increase in cAMP accelerates maturation (Gelerstein et
al., 1988; Noh and Han, 1998). Consistently, exposing oocytes to P4 for <30 min induced a
decrease in cAMP but did not induce oocyte maturation, which required >30 min incubation
with P4 (Nader et al., 2014). Hence there is strong support for the cAMP-PKA axis
maintaining the prolonged oocyte prophase I meiotic arrest. However, the evidence
supporting a drop of cAMP/PKA activity to release meiotic arrest is poor, with conflicting
reports regarding the timing and extent of the cAMP decrease in response to P4.
Here we use multiple approaches, including cAMP and PKA sensitive channels and FRET
sensors, to follow cAMP and PKA activity in individual oocytes in real time, and could not
detect any change during the induction of oocyte maturation. Unexpectedly, increasing or
buffering cAMP levels in the oocytes did not correlate with the ability of P4 to release
meiotic arrest. Based on these findings we proposed an alternative model where P4 stimulates
oocyte maturation independent of a decrease in cAMP-PKA.
Development • Advance article
RESULTS
cAMP decrease in levels does not correlate with the release of meiotic arrest
We and others have previously shown a decrease in cAMP levels after P4 addition using an
ELISA-based assay in a batch of oocytes. However the early cAMP drop was not sufficient to
induce maturation, since exposure to P4 from 15 sec to 10 min was not sufficient to induce
oocyte maturation, rather oocytes required a minimum of 30 min exposure to P4 to induce
significant oocyte maturation (Nader et al., 2014). However, oocytes may require an
integrated sustained level of lower cAMP to inhibit PKA activity and relieve meiotic arrest.
To explore this possibility, we used the ELISA assay to measure cAMP levels in batches of
10 oocytes at time points ranging from seconds (Fig. 1B, left panel), up to 60 minutes (Fig.
1B, right panel) after P4 addition. Briefly, Oocytes were treated transiently with 10-5 M P4 or
ethanol as the carrier control for the indicated time points after which 10 oocytes were
collected for the cAMP-ELISA assay and the rest of the oocytes were washed twice with the
culture media (L15) to remove P4, and oocytes allowed to mature overnight.
When compared to oocytes treated with ethanol, cAMP levels showed a significant drop at 5,
15 sec, 1, 4, 6, 10, 12, 30 and 60 min after P4 (Fig. 1B and 1C). Interestingly, cAMP levels at
1 min and 30 min after P4 where not significantly different from each other (p=0.5071) (Fig.
1C), however oocytes failed to mature after a 1 min exposure to P4, yet achieved full
maturation after 30 min exposure to P4 (Fig. 1D). These results suggest that the cAMP dip
observed within 30 min of P4 treatment is not the trigger for maturation.
Membrane channels do not detect changes in cAMP or PKA in response to P4
An important limitation of the correlation between cAMP levels and maturation is that the
analysis was performed on a batch of 10 oocytes, as typically reported in the literature. This
is problematic because the early stages of oocyte maturation (before GVBD) are not
synchronized at the single oocyte level; rather individual oocytes reach the GVBD stage with
disparate time courses. Therefore, changes in cAMP could be more pronounced at the single
oocyte level but are averaged out in a cell population, which may account for the absence of a
correlation between the cAMP levels and maturation.
We therefore devised approaches to follow cAMP and PKA levels in individual oocytes in
real time after P4 treatment. For that purpose we used two membrane channels as sensors for
cAMP and its effector PKA: the Cyclic Nucleotide Gated channel (CNG), a cAMP-gated
Development • Advance article
Ca2+ channel; and the Cystic Fibrosis Transmembrane Regulator (CFTR), a PKA-activated
Cl- channel (Biel and Michalakis, 2009; Hanrahan et al., 1996). Both channels have been
expressed and well characterized in Xenopus oocytes (Bear et al., 1991; Nache et al., 2012;
Webe et al., 2001; Young and Krougliak, 2004). A rise in cAMP gates CNG channels, which
permeate both mono- and divalent cations including Ca2+. Ca2+ in turn stimulates endogenous
Ca2+-Activated Chloride Channels (CaCC) thus amplifying the signal (Fig. 2A and 2B).
Expressing CNG in oocytes was associated with a large CaCC (Fig. 2B), indicating that the
channel is gated at resting cAMP levels. To test the sensitivity of CNG to changes in cAMP
we inhibited AC with 2’,5’-Dideoxyadenosine (DDA), to lower basal cAMP levels (Sadler
and Maller, 1983). DDA treatment significantly (p=0.021) decreased basal CNG-CaCC
current (ICNG-CaCC) (Fig. 2C), indicating that CNG effectively detects a drop in cAMP.
However, when CNG-expressing oocytes were treated with P4, we failed to detect any
significant differences in ICNG-CaCC, up to 30 min after P4 as compared to ethanol (Fig. 2D).
This suggests that basal cAMP levels, at least in the sub-plasma membrane domain detected
by the CNG channel, were unchanged up to 30 min after P4 treatment. However, the CNG
channel clone used is gated by both cGMP and cAMP (Young and Krougliak, 2004), so the
lack of response could be because the effect was masked by cGMP in the oocyte.
To test for changes in PKA activity we used the PKA-gated Cl- channel, CFTR, to follow
PKA activity in real time in the oocyte. In order to test whether CFTR can detect changes in
PKA activity, we injected db-cAMP to activate PKA, this showed a significant and
reproducible, although short lived, induction of the CFTR current (ICFTR) (Supplemental Fig.
1A-B). Inhibition of PKA with PKI under these condition, resulted in a significant decrease
of ICFTR (p=0.015) (Supplemental Fig. 1C and 1D). This shows that CFTR detects changes in
PKA activity in real time in the oocyte.
ICFTR showed significant run-down within a minute of establishing the voltage clamp (Fig.
2E), as previously reported (Button et al., 2001). To minimize ICFTR rundown, we pre-treated
oocytes with 3-isobutyl-1-methylxanthine (IBMX) (Ramu et al., 2007), a phosphodiesterase
inhibitor, to raise cAMP and activate PKA (Sadler and Maller, 1987; Schorderet-Slatkine and
Baulieu, 1982). IBMX pre-treatment significantly (p=0.0002) increased ICFTR, and limited its
rundown to maintain a significant current for up to 25 min (Fig. 2F). Though we cannot rule
out the possibility that IBMX, by inhibiting PDE might blunt changes in cAMP in response to
P4, we failed to detect any significant changes in ICFTR, as compared to the ethanol control,
for up to 30min after P4 treatment (Fig. 2G) in accordance with the CNG data. This argues
Development • Advance article
that PKA activity, at least in the sub-plasma membrane compartment sampled by CFTR, does
not change significantly in response to P4. Collectively, membrane bound sensors did not
detect significant changes in cAMP and PKA in response to P4.
Intracellular FRET sensors do not detect changes in cAMP or PKA in response to P4
Since we failed to detect changes cAMP levels or PKA activity using membrane channels,
we needed to rule out the possibility of P4 affecting cAMP/PKA in sub-cellular
compartments that cannot be sampled by membrane channels. To do this we expressed two
widely used intracellular FRET-based sensors, TEPACVV and AKAR2, to monitor cAMP
levels and PKA activity, respectively (Klarenbeek and Jalink, 2014; Zaccolo et al., 2000).
TEPACVV employs the mTurquoise-YFP FRET pair, and has improved dynamic range as a
function of cAMP concentrations (Fig. 3A) (Klarenbeek et al., 2011), whereas the A kinase
activity reporter (AKAR2) FRET sensor uses Clover-mRuby2, and has a phospho-amino acid
binding domain, a defined docking site for recruiting PKA and a substrate domain, resulting
in increased FRET as a function of PKA activity (Fig. 3D) (Lam et al., 2012; Ni et al., 2006).
Confocal imaging using oocytes expressing TEPACVV and AKAR2 showed that both sensors
are diffusely cytoplasmically distributed (Fig. 3A and 3D). We then calibrated both sensors in
oocytes. To that end, TEPACVV-FRET efficiency and AKAR2-mRuby/Clover ratio change
(ΔFRET%) were measured following injection with db-cAMP (Fig. 3B and 3E). Given that
basal cAMP values in Xenopus oocytes averages ~1.25 pmole/oocyte with a range of 0.7-2.7
pmole/oocyte (Maller et al., 1979; O'Connor and Smith, 1976; Schutter et al., 1975), we
injected 0.2-0.8 pmole of db-cAMP, which represent 10-80% increase in cAMP levels.
Though db-cAMP is a cell permeable analog, we wanted to control the exact amounts of db-
cAMP delivered to the oocyte; hence we opted for injection rather than incubation.
The FRET efficiency of TEPACVV decreased significantly by ~40% (p=0.0094) after injection
of 0.8 pmole db-cAMP (Fig. 3B), and AKAR2-ΔFRET% increased significantly in a dose-
dependent fashion after db-cAMP injection (0.2-0.8 pmole) (Fig. 3E).
However, when oocytes expressing either sensor were treated with P4 for up to 30 min, we
were unable to detect any significant changes in cAMP or PKA levels when compared to
control ethanol-treated oocytes (Fig. 3C and F) though the expression of the FRET sensors
did not affect the GVBD response (data not shown). These results support the data from the
Development • Advance article
membrane bound channel sensors, and argue that the release of meiotic arrest induced by P4
occurs independently of changes in cAMP or PKA.
PKA anchoring to AKAP is not essential for releasing meiotic arrest
Although we could not detect a repression of PKA using CFTR or AKAR2, we cannot rule
out the possibility that changes in PKA activity localize to sub-cellular microdomains. To
explore this possibility, we evaluated PKA catalytic subunit (PKAc) subcellular distribution
by immunostaining (Fig. 4A-C). The specificity of the PKAc antibody was confirmed by pre-
incubating the anti-PKA antibody with its peptide antigen (Fig 4A, right panel). PKAc
showed a diffuse cytoplasmic distribution with enrichment below the plasma membrane at
the animal, but not vegetal, pole of the oocyte (Fig. 4A and 4B). Line scan analyses confirm
PKAc enrichment in the sub-plasma membrane domain on the animal but not the vegetal pole
(Fig. 4C). To test whether sub-membrane PKA is specifically inhibited in response to P4, we
prepared membrane and cytosolic fractions by ultracentrifugation before and 30 min after P4
treatment, and tested PKA kinase activity (as described in Figure 5). The efficiency of the
biochemical separation of membrane and cytosolic fractions was confirmed by
immnoblotting for the plasma membrane protein, Na-K-pump (Fig. 4D). We failed to detect
any changes in PKA activity in either the cytosolic or membrane fractions 30 min after P4
treatment (Fig. 4E), or after normalizing for PKAc protein content in the two fractions (Fig.
4F).
To further test whether P4 treatment modulates PKA activity away from the sub-plasma
membrane space, we tested the role of A Kinase-Anchoring proteins (AKAP)s in oocyte
maturation. AKAPs are scaffold proteins that binds PKA regulatory subunit (PKAr), and
anchor it close to its physiological substrates (Colledge and Scott, 1999; Malbon, 2005;
Malbon et al., 2004). PKA interaction with AKAPs was found to be involved in mammalian
oocyte meiotic arrest (Brown et al., 2002; Kovo et al., 2006; Kovo et al., 2002; Newhall et
al., 2006; Nishimura et al., 2013). Injection of an AKAP inhibitory peptide that blocks
AKAP-PKAr interaction in mouse oocytes stimulates oocyte maturation in the presence of
high cAMP (Newhall et al., 2006). We therefore tested the effect blocking AKAP-PKA
interactions on Xenopus oocyte maturation. Oocytes injected with the AKAP inhibitory
peptide matured significantly (p=0.0349) faster as compared to oocytes injected with the
control peptide (Fig. 4G and 4H). However, interfering with AKAP-PKA interaction did not
Development • Advance article
affect maximal GVBD levels achieved in response to P4 (Fig. 4H). These results argue that
PKA anchoring modulates the rate of P4-induced oocyte maturation without affecting its
extent or promoting maturation independently of P4.
PKA activity does not change following the release of meiotic arrest
To further monitor total PKA activity at the single oocyte level we used a quantitative
enzymatic PKA assay (Fig. 5A). We first validated the sensitivity of the assay by injecting
individual oocytes with the PKA inhibitor, PKI, or with PKA catalytic subunit (PKAc) (Fig.
5A). Control oocytes show a basal level of PKA activity at 37±4% phosphorylation of the
substrate peptide (Fig. 5A, lower panel). PKI injection lead to a significant (p=0.0274)
decrease in percent phosphorylated peptide to 16±1% (Fig. 5A). Consistently, PKAc injection
increased the percent phosphorylated peptide to 89±3%, (p0.0001) (Fig. 5A). We then
checked the effect of P4 treatment on PKA activity in single oocytes at 15 sec, 1, 2, 12 and 30
min after P4 and could not detect any change (Fig. 5A).
Concomitant with the PKA activity assay, we tested the effect of PKI and PKAc on oocyte
maturation. This was done at sub-threshold (10-8 M) and saturating (10-5 M) P4 concentrations
(Fig. 5B). Although, PKI injection stimulates oocyte maturation, it was significantly less
effective than 10-5M P4 (Fig. 5B). However, PKI significantly (p=0.0044) enhanced the rate
of P4-induced oocyte maturation (Fig. 5C and 5D), as previously reported (Noh and Han,
1998). The poor activation of oocyte maturation by PKI (Fig. 5B and 5C) is surprising
especially given the robust repression of PKA activity in the same batch of oocytes (Fig. 5A).
These data show a lack of correlation between PKA activity and the ability of PKI and P4 to
release meiotic arrest and induce oocyte maturation, hence bringing into question the
accepted view that the release of meiotic arrest is mediated by a drop in PKA activity. It is
worth noting here that previous studies using PKI to induce maturation in Xenopus oocytes,
including a study from our Lab, reported full maturation when compared to P4 treatment
(Maller and Krebs, 1977; Sun and Machaca, 2004). In contrast, microinjection of PKI did not
cause oocyte maturation, but accelerated GVBD in the presence of P4 in another study (Noh
and Han, 1998). The release of meiotic arrest is an asynchronous process in oocytes from the
same female and show significantly variability among different females based on whether
they are Lab bred, or captured in the wild and whether the females were hormonally
stimulated to test for their ability to produce oocytes. All these factors could affect the
Development • Advance article
effectiveness of various stimuli to induce oocyte maturation. However, notwithstanding this
inherent variability, our results show a lack of correlation between the effectiveness of PKI
and P4 in inducing oocyte maturation and their ability to inhibit PKA activity in the same
batch of oocytes. PKI induces a robust inhibition of PKA activity (by 42.5%) with poor
maximal maturation rates (33+7.3%). In contrast, P4 treatment did not induce any detectable
change in PKA activity, yet fully induced oocyte maturation in the population (75+7.8%).
Note that even P4 in this batch of oocytes did not fully stimulate maturation, consistent with
the poor maturation levels with PKI.
Finally, PKAc injection abolished P4-induced maturation (Fig. 5B) as previously reported
(Daar et al., 1993; Matten et al., 1994), indicating that high PKA activity is sufficient to block
signals emanating from P4 stimulation that lead to oocyte maturation. Collectively our data
argue that the cAMP-PKA axis maintains oocyte meiotic arrest at steady state, but that it is
not modulated downstream of P4 to release meiotic arrest, suggesting a parallel pathway that
actively releases meiotic arrest despite constant high levels of cAMP/PKA.
Increasing cAMP levels does not affect the levels of oocyte maturation
The above data suggest that the cAMP-PKA axis acts as the ‘brake’ that maintains meiotic
arrest, but that the release of this meiotic arrest does not involve changes in cAMP-PKA. To
further assess this model, we tested the effects of the non-hydrolyzable cAMP analog, db-
cAMP, on the extent and kinetics of oocyte maturation. We used CNG-expressing oocytes to
determine the level of endogenous cAMP in the oocyte and the effectiveness of db-cAMP.
Young and Krougliak have previously characterized in Xenopus oocytes the cAMP dose
dependence of the recombinant CNG channel, x-fA4, we used here (Young and Krougliak,
2004). They showed a steep dose dependence of ICNG on cAMP concentrations with a K1/2 of
1 μM, and a saturating current above 200 μM cAMP (see Fig. 2C in Young and Krougliak,
2004). Consistently injecting oocytes with 0.8 or 400 pmole db-cAMP (which corresponds to
400 μM final cAMP concentration in the oocyte) dose-dependently increased ICNG-CaCC (Fig.
6A). Assuming a linear response of the CaCC to Ca2+ flowing through CNG channels, which
is reasonable given the large density of CaCC in the oocyte and their Ca2+ dependence
(Kuruma and Hartzell, 1998), we could translate ICNG-CaCC into cAMP concentration in the
oocyte using the Young and Krougliak dose response curve. We used the current induced by
400 pmole db-cAMP as the maximal current (Imax). With an I/Imax of 0.3 in control
oocytes, this translates to a cAMP concentration of 0.9 μM, within the range of 0.7-2.7
Development • Advance article
μM/oocyte obtained from direct measurements of cAMP by others (Maller et al., 1979;
O'Connor and Smith, 1976; Schutter et al., 1975). This validates the use of I/Imax of ICNG-
CaCC to estimate oocyte cAMP concentration. Furthermore, ICNG-CaCC recordings show that the
injected db-cAMP is stable and results in a sustained increase in cAMP levels for extended
time periods. Injection of 0.8 and 400 pmole db-cAMP dose-dependently and significantly
(p<0.0022) repressed oocyte maturation stimulated with 10-7 M P4 (Fig. 6B). Surprisingly
however, when oocytes were stimulated with the saturating P4 concentration of 10-5 M for
only 1 hour, neither concentration of db-cAMP produced any inhibition of oocyte maturation
(Fig. 6C). Although maximal GVBD levels were not reduced following injection of 400
pmole db-cAMP, the time course of maturation was slowed down (Fig. 6D), requiring 6±0.62
hours to reach 50% GVBD as compared to 3±0.62 hours in control oocytes (p=0.0146) (Fig.
6E). These data indicate that increasing cAMP levels modulate the rate and extent of
maturation. However, increasing the P4-dependent signal overcomes the cAMP-mediated
inhibition and releases meiotic arrest independently from high sustained levels of intracellular
cAMP, of ~400 fold higher levels then resting cAMP concentrations.
Buffering endogenous cAMP doesn’t induce spontaneous maturation
To further test our hypothesis that P4 induces oocyte maturation through an alternative
pathway rather that reducing the levels of cAMP-PKA, we wanted to test whether a reduction
of endogenous cAMP levels is sufficient to induce oocyte maturation independently of P4.
The highest reported P4-dependent reduction in cAMP levels is 60% (Maller et al., 1979).
The FRET sensor TEPACVV binds cAMP and as such, when expressed at high concentrations
in the oocyte (Fig. 7A), would be expected to buffer endogenous cAMP. Expression of
TEPACVV decreased ICNG-CaCC by more than 10-fold (p=0.018), resulting in a 66.7% reduction
of cAMP levels from 0.9 μM to 0.3 μM (Fig. 7B). This reduction in cAMP is in line with the
maximal 60% reduction reported in response to P4 (Maller et al., 1979). Further attesting to
the effectiveness of TEPACVV to buffer cAMP, cells expressing TEPACVV attenuate the
stimulation of ICNG-CaCC following db-cAMP injection by 3-fold (p=0.0413) (Fig. 7B).
Remarkably though, endogenous cAMP buffering by TEPACVV did not release oocyte meiotic
arrest, nor it did enhance maximal GVBD when suboptimal or optimal concentration of P4
were used (Fig. 7C). Additionally, it did not accelerate the rate of oocyte maturation (Fig.
7D). These data support the conclusion that a dip in cAMP is not the trigger to release
Xenopus oocyte meiotic arrest.
Development • Advance article
DISCUSSION
The current accepted model of releasing meiotic arrest in frog oocytes postulates a drop in
cAMP leading to inhibition PKA. Our results challenge this model and argue that P4 releases
meiotic arrest through a cAMP-PKA independent pathway. There is no correlation between
cAMP levels and oocyte maturation (Fig. 1). Furthermore, we could not detect any P4-
mediated changes in cAMP or PKA at the single oocyte level using CNG/CFTR channels
(Fig. 2), FRET sensors (Fig. 3) or a PKA kinase assay (Fig. 4 and 5). Despite these results,
we cannot rule out the possibility that the dip in cAMP and PKA is too small, too localized or
too transient to be detected. However several lines of evidence argue against the need for a
drop in cAMP/PKA to release meiotic arrest: 1) Injection of PKI inhibits PKA activity by
42.5% but results in poor maturation levels (Fig. 5). This argues that inhibition of PKA
activity is not sufficient to release meiotic arrest in the absence of a P4 dependent signal. As
discussed in the introduction whether PKA inhibition is sufficient to release meiotic arrest is
controversial with conflicting reports (Andersen et al., 1998; Daar et al., 1993; Eyers et al.,
2005; Noh and Han, 1998). This could be due to the priming state of the donor female as
discussed above. 2) Maintaining high sustained levels of cAMP inhibited oocyte maturation
but this inhibition was reversed by increasing P4 concentration (Fig. 6). A complete block of
maturation could be achieved only with injection of excess PKA catalytic subunit (Fig. 5). 3)
Buffering cAMP by over 65% was insufficient to induce maturation (Fig. 7).
Collectively, these results argue that meiotic arrest is released through a positive signal
downstream of the membrane P4 receptor (mPRβ), while the cAMP-PKA pathway acts a
“brake” (see model in Fig. 8). It is the balance between the ‘positive’ P4-dependent signal
and the ‘negative’ cAMP-PKA inhibitory signal that defines whether the oocyte commits to
maturation. PKA inhibits the maturation signaling pathway through at least two points:
mRNA translation and Cdc25C activation (Duckworth et al., 2002; Matten et al., 1994),
consistent with the idea that it is acting as a safety mechanism to eliminate spontaneous
maturation in the absence of a positive signal. Although challenging the accepted dogma, our
model is supported by earlier studies (Gelerstein et al., 1988; Nader et al., 2014; Noh and
Han, 1998; Schmitt and Nebreda, 2002), and the yin-yang balance between the mPRβ and
cAMP pathways effectively explains the discrepancies in the literature.
cAMP-PKA levels are maintained at high levels for the duration of the meiotic arrest through
the action of a constitutively active Gs-coupled G protein coupled receptor (GPCR) (Rios-
Cardona et al., 2008). GPR185 is the Xenopus homolog of mammalian GPR3, which has
Development • Advance article
been shown to maintain meiotic arrest in mouse oocytes (Freudzon et al., 2005; Mehlmann et
al., 2002; Mehlmann et al., 2004; Norris et al., 2009). Surprisingly, knockdown of GPR185 is
not sufficient to induce oocyte maturation (Deng et al., 2008; Rios-Cardona et al., 2008),
arguing for other GPCRs involved in maintaining meiotic arrest. We have previously shown
that blocking exocytosis, while maintaining endocytosis, releases meiotic arrest, that P4
results in the internalization of GPR185, and that a GPR185 mutant that does not internalize
is more effective at maintaining meiotic arrest (El Jouni et al., 2007; Nader et al., 2014).
These data argue that P4 induces internalization of GPR185 and potentially other GPCRs that
are involved in maintaining high cAMP-PKA and as such meiotic arrest. Consistent with the
regulation of GPR185 in Xenopus, a drop in cAMP has been shown to correlate with the
release of meiotic arrest in mouse oocytes, through a cGMP signal from the surrounding
somatic cells (Norris et al., 2009). Furthermore, there is evidence in the mouse oocyte for a
positive signal through an EGF-like pathway downstream of LH stimulation (Ashkenazi et
al., 2005; Park et al., 2004), which may represent the counter part for the P4 signal in the
frog.
The mPRβ (also named PAQR8) has seven-transmembrane and belongs to the progestin and
adiponectin receptor family (PAQR) (Tang et al., 2005). It is still unclear whether mPRβ is a
GPCR (Moussatche and Lyons, 2012; Thomas et al., 2007). For example, mPRβ does not
work through Gi to inhibit AC since pertussis toxin, a specific Gi inhibitor, failed to block
P4-induced maturation (Mulner et al., 1985; Olate et al., 1984; Sadler et al., 1984). This is
consistent with the idea that mPR acts through other pathways to release the meiotic arrest.
mPRs and AdipoQ receptors can functionally couple to the same signal transduction pathway
in yeast, suggesting a common mechanism of action (Kupchak et al., 2007). Adiponectin
receptors signal through multiple pathways, including 1) the adaptor protein APPL1, which
links to the trafficking GTPase Rab5 (Buechler et al., 2010; Mao et al., 2006), or p38MAPK
(Heiker et al., 2010); and 2) possess alkaline ceramidase activity to generate sphingosine 1-
phosphate (S1P) (Moussatche and Lyons, 2012). S1P is known to modulate GPR3 activity,
the homolog of GPR185 in mammals (Uhlenbrock et al., 2002; Zhang et al., 2012), and
ceramide is a potential mediator of P4-induced maturation in Xenopus oocytes (Strum et al.,
1995).
In summary, our results challenge the generally accepted initial signaling pathway
downstream of P4, which assumes a drop in cAMP and repression of PKA activity to release
meiotic arrest. Rather, they argue for the existence of a positive signal downstream of mPRβ
Development • Advance article
to overcome the negative inhibitory signal from cAMP-PKA to release meiotic arrest. As
such in the future it would be of great interest to better define the mPRβ signaling pathway.
Development • Advance article
MATERIALS AND METHODS
Molecular Biology. Human CFTR a gift from John Riordan (University of North Carolina)
was amplified using the following primers and subcloned in the Xenopus oocyte expression
vector pSGEM: 5’-CTGCAGGAATTCGATATGCAGAGGTCGCCTCTGGAAAAGGCC3’
(F) and 5’-ATCGATAAGCTTGATCTAAAGCCTTGTATCTTGCACCTCTTCTTC -3’
(R). TEPACVV (mTurq2Del-EPAC(dDEPCD)Q270E-tdcp173Venus(d)EPAC-SH187) a gift
from Kees Jalink (Netherlands Cancer Institute) (Klarenbeek et al., 2011), was subcloned into
the NotI-XbaI sites of pSGEM. The CNG channel chimeric clone X-fA4 was a gift from
Edgar Young (Simon Fraser University, Canada) (Young and Krougliak, 2004). AKAR2
(Lam et al., 2012) was purchased from addgene, and subcloned into the BamHI-EcoRI sites
of pSGEM. All constructs were verified by DNA sequencing and by analytical endonuclease
restriction digestion. mRNAs for all the clones were produced by in vitro transcription after
linearizing the vectors with NheI (CFTR, AKAR2), Not1 (CNG) or Sph1 (TEPACVV) using
the mMessage mMachine T7 kit (Ambion).
Xenopus oocytes. Stage VI Xenopus oocytes were obtained as previously described (Machaca
and Haun, 2002). The donor females were not hormonally stimulated prior to use. The
oocytes were used 24 to 72 hours after harvesting and digestion with collagenase to remove
follicular cells surrounding the oocytes. To study the role of AKAPs, cells were treated with
100 µM of AKAP St-Ht31 inhibitor peptide (Promega, Cat V8211) or its control peptide
(Promega, Cat V8221). Oocytes were injected with RNA and kept at 18°C for 1-2 days after
injection to allow for protein expression.
ELISA cAMP assay. We used the cAMP Complete Enzyme Immunometric Assay kit (Assay
Designs catalog no. 900-163). Ten oocytes were lysed by forcing them through a pipette tip,
in 250 μl of ice-cold 95% ethanol. Extracts were centrifuged at 15,000 g for 15 min at 4°C.
The supernatants were transferred to new tubes and dried under vacuum. The residue was
dissolved, and cAMP was measured according to the kit’s protocol.
CNG and CFTR recording. The ionic currents were recorded using standard two electrode
voltage-clamp recording technique. Recording electrodes were filled with 3 M KCl and
coupled to a Geneclamp 500B controlled with pClamp 10.5 (Axon instruments). The CNG
Development • Advance article
currents were measured indirectly by monitoring the activation of endogenous Ca2+-activated
Chloride Channels (CaCC). Those chloride channels work as biological sensors and give a
very precise and amplified indication of the sub-membrane Ca2+ concentration. The currents
were recorded at a 0.1 Hz frequency using a previously described “triple jump” protocol that
allows the measure of Ca2+ influx (Courjaret and Machaca, 2014; Machaca and Hartzell,
1998). The CNG-CaCC current was measured as the difference in the amplitude of the
currents at +40 mV before and after a voltage pulse to -140 mV that increases the driving
force for Ca2+ entry (Fig. 2B). For CNG recordings using DDA and P4 the extracellular Ca2+
concentration was lowered to 0.9 mM to limit Ca2+ influx. The standard extracellular saline
contained (in mM) 96 NaCl, 2.5 KCl, 1.8 CaCl2, 2 MgCl2, 10 HEPES, pH 7.4. CFTR currents
were recorded at a steady state membrane potential of -80 mV. For CNG currents
normalization was done on the last current trace before treatment, for CFTR experiments
normalization was performed using the average current over a 1 min period before treatment.
TEPACVV and AKAR2 -FRET Imaging. Confocal imaging of live cells was performed using
a LSM710 (Zeiss, Germany) fitted with a Plan Apo 40x/1.3 oil immersion objective. Z-stacks
were taken in 0.45 µm sections using a 1 Airy unit pinhole aperture. When using TEPACVV,
FRET efficiency was calculated using the FRET acceptor bleaching technique. Briefly, this
was done by comparing donor fluorescence intensity in the same sample before and after
photobleaching the acceptor. If FRET was initially present, a resultant increase in donor
fluorescence will occur after photobleaching the acceptor, and the FRET efficiency can be
calculated as follow FRET efficiency = (Dpost-Dpre)/Dpost where Dpos is the fluorescence
intensity of the Donor after acceptor photobleaching, and Dpre the fluorescence intensity of
the Donor before acceptor photobleaching. To that end, TEPACVV expressing-oocytes animal
pole was imaged with the pinhole fully open. mTurquoise was excited at 458 nm and
emission detected at 462-520 nm, and YFP was photobleached at 514 nm and detected at
520-620 nm. YFP photobleaching was done by selecting at least 5 regions and FRET
efficiency calculated using ZEN 2008 (Zeiss) software. When using AKAR2 the excitation
was performed at 488 nm and emission detected at 495-560 nm (Clover) and at 588-702 nm
(mRuby). AKAR2 fluorescence was analyzed using ImageJ software (Schneider et al., 2012)
and the mRuby/Clover fluorescence ratio change (ΔFRET%) were calculated.
Development • Advance article
PepTag assay. PKA kinase activity was measured in single oocyte using the PepTag® Non-
Radioactive Protein Kinase Assays from Promega according to the kit’s provided protocol.
Briefly, The PepTag® assay uses a highly specific fluorescent PKA peptide substrate that
when phosphorylated by changes the peptide’s net charge allowing easy electrophoretic
separation of the phosphorylated and non-phosphorylated peptide. This allows for
quantitative measurement of PKA catalytic activity in oocyte lysates. Phosphorylated and
non-phosphorylated bands were imaged using Geliance 600 Imaging system and the
intensities (I) of the bands were analyzed and corrected on the blank using Image J software
allowing calculation of % phosphorylation with correction on the negative control in the
absence of lysates or PKA.
Oocyte fractionation. Xenopus oocytes (
100) were lysed in Tris-HCL pH 8 (25 mM, EDTA
0.5 mM, EGTA 0.5 mM, protease inhibitor 1/100) using 5 µl per oocyte, followed by a spin
at 1000 g for 10 min. Supernatants were collected and spun for 1 hour at 150,000 g. The
supernatant was saved as the cytosolic fraction and the pellet (membrane fraction) was
dissolved with 50 µl PKA extraction buffer (freshly added to it 10 mM -mercaptoethanol,
protease inhibitor and PMSF). All spinning steps were done at 4C, and the tubes were kept
on ice during the whole procedures. The equivalent of 2 oocytes was used for the PepTag®
assay and also for the western blot to detect PKAc.
Development • Advance article
Western blots. Cells were dounced in MPF lysis buffer (0.08M -glycerophosphate, 20mM
Hepes (pH 7.5), 15mM MgCl2, 20mM EGTA, 1mM Na-Vanadate, 50mM NaF, 1mM DTT,
1mM PMSF and 0.1% protease Inhibitor (Sigma)) and centrifuged twice at 1000 x g for 10
minutes at 4°C to remove yolk granules. The lysates were then incubated with 4% NP40 at
4°C for 2 hours followed by centrifugation at maximum speed for 15 minutes at 4°C and the
supernatant was stored. Supernatants were run on 4-12% SDS-PAGE gels, transferred to
polyvinylidene difluoride) (PVDF) membranes (Millipore), blocked for one hour at room
temperature with 5% Milk in TBS-T buffer (150 mM NaCl and 20 mM Tris; pH 7.6, 0.1%
Tween) and then incubated overnight at 4°C in 3% BSA in TBS-T with one of the following
primary antibodies: anti-PKAc (1/1000, sc-903, Santa Cruz), anti-GFP (1/1000, 2955S, Cell
Signaling), anti-Actin (1/10,000, A1978, Sigma), and anti-Na-K pump (1/1000, 3010S, Cell
Signaling). Blots were washed three times with TBS-T and probed for 1 hour with infrared
fluorescence, IRDye® 800 and 680 secondary antibodies (1/10,000) and the Western Blots
were revealed using the quantitative system LiCor Odyssey Clx Infrared Imaging system.
Oocytes immunostaining. Oocytes were fixed using 4% paraformaldehyde, washed with
phosphate buffered saline containing 30 % of sucrose, and incubated overnight in 50 % OCT
(WVR, clear frozen section compound) with gentle shaking. The oocytes were then
transferred to 100% OCT and frozen into a plastic mold prior to slicing to 7 µm sections on a
cryostat. For immunostaining the oocyte slices were first saturated with 3% BSA and 1 %
goat serum for 1 hour and incubated with anti-PKA antibody (sc-903, Santa Cruz) at a 1/50
dilution for 2 hours. For experiments using the blocking peptide (sc-903p) the primary
antibody was incubated overnight with the blocking peptide at a 5 times concentrations
according to manufacturer’s instructions. The secondary antibody was anti-rabbit IgG
coupled to and Ax488 fluorochrome (A11008, Fisher Scientific) to a 1/500 dilution for 2
hours. Imaging was performed on a Leica SP5 microscope controlled by Leica LAS software
using a 40x/1.3 lens and the pinhole slightly open to an optical slice of 1.3 µm.
Statistics. Values are given as means ± s.e.m. Statistical analysis was performed when
required using student paired and unpaired t-test. P values are indicated as follows: *
(p<0.05), ** (p<0.01), *** (p<0.001), and ns (not significant).
Development • Advance article
Ethical approval. Animals were handled according to Weill Cornell Medicine College
IACUC approved procedures (protocol #2011-0035).
Development • Advance article
ACKNOWLEDGEMENTS
We thank Dr Lu Sun for helping with the TEPACVV FRET confocal imaging and analysis,
and Drs John J. Riordan, Kees Jalink, Edgar C. Young for sharing the CFTR, TEPACVV and
CNG, clones respectively. We thank the Histology Core of Weill Cornell Medicine Qatar for
contributing to these studies. The Histology Core is supported by the BMRP program funded
by Qatar Foundation. This work was funded by NPRP 7-709-3-195 from the Qatar national
research fund (QNRF). The statements made herein are solely the responsibility of the
authors. Additional support for the authors comes from the Biomedical Research Program
(BMRP) at Weill Cornell Medical College in Qatar, a program funded by Qatar Foundation.
Competing Interests
The authors declare no competing interests
Author Contributions
NN designed and performed experiments, analyzed data and wrote the paper. RC designed
and performed experiments. MD and RPK performed experiments. KM developed the
concepts, analyzed data and wrote the paper.
Development • Advance article
REFERENCES
Andersen, C. B., Roth, R. A. and Conti, M. (1998). Protein kinase B/Akt induces
resumption of meiosis in Xenopus oocytes. Journal of Biological Chemistry 273,
18705-18708.
Ashkenazi, H., Cao, X., Motola, S., Popliker, M., Conti, M. and Tsafriri, A. (2005).
Epidermal growth factor family members: endogenous mediators of the ovulatory
response. Endocrinology 146, 77-84.
Bayaa, M., Booth, R. A., Sheng, Y. and Liu, X. J. (2000). The classical progesterone
receptor mediates Xenopus oocyte maturation through a nongenomic mechanism.
Proc.Natl.Acad.Sci.U.S.A 97, 12607-12612.
Bear, C. E., Duguay, F., Naismith, A. L., Kartner, N., Hanrahan, J. W. and Riordan, J.
R. (1991). Cl- channel activity in Xenopus oocytes expressing the cystic fibrosis
gene. The Journal of biological chemistry 266, 19142-19145.
Bement, W. M. and Capco, D. G. (1990). Transformation of the amphibian oocyte into the
egg: structural and biochemical events. Journal of Electron Microscopy Technique 16,
202-234.
Biel, M. and Michalakis, S. (2009). Cyclic nucleotide-gated channels. Handb Exp
Pharmacol, 111-136.
Bravo, R., Otero, C., Allende, C. C. and Allende, J. E. (1978). Amphibian oocyte
maturation and protein synthesis: related inhibition by cyclic AMP, theophylline, and
papaverine. Proc.Natl.Acad.Sci.U.S.A 75, 1242-1246.
Brown, R. L., Ord, T., Moss, S. B. and Williams, C. J. (2002). A-kinase anchor proteins as
potential regulators of protein kinase A function in oocytes. Biol Reprod 67, 981-987.
Buechler, C., Wanninger, J. and Neumeier, M. (2010). Adiponectin receptor binding
proteins--recent advances in elucidating adiponectin signalling pathways. FEBS Lett
584, 4280-4286.
Button, B., Reuss, L. and Altenberg, G. A. (2001). PKC-mediated stimulation of amphibian
CFTR depends on a single phosphorylation consensus site. insertion of this site
confers PKC sensitivity to human CFTR. The Journal of general physiology 117, 457-
468.
Cicirelli, M. F. and Smith, L. D. (1985). Cyclic AMP levels during the maturation of
Xenopus oocytes. Developmental Biology 108, 254-258.
Colledge, M. and Scott, J. D. (1999). AKAPs: from structure to function. Trends Cell Biol
9, 216-221.
Courjaret, R. and Machaca, K. (2014). Mid-range Ca2+ signalling mediated by functional
coupling between store-operated Ca2+ entry and IP3-dependent Ca2+ release. Nature
communications 5, 3916.
Development • Advance article
Daar, I., Yew, N. and Vande Woude, G. F. (1993). Inhibition of mos-induced oocyte
maturation by protein kinase A. Journal of Cell Biology 120, 1197-1202.
Deng, J., Lang, S., Wylie, C. and Hammes, S. R. (2008). The Xenopus laevis isoform of G
protein-coupled receptor 3 (GPR3) is a constitutively active cell surface receptor that
participates in maintaining meiotic arrest in X. laevis oocytes. Mol Endocrinol. 22,
1853-1865.
Duckworth, B. C., Weaver, J. S. and Ruderman, J. V. (2002). G2 arrest in Xenopus
oocytes depends on phosphorylation of cdc25 by protein kinase A.
Proc.Natl.Acad.Sci.U.S.A 99, 16794-16799.
El Jouni, W., Haun, S., Hodeify, R., Hosein, W. A. and Machaca, K. (2007). Vesicular
traffic at the cell membrane regulates oocyte meiotic arrest. Development 134, 3307-
3315.
Eyers, P. A., Liu, j., Hayashi, N. R., Lewellyn, A. L., Gautier, J. and Maller, J. L. (2005).
Regulation of the G(2)/M transition in Xenopus oocytes by the cAMP-dependent
protein kinase. Journal of Biological Chemistry 280, 24339-24346.
Freudzon, L., Norris, R. P., Hand, A. R., Tanaka, S., Saeki, Y., Jones, T. L., Rasenick,
M. M., Berlot, C. H., Mehlmann, L. and Jaffe, L. A. (2005). Regulation of meiotic
prophase arrest in mouse oocytes by GPR3, a constitutive activator of the Gs G
protein. Journal of Cell Biology 171, 255-265.
Gelerstein, S., Shapira, H., Dascal, N., Yekuel, R. and Oron, Y. (1988). Is a decrease in
cyclic AMP a necessary and sufficient signal for maturation of amphibian oocytes?
Developmental Biology 127, 25-32.
Han, S. J., Vaccari, S., Nedachi, T., Andersen, C. B., Kovacina, K. S., Roth, R. A. and
Conti, M. (2006). Protein kinase B/Akt phosphorylation of PDE3A and its role in
mammalian oocyte maturation. EMBO J 25, 5716-5725.
Hanrahan, J. W., Mathews, C. J., Grygorczyk, R., Tabcharani, J. A., Grzelczak, Z.,
Chang, X. B. and Riordan, J. R. (1996). Regulation of the CFTR chloride channel
from humans and sharks. J Exp Zool 275, 283-291.
Heiker, J. T., Kosel, D. and Beck-Sickinger, A. G. (2010). Molecular mechanisms of signal
transduction via adiponectin and adiponectin receptors. Biol Chem 391, 1005-1018.
Huchon, D., Ozon, R., Fischer, E. H. and Demaille, J. G. (1981). [Meiotic maturation of
the ovocyte of Xenopus laevis : a 4-step mechanism]. C R Seances Acad Sci III 292,
65-67.
Josefsberg Ben-Yehoshua, L., Lewellyn, A. L., Thomas, P. and Maller, J. L. (2007). The
role of Xenopus membrane progesterone receptor beta in mediating the effect of
progesterone on oocyte maturation. Mol.Endocrinol. 21, 664-673.
Klarenbeek, J. and Jalink, K. (2014). Detecting cAMP with an EPAC-based FRET sensor
in single living cells. Methods in molecular biology 1071, 49-58.
Development • Advance article
Klarenbeek, J. B., Goedhart, J., Hink, M. A., Gadella, T. W. and Jalink, K. (2011). A
mTurquoise-based cAMP sensor for both FLIM and ratiometric read-out has
improved dynamic range. PLoS One 6, e19170.
Kovo, M., Kandli-Cohen, M., Ben-Haim, M., Galiani, D., Carr, D. W. and Dekel, N.
(2006). An active protein kinase A (PKA) is involved in meiotic arrest of rat growing
oocytes. Reproduction 132, 33-43.
Kovo, M., Schillace, R. V., Galiani, D., Josefsberg, L. B., Carr, D. W. and Dekel, N.
(2002). Expression and modification of PKA and AKAPs during meiosis in rat
oocytes. Mol Cell Endocrinol 192, 105-113.
Kupchak, B. R., Garitaonandia, I., Villa, N. Y., Mullen, M. B., Weaver, M. G., Regalla,
L. M., Kendall, E. A. and Lyons, T. J. (2007). Probing the mechanism of FET3
repression by Izh2p overexpression. Biochim Biophys Acta 1773, 1124-1132.
Kuruma, A. and Hartzell, H. C. (1998). Dynamics of Calcium Regulation of Cl currents in
Xenopus oocytes. American Journal of Physiology 276, C161-C175.
Lam, A. J., St-Pierre, F., Gong, Y., Marshall, J. D., Cranfill, P. J., Baird, M. A.,
McKeown, M. R., Wiedenmann, J., Davidson, M. W., Schnitzer, M. J., et al.
(2012). Improving FRET dynamic range with bright green and red fluorescent
proteins. Nat Methods 9, 1005-1012.
Machaca, K. and Hartzell, H. C. (1998). Asymmetrical distribution of Ca-activated Cl
channels in Xenopus oocytes. Biophys J 74, 1286-1295.
Machaca, K. and Haun, S. (2002). Induction of maturation-promoting factor during
Xenopus oocyte maturation uncouples Ca 2+ store depletion from store-operated Ca
2+ entry Journal of Cell Biology 156, 75-85.
Malbon, C. C. (2005). G proteins in development. Nature reviews. Molecular cell biology 6,
689-701.
Malbon, C. C., Tao, J. and Wang, H. Y. (2004). AKAPs (A-kinase anchoring proteins) and
molecules that compose their G-protein-coupled receptor signalling complexes.
Biochemical Journal 379, 1-9.
Maller, J. L., Butcher, F. R. and Krebs, E. G. (1979). Early effect of progesterone on
levels of cyclic adenosine 3':5'-monophosphate in Xenopus oocytes. Journal of
Biological Chemistry 254, 579-582.
Maller, J. L. and Krebs, E. G. (1977). Progesterone-stimulated meiotic cell division in
Xenopus oocytes. Induction by regulatory subunit and inhibition by catalytic subunit
of adenosine 3':5'-monophosphate-dependent protein kinase. Journal of Biological
Chemistry 252, 1712-1718.
Mao, X., Kikani, C. K., Riojas, R. A., Langlais, P., Wang, L., Ramos, F. J., Fang, Q.,
Christ-Roberts, C. Y., Hong, J. Y., Kim, R. Y., et al. (2006). APPL1 binds to
adiponectin receptors and mediates adiponectin signalling and function. Nat Cell Biol
8, 516-523.
Development • Advance article
Matten, W., Daar, I. and Vande Woude, G. F. (1994). Protein kinase A acts at multiple
points to inhibit Xenopus oocyte maturation. Molecular and Cell Biology 14, 4419-
4426.
Mehlmann, L., Jones, T. L. and Jaffe, L. A. (2002). Meiotic arrest in the mouse follicle
maintained by a Gs protein in the oocyte. Science 297, 1343-1345.
Mehlmann, L., Saeki, Y., Tanaka, S., Brennan, T. J., Evsikov, A. V., Pendola, F. L.,
Knowles, B. B., Eppig, J. J. and Jaffe, L. A. (2004). The Gs-linked receptor GPR3
maintains meiotic arrest in mammalian oocytes. Science 306 1947-1950.
Moussatche, P. and Lyons, T. J. (2012). Non-genomic progesterone signalling and its non-
canonical receptor. Biochemical Society transactions 40, 200-204.
Mulner, O., Huchon, D., Thibier, C. and Ozon, R. (1979). Cyclic AMP synthesis in
Xenopus laevis oocytes: inhibition by progesterone. Biochim Biophys Acta 582, 179-
184.
Mulner, O., Megret, F., Alouf, J. E. and Ozon, R. (1985). Pertussis toxin facilitates the
progesterone-induced maturation of Xenopus oocyte. Possible role of protein
phosphorylation. FEBS Lett 181, 397-402.
Nache, V., Zimmer, T., Wongsamitkul, N., Schmauder, R., Kusch, J., Reinhardt, L.,
Bonigk, W., Seifert, R., Biskup, C., Schwede, F., et al. (2012). Differential
regulation by cyclic nucleotides of the CNGA4 and CNGB1b subunits in olfactory
cyclic nucleotide-gated channels. Sci Signal 5, ra48.
Nader, N., Dib, M., Daalis, A., Kulkarni, R. P. and Machaca, K. (2014). Role for
endocytosis of a constitutively active GPCR (GPR185) in releasing vertebrate oocyte
meiotic arrest. Dev Biol 395, 355-366.
Nader, N., Kulkarni, R. P., Dib, M. and Machaca, K. (2013). How to make a good egg!:
The need for remodeling of oocyte Ca(2+) signaling to mediate the egg-to-embryo
transition. Cell Calcium 53, 41-54.
Newhall, K. J., Criniti, A. R., Cheah, C. S., Smith, K. C., Kafer, K. E., Burkart, A. D.
and McKnight, G. S. (2006). Dynamic anchoring of PKA is essential during oocyte
maturation. Current Biology 16, 321-327.
Ni, Q., Titov, D. V. and Zhang, J. (2006). Analyzing protein kinase dynamics in living cells
with FRET reporters. Methods 40, 279-286.
Nishimura, T., Sugiura, K. and Naito, K. (2013). A-kinase anchor protein 1 (AKAP1)
regulates cAMP-dependent protein kinase (PKA) localization and is involved in
meiotic maturation of porcine oocytes. Biol Reprod 88, 85.
Noh, S. J. and Han, J. K. (1998). Inhibition of the adenylyl cyclase and activation of the
phosphatidylinositol pathway in oocytes through expression of serotonin receptors
does not induce oocyte maturation. J Exp Zool 280, 45-56.
Development • Advance article
Norris, R. P., Ratzan, W. J., Freudzon, M., Mehlmann, L. M., Krall, J., Movsesian, M.
A., Wang, H., Ke, H., Nikolaev, V. O. and Jaffe, L. A. (2009). Cyclic GMP from
the surrounding somatic cells regulates cyclic AMP and meiosis in the mouse oocyte.
Development 136, 1869-1878.
O'Connor, C. M. and Smith, L. D. (1976). Inhibition of oocyte maturation by theophylline:
possible mechanism of action. Dev Biol 52, 318-322.
Olate, J., Allende, C. C., Allende, J. E., Sekura, R. D. and Birnbaumer, L. (1984). Oocyte
adenylyl cyclase contains Ni, yet the guanine nucleotide-dependent inhibition by
progesterone is not sensitive to pertussis toxin. FEBS Lett 175, 25-30.
Park, J. Y., Su, Y. Q., Ariga, M., Law, E., Jin, S. L. and Conti, M. (2004). EGF-like
growth factors as mediators of LH action in the ovulatory follicle. Science 303, 682-
684.
Ramu, Y., Xu, Y. and Lu, Z. (2007). Inhibition of CFTR Cl- channel function caused by
enzymatic hydrolysis of sphingomyelin. Proc Natl Acad Sci U S A 104, 6448-6453.
Rios-Cardona, D., Ricardo-Gonzalez, R. R., Chawla, A. and Ferrell, J. E., Jr. (2008). A
role for GPRx, a novel GPR3/6/12-related G-protein coupled receptor, in the
maintenance of meiotic arrest in Xenopus laevis oocytes. Developmental Biology 317,
380-388.
Sadler, S. E. and Maller, J. L. (1981). Progesterone inhibits adenylate cyclase in Xenopus
oocytes. Action on the guanine nucleotide regulatory protein. Journal of Biological
Chemistry 256, 6368-6373.
---- (1983). Inhibition of Xenopus oocyte adenylate cyclase by progesterone and 2',5'-
dideoxyadenosine is associated with slowing of guanine nucleotide exchange. The
Journal of biological chemistry 258, 7935-7941.
---- (1985). Inhibition of Xenopus oocyte adenylate cyclase by progesterone: a novel
mechanism of action. Adv.Cyclic.Nucleotide.Protein Phosphorylation.Res. 19, 179-
194.
---- (1987). In vivo regulation of cyclic AMP phosphodiesterase in Xenopus oocytes.
Stimulation by insulin and insulin-like growth factor 1. Journal of Biological
Chemistry 262, 10644-10650.
Sadler, S. E., Maller, J. L. and Cooper, D. M. (1984). Progesterone inhibition of Xenopus
oocyte adenylate cyclase is not mediated via the Bordetella pertussis toxin substrate.
Molecular Pharmacology 26, 526-531.
Sadler, S. E., Schechter, A. L., Tabin, C. J. and Maller, J. L. (1986). Antibodies to the ras
gene product inhibit adenylate cyclase and accelerate progesterone-induced cell
division in Xenopus laevis oocytes. Mol Cell Biol 6, 719-722.
Schmitt, A. and Nebreda, A. R. (2002). Inhibition of Xenopus oocyte meiotic maturation by
catalytically inactive protein kinase A. Proc.Natl.Acad.Sci.U.S.A 99, 4361-4366.
Development • Advance article
Schneider, C. A., Rasband, W. S. and Eliceiri, K. W. (2012). NIH Image to ImageJ: 25
years of image analysis. Nature Methods 9, 671-675.
Schorderet-Slatkine, S. and Baulieu, E. E. (1982). Forskolin increases cAMP and inhibits
progesterone induced meiosis reinitiation in Xenopus laevis oocytes. Endocrinology
111, 1385-1387.
Schorderet-Slatkine, S., Schorderet, M. and Baulieu, E. E. (1982). Cyclic AMP-mediated
control of meiosis: effects of progesterone, cholera toxin, and membrane-active drugs
in Xenopus laevis oocytes. Proc.Natl.Acad.Sci.U.S.A 79, 850-854.
Schorderet-Slatkine, S., Schorderet, M., Boquet, P., Godeau, F. and Baulieu, E. E.
(1978). Progesterone-induced meiosis in Xenopus laevis oocytes: a role for cAMP at
the "maturation-promoting factor" level. Cell 15, 1269-1275.
Schutter, A. P., Kram, R., Hubert, E. and Brachet, J. (1975). Cyclic nucleotides and
amphibian development. Exp Cell Res 96, 7-14.
Smith, L. D. (1989). The induction of oocyte maturation: transmembrane signaling events
and regulation of the cell cycle. Development 107, 685-699.
Strum, J. C., Swenson, K. I., Turner, J. E. and Bell, R. M. (1995). Ceramide triggers
meiotic cell cycle progression in Xenopus oocytes. A potential mediator of
progesterone-induced maturation. The Journal of biological chemistry 270, 13541-
13547.
Sun, L. and Machaca, K. (2004). Ca 2+ cyt negatively regulates the initiation of oocyte
maturation Journal of Cell Biology 165 63
-75.
Tang, Y. T., Hu, T., Arterburn, M., Boyle, B., Bright, J. M., Emtage, P. C. and Funk, W.
D. (2005). PAQR proteins: a novel membrane receptor family defined by an ancient
7-transmembrane pass motif. J Mol Evol 61, 372-380.
Thibier, C., Mulner, O. and Ozon, R. (1982). In vitro effects of progesterone and estradiol-
17 beta on choleragen activated Xenopus oocyte adenylate cyclase. J Steroid Biochem
17, 191-196.
Thomas, P., Pang, Y., Dong, J., Groenen, P., Kelder, J., de Vlieg, J., Zhu, Y. and Tubbs,
C. (2007). Steroid and G protein binding characteristics of the seatrout and human
progestin membrane receptor alpha subtypes and their evolutionary origins.
Endocrinology 148, 705-718.
Tian, J., Kim, S., Heilig, E. and Ruderman, J. V. (2000). Identification of XPR-1, a
progesterone receptor required for Xenopus oocyte activation.
Proc.Natl.Acad.Sci.U.S.A 97, 14358-14363.
Uhlenbrock, K., Gassenhuber, H. and Kostenis, E. (2002). Sphingosine 1-phosphate is a
ligand of the human gpr3, gpr6 and gpr12 family of constitutively active G protein-
coupled receptors. Cellular signalling 14, 941-953.
Development • Advance article
Voronina, E. and Wessel, G. M. (2003). The regulation of oocyte maturation.
Curr.Top.Dev.Biol 58, 53-110.
Webe, W. M., Segal, A., Vankeerberghen, A., Cassiman, J. J. and Van Driessche, W.
(2001). Different activation mechanisms of cystic fibrosis transmembrane
conductance regulator expressed in Xenopus laevis oocytes. Comp Biochem Physiol A
Mol Integr Physiol 130, 521-531.
Young, E. C. and Krougliak, N. (2004). Distinct structural determinants of efficacy and
sensitivity in the ligand-binding domain of cyclic nucleotide-gated channels. The
Journal of biological chemistry 279, 3553-3562.
Zaccolo, M., De Giorgi, F., Cho, C. Y., Feng, L., Knapp, T., Negulescu, P. A., Taylor, S.
S., Tsien, R. Y. and Pozzan, T. (2000). A genetically encoded, fluorescent indicator
for cyclic AMP in living cells. Nat Cell Biol 2, 25-29.
Zhang, B. L., Li, Y., Ding, J. H., Dong, F. L., Hou, Y. J., Jiang, B. C., Shi, F. X. and Xu,
Y. X. (2012). Sphingosine 1-phosphate acts as an activator for the porcine Gpr3 of
constitutively active G protein-coupled receptors. Journal of Zhejiang University.
Science. B 13, 555-566.
Development • Advance article
Figures
Figure 1. Lack of correlation between cAMP levels and oocyte maturation. A) Current
model by which P4 releases meiotic arrest. B) cAMP levels in oocytes before and seconds
(left panel) or minutes (right panel) after treatment with P4 (n=7) or ethanol (n=4). Data are
normalized to untreated oocytes. C) cAMP levels at 15 sec, 1 and 30 min after P4 treatment,
normalized to untreated oocytes. D) Percentage of oocytes reaching the GVBD stage
following P4 treatment at different time points (n=6). Oocytes were treated transiently with
10-5 M P4 for 15 sec, 1 min, 30 min and overnight (O/N) or with its carrier ethanol O/N.
After the exposure to P4 for 15 sec, 1 min and 30 min the oocytes were immediately washed
twice with the culture media L15. Oocytes were then left overnight and the following day, the
GVBD count was done. P4 (P4), PKA (Protein Kinase A), AC (Adenylate Cyclase), MPF
(Maturation Promoting Factor).
Development • Advance article
Development • Advance article
Figure 2. CNG and CFTR do not detect changes in cAMP or PKA activity after P4. A)
Schematic representation of the regulation of CFTR, CNG and CaCC. B) Oocytes were
injected with CNG-RNA (10 ng/oocyte), and the CNG-CaCC current (ICNG-CaCC) (6 oocytes)
recorded at least 24 hours later as compared to Naïve oocytes (3 oocytes). C) CNG
expressing oocytes were treated with 0.64 mM DDA, and ICNG-CaCC recorded before and after
treatment (4 oocytes). Data are normalized to before treatment. D) CNG expressing oocytes
were treated with P4 (P4) (19 oocytes) or ethanol (EtOH) (18 oocytes) for up to 30 min and
ICNG-CaCC recorded continuously. Data are normalized to untreated oocytes. E) oocytes were
injected with CFTR-RNA (10 ng/oocyte) or water and the CFTR current (ICFTR) recorded at
least 48 hours later (9 oocytes). F) CFTR-expressing oocytes were pretreated with IBMX (1
mM) for 15 minutes and ICFTR recorded in IBMX for 25 min recording (no perfusion). Time
course of current run down (top panel) and current amplitude after 25 min (bottom panel) in
IBMX treated (IBMX, 16 oocytes) or untreated (Control, 7 oocytes) CFTR expressing
oocytes. G) Time course of ICFTR after P4 (6 oocytes) or EtOH (5 oocytes) addition in CFTR
expressing oocytes treated with IBMX.
Development • Advance article
Figure 3. cAMP and PKA FRET sensors do not detect changes in response to P4. A and
D) Upper panels: Cartoon of the two FRET sensors TEPACVV (A) and AKAR2 (D) used for
cAMP and PKA respectively. Lower panels: Representative confocal images 48 hrs after
injecting TEPACVV (50ng/oocyte) or AKAR2 (10ng/oocyte) RNAs. B) FRET efficiency from
Development • Advance article
TEPACVV expressing oocytes 10 min after db-cAMP (0.8 pmole) injection (3 oocytes). Data
are normalized to the FRET efficiency before db-cAMP injection. C) Time course of
TEPACVV FRET efficiency after P4 or ethanol treatment. E) ΔFRET changes in AKAR2 10
min after injection of increasing amounts of db-cAMP as indicated (3 oocytes). Data are
normalized to ΔFRET before treatment. F) Time course of AKAR2 ΔFRET changes in
response to P4 or ethanol. Data are normalized to before treatment.
Development • Advance article
Figure 4. Subcellular PKA distribution and activity. A) Representative confocal images
from oocyte sections immunostained using an anti-PKAαc antibody in the absence (left
panel) or presence of blocking peptide (right panel). B) Bright field (BF) (left panel) and
fluorescence PKAαc immunostaining (right panel) of an oocyte at the equator showing both
the vegetal pole (VP) and the animal pole (AP). C) Zoomed focal planes of an oocyte at the
animal and vegetal poles showing PKA enrichment below the plasma membrane at the
animal pole (left panel). Average intensity of the immunostaining signal along linescans
across the plasma membrane as illustrated on the images in the left panels (right panel). D)
Development • Advance article
Na-K-pump Western Blot (WB) after separation of the membrane and cytoplasmic fractions,
before and after P4 treatment. E) PKA kinase activity measured using the PepTag® assay in
whole lysates, membrane and cytoplasmic fractions before and after P4 treatment (n=3). F)
Cytosolic and membrane fractions were examined by western blotting using the PKAαc
antibody (top panel). Relative PKA activity corrected on the amount of PKA protein
quantified from WB (bottom panel). G) Representative GVBD-time course after P4 treatment
in the presence of AKAP inhibitor or control peptide. H) Oocytes were treated with P4 in the
presence of AKAP inhibitor or its control peptide. The time required for 50% of the oocytes
in the population to reach the GVBD (left y axis) and the maximum GVBD (right y axis) are
shown (n=6).
Development • Advance article
Figure 5. P4 does not affect PKA kinase activity. A) Oocytes were injected with PKI (60
ng), PKAc (50 ng) or treated with P4 for different time points and PKA activity was
measured using the PepTag® assay. Representative agarose gels, showing the phosphorylated
and the non-phosphorylate peptide are presented on the top panel, and a quantification of
PKA as % phosphorylated peptide in naïve, PKI injected (n=3), PKAc injected (n=3), and
treated with P4 for different time points (n=4) are shown on the lower panel. B) Oocyte
maturation levels following: PKI injection, P4 treatments at the indicated concentrations in
the presence or absence of PKI (60 ng) or PKAc (50 ng) (minimum of n=3). The
Development • Advance article
injections/treatments were done early in the morning and GVBD followed throughout the
day. C) Representative GVBD-time course after PKI injection alone and P4 treatment in the
presence or absence of PKI. D) The time required for 50% of the oocytes in the population to
reach the GVBD in oocytes treated with P4 alone or in the presence of PKI (n=3).
Development • Advance article
Figure 6. High sustained cAMP slows down but does not block maturation. A) CNG
expressing oocytes were injected with 0.8 (5 oocytes) or 400 pmole (4 oocytes) db-cAMP
and ICNG-CaCC measured. B-C) Maturation levels following overnight P4 treatment at 10-7M
(n=7) (B), or for one hour at 10-5M (n=3) (C) in the presence or absence of 0.8 or 400 pmole
db-cAMP. D) Representative GVBD-time course after P4 (10-5M) in the presence or absence
of 400 pmole db-cAMP. E) Oocyte maturation rates in oocytes treated with P4 alone or in the
presence of 400 pmole db-cAMP (n=4).
Development • Advance article
Figure 7. Buffering endogenous cAMP doesn’t induce spontaneous maturation. A)
Oocytes were injected with TEPACVV-RNA (10 ng/oocytes), and 48 hours later, Naïve and
TEPACVV–expressing oocytes were processed and TEPACVV expression was assessed by
western blot using GFP antibody. B) ICNG-CaCC measured in CNG-expressing oocytes in the
presence (4 oocytes) or absence of TEPACVV (5 oocytes), before or after injection of 0.8
pmole db-cAMP. C) Levels of oocyte maturation in the presence or absence of TEPACVV
and/or P4 (10-8 and 10-7M) (minimum of n=5) after overnight incubation. D) Representative
GVBD-time course after P4 treatment (10-7M ) in the presence or absence of TEPACVV. P4
was added in the morning and the time-course GVBD was followed throughout the day.
Development • Advance article
Figure 8. Proposed model for the signal transduction cascade leading to the release of
Xenopus oocyte meiotic arrest. See text for details.
Development • Advance article
... Despite this general consensus, it has been proposed that progesterone triggers meiotic maturation by bypassing the negative inhibitory signal imposed by the cAMP-PKA via an independent positive signal. 18,19 This hypothesis suggests that multiple progesterone receptors are involved in the control of meiosis, one for maintaining the prophase arrest [20][21][22] and the other(s) for triggering meiotic resumption, 23,24 possibly acting through both cAMP-dependent and -independent pathways. 5,11,[25][26][27][28] In Xenopus, PKA inhibition occurs within 1 h in response to progesterone. ...
... Progesterone treatment induces a drop of cAMP levels that is followed by PKA inactivation. [4][5][6][7][8][9] Since this model has been recently challenged, 18,19,25,27 we confirmed that, under our experimental conditions, PKA inactivation is induced by progesterone and is sufficient to induce the ''upstream'' events and meiosis resumption. For this purpose, we measured in ovo the activity of PKA by injecting in the oocyte at different times during meiosis resumption, its substrate, recombinant Arpp19, and by monitoring its phosphorylation at S109 (Figure 5A and 5B). ...
... Additionally, it has been proposed that progesterone induces events that are independent of PKA downregulation. 19,25,27 Our Cip1-based approach can be applied to understand whether these events control the activation of the starting activity of Cdk1 or regulate the autoamplification loop. ...
... Dans l'ovocyte de xénope, l'un des premiers événements intracellulaires bien décrit et déclenché par la progestérone est une chute de 20% de l'AMPc suite à l'inhibition de l'AC, sans intervention d'une régulation de l'activité des PDEs (Mulner et al., 1980;Mulner et al., 1979;Sadler and Maller, 1981;Sadler and Maller, 1983). La chute de l'AMPc intervient dans la demi-heure qui suit la stimulation hormonale (Cicirelli and Smith, 1985;Mulner et al., 1979;Thibier et al., 1982) (Nader et al., 2016). Elle est immédiatement suivie par l'inhibition totale de PKA qui demeure ensuite inactive tout au long de la maturation méiotique ) ( Figure 11). ...
... L'activité catalytique de PKA est donc bien au centre de son action inhibitrice sur la division méiotique. Plus récemment, il a été proposé que la progestérone à de fortes concentrations soit capable d'induire l'activation de Cdk1 indépendamment de la chute des niveaux intracellulaires en AMPc et de l'inhibition de PKA (Nader et al., 2016). Cette conclusion est principalement basée sur le fait que l'injection de PKI n'est pas capable d'induire la reprise de la méiose en absence de stimulation hormonale alors que, dans ce même article, 25% des ovocytes atteignent néanmoins GVBD (Nader et al., 2016). ...
... Plus récemment, il a été proposé que la progestérone à de fortes concentrations soit capable d'induire l'activation de Cdk1 indépendamment de la chute des niveaux intracellulaires en AMPc et de l'inhibition de PKA (Nader et al., 2016). Cette conclusion est principalement basée sur le fait que l'injection de PKI n'est pas capable d'induire la reprise de la méiose en absence de stimulation hormonale alors que, dans ce même article, 25% des ovocytes atteignent néanmoins GVBD (Nader et al., 2016). Il est donc possible que la progestérone à des doses sub-luminaires (10 µM) puisse déclencher un mécanisme alternatif à la chute des niveaux en AMPc et à l'inhibition de PKA. ...
Thesis
Ma thèse a visé à comprendre comment la cellule germinale femelle, ou ovocyte, reprend la division méiotique. Dans tout le règne animal, l’ovocyte est bloqué en prophase de première division méiotique et reprend la division au moment de l’ovulation, sous l’effet d’un signal externe, la progestérone chez le xénope. Cette hormone déclenche une voie de signalisation qui aboutit après 3 à 5 heures à l’activation de la kinase Cdk1. Les premières molécules activées de Cdk1 déclenchent une boucle d’auto-amplification permettant son activation totale et l’entrée en division. Cette boucle repose sur des phosphorylations catalysées par Cdk1 et ne fonctionne que si la phosphatase qui contrecarre son activité, PP2A-B55δ, est inhibée. Cette inhibition dépend de la protéine Arpp19, un inhibiteur spécifique de PP2A-B55δ quand elle est phosphorylée sur la S67 par la kinase Greatwall. Arpp19 a un second rôle lors de la reprise de la méiose. Chez tous les vertébrés, l’arrêt en prophase est maintenu par une forte activité de PKA. Chez le xénope, Arpp19 est phosphorylée par PKA sur un résidu distinct de celui ciblé par Greatwall, la S109. En réponse à la progestérone, l’inhibition de PKA provoque la déphosphorylation d’Arpp19, ce qui permet d'enclencher la voie de signalisation conduisant à l'activation de Cdk1. Une phosphatase est donc requise pour déphosphoryler Arpp19 sur S109 et lever le verrou exercé sur l'activation de Cdk1 en prophase. Mon travail a visé à élucider l’identité moléculaire de cette phosphatase inconnue. Par des approches de biochimie et protéomique, j’ai identifié cette phosphatase comme étant PP2A-B55δ. Par des approches fonctionnelles dans des extraits acellulaires et des ovocytes, j’ai établi que PP2A-B55δ est active en prophase. La phosphorylation d’Arpp19 sur la S109 dépend d'une balance entre les activités de PKA et PP2A-B55δ, en faveur de la kinase. En réponse à la progestérone, l’activité de PP2A-B55δ n’est pas affectée et l’inhibition de PKA suffit à la déphosphorylation d’Arpp19. PP2A-B55δ orchestre donc la levée de l’arrêt en prophase et l'entrée en phase M en agissant à deux périodes et sur deux sites distincts d’Arpp19, respectivement ciblés par PKA et Greatwall.
... In fish and amphibians, follicle cells secrete steroid hormones that act on one or more types of oocyte membrane receptor. The extent and importance of the cAMP drop observed at maturation initiation vary in different systems studied, and other signaling pathways are likely to be important, at least in some species [39,40,43,44,45,46]. Upstream of the vertebrate follicle, LH production from the pituitary is under control of the master reproductive regulator GnRH receptor acting via Gα q and cytoplasmic Ca ++ release. ...
... In mouse oocytes, a cAMP decrease upon hormone stimulation triggers maturation; however, in fish and frog oocytes the degree and role of this decrease is debated. Several types of oocytes receptor (orange) may respond to steroid hormones (Pg) in different species of amphibians and fish, but the relative importance of multiple downstream signalling pathways remains to be clarified [1,43,44,45,46]. See text for discussion. ...
Article
Full-text available
The reproductive hormones that trigger oocyte meiotic maturation and release from the ovary vary greatly between animal species. Identification of receptors for these maturation-inducing hormones (MIHs) and understanding how they initiate the largely conserved maturation process remain important challenges. In hydrozoan cnidarians including the jellyfish Clytia hemisphaerica, MIH comprises neuropeptides released from somatic cells of the gonad. We identified the receptor (MIHR) for these MIH neuropeptides in Clytia using cell culture–based “deorphanization” of candidate oocyte-expressed G protein–coupled receptors (GPCRs). MIHR mutant jellyfish generated using CRISPR-Cas9 editing had severe defects in gamete development or in spawning both in males and females. Female gonads, or oocytes isolated from MIHR mutants, failed to respond to synthetic MIH. Treatment with the cAMP analogue Br-cAMP to mimic cAMP rise at maturation onset rescued meiotic maturation and spawning. Injection of inhibitory antibodies to the alpha subunit of the Gs heterodimeric protein (GαS) into wild-type oocytes phenocopied the MIHR mutants. These results provide the molecular links between MIH stimulation and meiotic maturation initiation in hydrozoan oocytes. Molecular phylogeny grouped Clytia MIHR with a subset of bilaterian neuropeptide receptors, including neuropeptide Y, gonadotropin inhibitory hormone (GnIH), pyroglutamylated RFamide, and luqin, all upstream regulators of sexual reproduction. This identification and functional characterization of a cnidarian peptide GPCR advances our understanding of oocyte maturation initiation and sheds light on the evolution of neuropeptide-hormone systems.
... Interestingly, GPR185 is internalized in response to P4, which inhibits its ability to activate adenylate cyclase and thus renders the oocyte more permissive for maturation (22). However, a global decrease in cAMP levels is not detectable in response to P4 and P4-dependent maturation proceeds even when cAMP is high (25). This argues that P4 signals through parallel pathways can overwhelm the cAMP-PKA-dependent inhibition. ...
Preprint
Full-text available
The steroid hormone progesterone (P4) regulates multiple aspects of reproductive and metabolic physiology. Classical P4 signaling operates through nuclear receptors that regulate transcription of target genes. In addition, P4 signals through membrane P4 receptors (mPRs) in a rapid nongenomic modality. Despite the established physiological importance of P4 nongenomic signaling, its detailed signal transduction remains elusive. Here, using Xenopus oocyte maturation as a well-established physiological readout of nongenomic P4 signaling, we identify the lipid hydrolase ABHD2 (α/β hydrolase domain-containing protein 2) as an essential mPRβ co-receptor to trigger meiosis. We show using functional assays coupled to unbiased and targeted cell-based lipidomics that ABHD2 possesses a phospholipase A2 (PLA2) activity that requires both P4 and mPRβ. This PLA2 activity bifurcates P4 signaling by inducing mPRβ clathrin-dependent endocytosis and producing lipid messengers that are G-protein coupled receptors agonists. Therefore, P4 drives meiosis by inducing the ABHD2 PLA2 activity that requires both mPRβ and ABHD2 as obligate co-receptors. Significance Statement Nongenomic progesterone signaling is important for many physiological functions yet its detailed early signaling steps remain elusive. Here we show that progesterone induces rapid signaling to drive Xenopus oocyte meiosis. Progesterone requires two cell membrane receptor to work in unison to signal. The co-receptor complex possesses lipase activity that produces lipid messenger and changes cell membrane structure. This leads to receptor endocytosis and activation of additional pathway to trigger meiosis. Nongenomic progesterone signaling operates in every tissue and regulates reproductive and metabolic physiology. Therefore, our findings have broad implications on our understanding of nongenomic progesterone signaling.
... Oocytes arrest in prophase of meiosis I across species. [1][2][3][4][5] In humans, this arrest often lasts for decades. 6 During the arrest, oocytes grow and acquire competence (an ability to complete meiosis and undergo fertilization to produce healthy progeny). ...
Article
Full-text available
Maternal RNAs are stored from minutes to decades in oocytes throughout meiosis I arrest in a transcriptionally quiescent state. Recent reports, however, propose a role for nascent transcription in arrested oocytes. Whether arrested oocytes launch nascent transcription in response to environmental or hormonal signals while maintaining the meiosis I arrest remains undetermined. We test this by integrating single-cell RNA sequencing, RNA velocity, and RNA fluorescence in situ hybridization on C. elegans meiosis I arrested oocytes. We identify transcripts that increase as the arrested meiosis I oocyte ages, but rule out extracellular signaling through ERK MAPK and nascent transcription as a mechanism for this increase. We report transcript acquisition from neighboring somatic cells as a mechanism of transcript increase during meiosis I arrest. These analyses provide a deeper view at single-cell resolution of the RNA landscape of a meiosis I arrested oocyte and as it prepares for oocyte maturation and fertilization.
... Mos, on the other hand, triggers ERK activation which in turn inactivates Myt1, the kinase that blocks Cdk1 activity (Palmer et al., 1998;Peter et al., 2002). Additionally, cAMP/PKA-independent progesterone-induced non-genomic signaling is also documented (Nader et al., 2016), where a signaling endosome of mPRβ interacts with adaptor proteins and Akt2 to promote maturation (Nader et al., 2020). Thus, activation of MPF possibly proceeds through a parallel cAMP/PKA-dependent and independent pathway and may converge to promote Xenopus oocyte maturation (Figure 3). ...
Article
During oogenesis, oocytes arrest at meiotic prophase I to acquire competencies for resuming meiosis, fertilization, and early embryonic development. Following this arrested period, oocytes resume meiosis in response to species‐specific hormones, a process known as oocyte maturation, that precedes ovulation and fertilization. Involvement of endocrine and autocrine/paracrine factors and signaling events during maintenance of prophase I arrest, and resumption of meiosis is an area of active research. Studies in vertebrate and invertebrate model organisms have delineated the molecular determinants and signaling pathways that regulate oocyte maturation. Cell cycle regulators, such as cyclin‐dependent kinase (CDK1), polo‐like kinase (PLK1), Wee1/Myt1 kinase, and the phosphatase CDC25 play conserved roles during meiotic resumption. Extracellular signal‐regulated kinase (ERK), on the other hand, while activated during oocyte maturation in all species, regulates both species‐specific, as well as conserved events among different organisms. In this review, we synthesize the general signaling mechanisms and focus on conserved and distinct functions of ERK signaling pathway during oocyte maturation in mammals, non‐mammalian vertebrates, and invertebrates such as Drosophila and Caenorhabditis elegans.
... Furthermore, mPR in the frog oocyte does not signal through Gα i because pertussis toxin does not block maturation [17,18]. Recent data also argue against mPR mediating its signaling by acting on the adenylate cyclase-cAMP pathway [19]. Upstream of the MAPK cascade, P4 triggers polyadenylation and translation of Mos, an oocyte-specific MAPK kinase activator [15,16]. ...
Article
Full-text available
The steroid hormone progesterone (P4) mediates many physiological processes through either nuclear receptors that modulate gene expression or membrane P4 receptors (mPRs) that mediate nongenomic signaling. mPR signaling remains poorly understood. Here we show that the topology of mPRβ is similar to adiponectin receptors and opposite to that of G-protein-coupled receptors (GPCRs). Using Xenopus oocyte meiosis as a well-established physiological readout of nongenomic P4 signaling, we demonstrate that mPRβ signaling requires the adaptor protein APPL1 and the kinase Akt2. We further show that P4 induces clathrin-dependent endocytosis of mPRβ into signaling endosome, where mPR interacts transiently with APPL1 and Akt2 to induce meiosis. Our findings outline the early steps involved in mPR signaling and expand the spectrum of mPR signaling through the multitude of pathways involving APPL1.
... Conversely, the inhibition of PKA activity by microinjection of either the regulatory R subunit of PKA or the specific inhibitor of PKA, PKI, leads to the re-initiation of oocyte maturation in the absence of progesterone [103,104]. Despite this general consensus that a decrease in cAMP and PKA activity is necessary and sufficient for meiosis reinitiation, one research group has proposed that progesterone triggers meiotic maturation independently of GαS-adenylate cyclase-cAMP-PKA, bypassing the negative inhibitory signal imposed by the cAMP-PKA via an independent positive signal [171]. This hypothesis is difficult to reconcile with the evidence presented above, accrued from more than 40 years of research in many laboratories. ...
Article
Full-text available
During oocyte development, meiosis arrests in prophase of the first division for a remarkably prolonged period firstly during oocyte growth, and then when awaiting the appropriate hormonal signals for egg release. This prophase arrest is finally unlocked when locally produced maturation initiation hormones (MIHs) trigger entry into M-phase. Here, we assess the current knowledge of the successive cellular and molecular mechanisms responsible for keeping meiotic progression on hold. We focus on two model organisms, the amphibian Xenopus laevis, and the hydrozoan jellyfish Clytia hemisphaerica. Conserved mechanisms govern the initial meiotic programme of the oocyte prior to oocyte growth and also, much later, the onset of mitotic divisions, via activation of two key kinase systems: Cdk1-Cyclin B/Gwl (MPF) for M-phase activation and Mos-MAPkinase to orchestrate polar body formation and cytostatic (CSF) arrest. In contrast, maintenance of the prophase state of the fully-grown oocyte is assured by highly specific mechanisms, reflecting enormous variation between species in MIHs, MIH receptors and their immediate downstream signalling response. Convergence of multiple signalling pathway components to promote MPF activation in some oocytes, including Xenopus, is likely a heritage of the complex evolutionary history of spawning regulation, but also helps ensure a robust and reliable mechanism for gamete production.
... Our results shown that 50 µM of AA stimulated the proliferation of granulosa cells by activating both ERK1/2 and Akt signaling pathways that regulate target genes expression. In the mammalian oocyte, AA has the potential to stimulate cAMP/PKA by increased cPLA2 activity [44,45]. Maitra and colleagues concluded that PKA inhibition may have a potent role in MAPK (also known as extracellular signal-regulated kinases, ERK1/2) activation for initiation of oocyte maturation [46]. ...
Article
Full-text available
In the present study, AA was used to challenge bovine ovarian granulosa cells in vitro and the related parameters of cellular and molecular biology were measured. The results indicated that lower doses of AA increased survival of bovine granulosa cells whereas higher doses of AA suppressed survival. While lower doses of AA induced accumulation of lipid droplet in granulosa cells, the higher dose of AA inhibited lipid accumulation, and AA increased abundance of FABP3, CD36 and SLC27A1 mRNA. Higher doses of AA decreased the secretion of E2 and increased the secretion of P4 accompanied by down-regulation of the mRNA abundance of CYP19A1, FSHR, HSD3B1 and STAR in granulosa cells. The signaling pathways employed by AA in the stimulation of genes expression included both ERK1/2 and Akt. Together, AA specifically affects physiological features, gene expression levels and steroid hormone secretion, and thus altering the functionality of granulosa cells of cattle.
Article
Full-text available
Meiosis is a specialized cell cycle that requires sequential changes to the cell division machinery to facilitate changing functions. To define the mechanisms that enable the oocyte-to-embryo transition, we performed time-course proteomics in synchronized sea star oocytes from prophase I through the first embryonic cleavage. Although we find that protein levels are broadly stable, our analysis reveals that dynamic waves of phosphorylation underlie each meiotic stage. We find that the phosphatase PP2A-B55 is reactivated at the meiosis I/II transition resulting in the preferential dephosphorylation of threonine residues. Selective dephosphorylation is critical for directing the MI / MII transition as altering PP2A-B55 substrate preferences disrupts key cell cycle events after meiosis I. In addition, threonine to serine substitution of a conserved phosphorylation site in the substrate INCENP prevents its relocalization at anaphase I. Thus, through its inherent phospho-threonine preference, PP2A-B55 imposes specific phosphoregulated behaviors that distinguish the two meiotic divisions.
Article
Full-text available
The relationship between the mos protoonco-gene protein and cAMP-dependent protein kinase (PKA) during the maturation of Xenopus oocytes was investigated. Microinjection of the PKA catalytic sub-unit (PKA0 into Xenopus oocytes inhibited oocyte mat-uration induced by the mos product but did not markedly affect the autophosphorylation activity of injected mos protein. By contrast, PKA~ did not inhibit matu-ration promoting factor (MPF) activation or germinal vesicle breakdown (GVBD) that was initiated by injecting crude MPF preparations. In addition, inhibiting endogenous PKA activity by microinjecting the PKA regulatory subunit (PKAr) induced oocyte maturation that was dependent upon the presence of the endog-enous mos product. Moreover, PKA~ potentiated mos protein-induced MPF activation in the absence of pro-gesterone and protein synthesis. These data are consistent with the hypothesis that progesterone-induced release from G2/M is regulated via PKA~ and that PKA~ negatively regulates a downstream target that is positively regulated by mos.
Article
Full-text available
Vertebrate oocytes are naturally arrested at prophase of meiosis I for sustained periods of time before resuming meiosis in a process called oocyte maturation that prepares the egg for fertilization. Members of the constitutively active GPR3/6/12 family of G-protein coupled receptors represent important mediators of meiotic arrest. In the frog oocyte the GPR3/12 homolog GPRx (renamed GPR185), has been shown to sustain meiotic arrest by increasing intracellular cAMP levels through GαSβγ. Here we show that GPRx is enriched at the cell membrane (∼80%), recycles through an endosomal compartment at steady state, and loses its ability to signal once trapped intracellularly. Progesterone-mediated oocyte maturation is associated with significant internalization of both endogenous and overexpressed GPRx. Furthermore, a GPRx mutant that does not internalize in response to progesterone is significantly more efficient than wild-type GPRx at blocking oocyte maturation. Collectively our results argue that internalization of the constitutively active GPRx is critical to release oocyte meiotic arrest.
Article
Full-text available
The versatility and universality of Ca(2+) signals stem from the breadth of their spatial and temporal dynamics. Spatially, Ca(2+) signalling is well studied in the microdomain scale, close to a Ca(2+) channel, and at the whole-cell level. However, little is known about how local Ca(2+) signals are regulated to specifically activate spatially distant effectors without a global Ca(2+) rise. Here we show that an intricate coupling between the inositol 1,4,5 trisphosphate (IP3) receptor, SERCA pump and store-operated Ca(2+) entry (SOCE) allows for efficient mid-range Ca(2+) signalling. Ca(2+) flowing through SOCE is taken up into the ER lumen by the SERCA pump, only to be re-released by IP3Rs to activate distal Ca(2+)-activated Cl(-) channels (CaCCs). This CaCC regulation contributes to setting the membrane potential of the cell. Hence functional coupling between SOCE, SERCA and IP3R limits local Ca(2+) diffusion and funnels Ca(2+) through the ER lumen to activate a spatially separate Ca(2+) effector.
Article
Full-text available
A variety of genetically encoded reporters use changes in fluorescence (or Förster) resonance energy transfer (FRET) to report on biochemical processes in living cells. The standard genetically encoded FRET pair consists of CFPs and YFPs, but many CFP-YFP reporters suffer from low FRET dynamic range, phototoxicity from the CFP excitation light and complex photokinetic events such as reversible photobleaching and photoconversion. We engineered two fluorescent proteins, Clover and mRuby2, which are the brightest green and red fluorescent proteins to date and have the highest Förster radius of any ratiometric FRET pair yet described. Replacement of CFP and YFP with these two proteins in reporters of kinase activity, small GTPase activity and transmembrane voltage significantly improves photostability, FRET dynamic range and emission ratio changes. These improvements enhance detection of transient biochemical events such as neuronal action-potential firing and RhoA activation in growth cones.
Article
Full-text available
Olfactory cyclic nucleotide-gated (CNG) ion channels are essential contributors to signal transduction of olfactory sensory neurons. The activity of the channels is controlled by the cyclic nucleotides guanosine 3',5'-monophosphate (cGMP) and adenosine 3',5'-monophosphate (cAMP). The olfactory CNG channels are composed of two CNGA2 subunits, one CNGA4 and one CNGB1b subunit, each containing a cyclic nucleotide-binding domain. Using patch-clamp fluorometry, we measured ligand binding and channel activation simultaneously and showed that cGMP activated olfactory CNG channels not only by binding to the two CNGA2 subunits but also by binding to the CNGA4 subunit. In a channel in which the CNGA2 subunits were compromised for ligand binding, cGMP binding to CNGA4 was sufficient to partly activate the channel. In contrast, in heterotetrameric channels, the CNGB1b subunit did not bind cGMP, but channels with this subunit showed activation by cAMP. Thus, the modulatory subunits participate actively in translating ligand binding to activation of heterotetrameric olfactory CNG channels and enable the channels to differentiate between cyclic nucleotides.
Article
Full-text available
We cloned the complete coding sequences of porcine Gpr3, Gpr6, and Gpr12 genes. Further, on the basis of their high levels of sequence similarity, these genes are identified as a subfamily of G protein-coupled receptors. These putative protein sequences also showed high sequence identity with other mammalian orthologs, including several highly conserved motifs. A wide expression of the Gpr3 gene in pigs was observed through tissue distribution analysis by reverse transcriptase-polymerase chain reaction (RT-PCR) and real-time PCR, specially in the brain, pituitary, fat, liver and oocyte, where its strong expression was observed. The Gpr3 gene was found to be located on chromosome 6 and a single exon coded for the entire open-reading frame. Expression of porcine Gpr3 in HEK293 cells resulted in constitutive activation of adenylate cyclase (AC) similar in amplitude to that produced by fully stimulated G(s)-coupled receptors. Moreover, sphingosine 1-phosphate (S1P) could increase AC activation via the constitutively active Gpr3 receptor. When a Gpr3-green fluorescent protein (GFP) construct was expressed in HEK293 cells, GFP-labeled Gpr3 protein was shown to be localized in the plasmalemma and subcellular membranes. After S1P treatment, agonist-mediated internalization could be visualized by confocal microscopy. In short, our findings suggest the porcine Gpr3, Gpr6, and Gpr12 genes as a subfamily of G protein-coupled receptors, and porcine Gpr3 was a constitutively active G protein-coupled receptor. Constitutive activation of AC and agonist-mediated internalization of Gpr3 receptor could be modulated by the S1P, suggesting that S1P might act as an activator for porcine Gpr3 receptor.
Article
In the mammalian oocyte, the cAMP-dependent protein kinase (PKA) has critical functions in the maintenance of meiotic arrest and oocyte maturation. Because PKA is spatially regulated, its localization was examined in developing oocytes. Both regulatory subunits (RI and RII) and the catalytic subunit (C) of PKA were found in oocytes and metaphase II-arrested eggs. In the oocyte, RI and C were predominantly localized in the cortical region, while RII showed a punctate distribution within the cytoplasm. After maturation to metaphase II, RI remained in the cortex and was also localized to the meiotic spindle, while RII was found adjacent to the spindle. C was diffuse within the cytoplasm of the egg but was enriched in the cytoplasm surrounding the metaphase spindle, much like RII. The polarized localization and redistribution of RI, RII, and C suggested that PKA might be tethered by A-kinase anchor proteins (AKAPs), proteins that tether PKA close to its physiological substrates. An AKAP, AKAP140, was identified that was developmentally regulated and phosphorylated in oocytes and eggs. AKAP140 was shown to be a dual-specific AKAP, having the ability to bind both RI and RII. By compartmentalizing PKA, AKAP140 and/or other AKAPs could spatially regulate PKA activity during oocyte development.
Article
Cyclic nucleotides such as cGMP and cAMP play pivotal roles as second messengers in many biological processes. Upon stimulation of appropriate signal transduction pathways, the levels of these messengers change rapidly. Such variations in second messenger level may also be spatially restricted within the cell. To detect dynamic and local changes in second messengers, we need to study them in living cells with high spatial and temporal resolution. Focusing on cAMP, here we describe how imaging of an EPAC-based FRET sensor in single cells provides that spatiotemporal resolution.
Article
In mammalian oocytes, cAMP-dependent protein kinase (PKA) has critical functions in meiotic arrest and meiotic maturation. Although it is known that subcellular localization of PKA is regulated by A-kinase anchor proteins (AKAPs) and that PKA compartmentalization is essential for PKA functions, the role of AKAPs in meiotic regulation has not been fully elucidated. In the present study, we performed far-Western blotting using porcine PRKAR2A for detection of AKAPs and found several signals in porcine oocytes for the first time. Among these signals, a 150 kDa-AKAP showed the major expression; we found this AKAP to be the product of porcine AKAP1. Overexpression of AKAP1 changed the PKA localization and promoted meiotic resumption of porcine oocytes even in the presence of a high concentration of cAMP, which inhibits meiotic resumption by inducing high PKA activity. On the contrary, knockdown of AKAP1 showed inhibitory effects on meiotic resumption and oocyte maturation. We also revealed that the expression level of AKAP1 in porcine growing oocytes, which show meiotic incompetence and PKA mislocalization, was significantly lower than that of fully grown oocytes. However AKAP1 insufficiency was not the primary cause of the meiotic incompetence of the growing oocytes. These results suggest the possibility that the regulation of PKA localization by AKAP1 is involved in the meiotic resumption and oocyte maturation but not in the meiotic incompetence of porcine growing oocytes.
Article
The egg-to-embryo transition marks the initiation of multicellular organismal development and is mediated by a specialized Ca(2+) transient at fertilization. This explosive Ca(2+) signal has captured the interest and imagination of scientists for many decades, given its cataclysmic nature and necessity for the egg-to-embryo transition. Learning how the egg acquires the competency to generate this Ca(2+) transient at fertilization is essential to our understanding of the mechanisms controlling egg and the transition to embryogenesis. In this review we discuss our current knowledge of how Ca(2+) signaling pathways remodel during oocyte maturation in preparation for fertilization with a special emphasis on the frog oocyte as additional reviews in this issue will touch on this in other species.