Mathematical model of FSH-induced cAMP production in ovarian
follicles
F.
Clément1,
D.
Monniaux2,
J.
Stark4,
K.
Hardy5,
J. C.
Thalabard3,
S.
Franks6, and
D.
Claude1
1 Institut National de Recherche en Informatique et
Automatique, Unité de Recherche de Rocquencourt, Domaine de
Voluceau, Rocquencourt, 78153 Le Chesnay Cedex; 2 Unité de
Physiologie de la Reproduction et des Comportements, UMR 6073 Institut
National de la Recherche Agronomique-Centre National de la Recherche
Scientifique-Université F. Rabelais de Tours, 37380 Nouzilly;
3 Unité de Formation et de Recherche Necker,
Biostatistiques-Informatique Médicale,
Endocrinologie-Médecine de la Reproduction, Université
Paris V, Groupe Hospitalier Necker-Enfants Malades, 75743 Paris,
France; 4 Centre for Nonlinear Dynamics and its Applications,
University College London, London WC1E 6BT; 5 Division of
Pediatrics, Obstetrics and Gynecology, Department of Reproductive
Science and Medicine, Imperial College of Science, Technology and
Medicine, Hammersmith Hospital, London W12 0NN; and 6 Division
of Pediatrics, Obstetrics and Gynecology, Department of Reproductive
Science and Medicine, Imperial College of Science, Technology and
Medicine, St. Mary's Hospital Medical School, London W2 1PG, United
Kingdom
 |
ABSTRACT |
During the terminal part of their development,
ovarian follicles become totally dependent on gonadotropin supply to
pursue their growth and maturation. Both gonadotropins,
follicle-stimulating hormone (FSH) and luteining hormone (LH), operate
mainly through stimulatory G protein-coupled receptors, their signal
being transduced by the activation of the enzyme adenylyl cyclase and
the production of second-messenger cAMP. In this paper, we develop a
mathematical model of the dynamics of the coupling between FSH receptor
stimulation and cAMP synthesis. This model takes the form of a set of
nonlinear, ordinary differential equations that describe the changes in
the different states of FSH receptors (free, bound, phosphorylated, and
internalized), coupling efficiency (activated adenylyl cyclase), and
cAMP response. Classical analysis shows that, in the case of constant
FSH signal input, the system converges to a unique, stable equilibrium
state, whose properties are here investigated. The system also appears
to be robust to nonconstant input. Particular attention is given to the
influence of biologically relevant parameters on cAMP dynamics.
signal transduction; granulosa cells; follicle-stimulating hormone; cyclic adenosine monophosphate
 |
INTRODUCTION |
FOLLICULOGENESIS IS THE
PROCESS of growth and functional maturation undergone by ovarian
follicles, from the time they leave the pool of primordial (quiescent)
follicles until ovulation, at which point they release a fertilizable
oocyte. Most of the developing follicles never reach the ovulatory
stage but degenerate by a process known as atresia (12).
The gonadotropic hormones follicle-stimulating hormone (FSH) and
luteinizing hormone (LH) play a major role in the regulation of
terminal follicular development through the control of proliferation
and differentiation of the granulosa cells surrounding the oocyte
(29). Gonadotropin secretion is, in turn, modulated by
granulosa cell products such as estradiol and inhibin. During the
follicular phase of the ovarian cycle, negative feedback is responsible
for reducing FSH secretion, leading to the degeneration of all but
those follicles selected for ovulation. Finally, positive feedback is
responsible for triggering the LH ovulatory surge leading to ovulation
(10). Both FSH and LH operate mainly through G
protein-coupled transmembrane receptors, transducing their signal by
activation of the enzyme adenylyl cyclase and production of
second-messenger cyclic adenosine monophosphate (cAMP)
(30).
In previous studies (6, 7), we investigated the divergent
commitment of granulosa cells toward proliferation, differentiation, or
apoptosis in response to their hormonal environment. Under cumulative exposure to gonadotropins, granulosa cells progressively lose their ability to proliferate and acquire a fully differentiated state. The accumulation of intracellular cAMP beyond a threshold seems
to be a key point in cell cycle arrest (26), because it is
believed to lead to the activation of cyclin kinase inhibitors (11). We thus believe that a better understanding of
gonadotropin-induced cAMP production will help gain insight into
changes in the rate of differentiation among granulosa cells during
terminal development (8).
In CONSTRUCTION OF A SIGNAL TRANSDUCTION MODEL, we describe
the mathematical model after stating the biological assumptions on
which it is based; in STABILITY ANALYSIS FOR CONSTANT FSH
INPUT, we focus on the analysis of the model in the case of a
constant FSH level; CONTROL OF FSH-INDUCED CAMP LEVELS is
devoted to the numerical application of the model; the physiological
implications of these results are discussed in DISCUSSION;
and mathematical details are given in APPENDIX.
 |
CONSTRUCTION OF A SIGNAL TRANSDUCTION MODEL |
Physiological Background: Transduction of the Gonadotropic Signal
in Granulosa Cells
Terminal follicular development is strictly dependent on FSH
supply. Before the selection of the follicle for ovulation, granulosa cells are responsive only to FSH. As follicular maturation progresses, the coupling between FSH receptor stimulation and the activation of
adenylyl cyclase becomes more and more efficient, leading to a steady
increase in cAMP production (13). The accumulation of
FSH-induced cAMP coincides with the appearance of and subsequent dramatic increase in LH receptors, allowing LH to act as a surrogate for FSH in granulosa cells (36). Conversely, when
gonadotropin, and especially FSH, plasma levels are too low to meet the
follicle's trophic requirements, uncoupling of receptor stimulation
with cAMP production is one of the earliest events occurring during granulosa cell death and follicular atresia (15).
The binding of FSH to its transmembrane receptors triggers an
intracellular signal via the heterotrimeric G proteins. The FSH-bound
receptor activates the G
-stimulatory (G
s)
subunit, which interacts with adenylyl cyclase to generate
an increase in cyclic AMP. Once cAMP is synthesized, it either binds
and activates specific protein kinases such as protein kinase A or is
degraded by cyclic nucleotide phosphodiesterase (PDE)
(30).
The control of cAMP levels in granulosa cells involves both fast
biochemical processes, such as binding and desensitization, occurring
on a time scale of a few minutes, and slower physiological processes
lasting hours or even a few days, which result mainly in changing the
efficiency of the enhancement of cAMP synthesis by stimulated FSH
receptors via adenylyl cyclase activation. The increase in this
coupling efficiency is a progressive, hormonally regulated process
(29), so that the degree of maturation of a follicle can
be characterized by the average cAMP level in its granulosa cells. The
design of our model follows from the interactions between these
contrasting biochemical and physiological dynamics. From here onward,
we will focus on the dynamics of intracellular cAMP in an average
granulosa cell from the time the follicle becomes able to respond to
FSH in term of cAMP production.
Biological Assumptions
The model is based on the following assumptions, which are
supported by the available biological knowledge on FSH signal
transduction in granulosa cells during the first part of terminal
follicular development.
Binding of FSH to its receptor (RFSH) results in
the formation of an active complex (XFSH)
Bound receptors activate adenylyl cyclase (E) through a
conformational change in the associated G protein
Activated adenylyl cyclase (EFSH)
synthesizes cAMP from the substrate Mg2+ ATP
cAMP is hydrolyzed into AMP by PDE
Bound receptors are subjected to a desensitization process through
cAMP-mediated phosphorylation
Phosphorylated inactive complexes (XpFSH)
undergo internalization into the cell, where receptors are dissociated
from FSH
Internalized receptors (Ri) are recycled
back to the cell membrane, whereas FSH is hydrolyzed
Consideration of only those reactions relevant to follicular
development allows some simplifications to be made. Reactions generating short-lived intermediary species are neglected. In particular, the cycle of G protein activation/deactivation is not
modeled explicitly. The process of receptor synthesis is assumed to
compensate both for intracellular receptor degradation and for the
depletion of the receptor pool during cell division, so that the total
number of FSH receptors in different states (free, active,
phosphorylated, and internalized) remains constant (4), leading to the following cellular cycle for FSH receptors under different states
Finally, cAMP-independent desensitization is not taken into
consideration, because its behavior during the maturation of granulosa
cells is not yet known. In addition, the amount of FSH is assumed to be
sufficiently large that its concentration is unaffected by binding to receptors.
Model Equations
Let RFSH, XFSH,
XpFSH, and Ri be,
respectively, the concentrations of free, bound active, bound inactive
(phosphorylated), and internalized FSH receptors (italics indicate
concentrations). Let EFSH be the
concentration of activated adenylyl cyclase, and let cAMP be
the concentration of intracellular cAMP. Let k+, k
, ki, and
kr be the rate constants for FSH binding, FSH unbinding, bound complex internalization, and receptor recycling to the
cell membrane, respectively. The function
describes the (cAMP-dependent) rate of receptor desensitization. The rates of change
of the concentrations are given by the following ordinary differential
equations
|
(1)
|
|
(2)
|
|
(3)
|
|
(4)
|
|
(5)
|
|
(6)
|
Equations 1, 4, and 6 result from applying
the principle of mass action. In Eq. 4, ATP is treated as a
nonlimiting substrate at a constant concentration, and its effect is
included in the kinetic constant
. In the same way, the
concentration of the enzyme PDE is included in
kPDE.
The desensitization rate
in Eqs. 2 and 5 is a
Hill function of intracellular cAMP
with saturation value (
), half-saturating cAMP concentration
(
), and slope of the increase in
(
) as real parameters.
This sigmoidal dependence accounts in a compact way for the
phosphorylation cascade occurring downstream of cAMP, including transmembrane receptors as phosphorylation targets. Thus
phosphorylation in the model is assumed to be cAMP mediated in a
dose-dependent, increasing, and saturated manner. On qualitative
grounds, this choice was substantiated by the critical importance of
cAMP-dependent postreceptor events for desensitization
(21). On quantitative grounds, the Hill function allows
either for a progressive effect of cAMP level or for a rather
all-or-nothing effect, depending on the value of the slope parameter
. Besides, the phosphorylation rate is assumed to be bounded by the
saturation value
, which reflects the limits in the phosphorylation
capacity resulting from the balance between phosphorylation through
kinases and dephosphorylation through phosphatases.
Equation 3 governs the change in the coupling variable
EFSH and is designed to be understood from a
physiological rather than a biochemical viewpoint.
acts as a time
scale parameter. Whenever it takes a low value (
1), the changes
in the coupling variable EFSH are slower than
those of the other variables of the model. The amplification parameter
measures the degree of signal amplification and represents the
average number of adenylyl cyclase molecules activated by one bound
receptor at steady state.
The choice for the right-hand term of Eq. 3 is subject to
the following physiological constraints, which make it specific to
granulosa cells: 1) there is a basal concentration of
activated adenylyl cyclase (E0 > 0), due
to minor constitutive activity of G proteins; 2) under
cumulative exposure to FSH, the capacity for cAMP production in
response to FSH stimulation increases during terminal follicular
development (EFSH is an increasing function as
long as
XFSH > EFSH); 3) the increase in the
efficiency of coupling is correlated with an increase in the
follicle's vulnerability toward FSH supply (as soon as
XFSH < EFSH,
EFSH starts decreasing); 4) coupling
and uncoupling are autoamplified processes, due to paracrine and
autocrine mechanisms enhancing the follicular sensitivity to FSH (right
EFSH term of amplification).
Model Reduction
The total number of receptors remains constant; hence, Eqs.
1, 2, 5, and 6 are subject to the conservation law
|
(7)
|
where RT is the constant size of the global receptor
pool. We can thus replace Ri in Eq. 1
by RT
(RFSH + XFSH + XpFSH), reducing the system to five equations
|
(8)
|
|
(9)
|
|
(10)
|
|
(11)
|
|
(12)
|
Initial values of the variables will be denoted, respectively, as
R0, X0,
E0, cAMP0, and
Xp0.
Boundedness
A basic requirement for a physiological model to be plausible is
that solutions should remain bounded for all time and that concentrations should remain nonnegative. It is easy to verify (for
details see APPENDIX, Upper Bounds of Solutions)
that, as long as kPDE > 0, this is the
case in the above model for constant FSH input, so that, for any
> 0, there exists a T
0, such that for all
t
T
If kPDE = 0, there is no mechanism for
removing cAMP from the system; hence, cAMP concentrations can grow
without bound. This is obviously not a physiologically realistic case.
 |
STABILITY ANALYSIS FOR CONSTANT FSH INPUT |
Quasi-Steady-State Model and Steady States
When
1, the changes in RFSH,
XFSH, cAMP and
XpFSH can be considered fast compared with the
change in EFSH. Applying a quasi-steady-state
approximation to Eqs. 9, 11, and 12 in case of constant FSH input leads respectively to the following relations
|
(13)
|
|
(14)
|
|
(15)
|
Substituting this into Eq. 8, we obtain
|
(16)
|
where
Hence, the quasi-steady-state assumption defines a
quasi-steady-state model reducing system 8-12 to a
one-variable, nonlinear differential equation
|
(17)
|
Figure 1 illustrates the
degree of discrepancy, as far as the changes in
EFSH are concerned, between the complete and the reduced models.

View larger version (17K):
[in this window]
[in a new window]
|
Fig. 1.
Comparison between approximate and exact solution of activated
adenylyl cyclase concentration (EFSH). Solid
lines correspond to the solutions obtained from Eq. 10 in
the complete model; dotted lines represent the solutions obtained from
Eq. 17, assuming quasi-steady state on the other variables.
From left to right, the 3 pairs of curves are
respectively associated with time scale parameter ( ) values of 0.1, 0.01, and 0.005. Time unit is 102 s, and
E0 = 0.1 × 104
molecules/cell. Other parameter values are displayed in Table 2. The
amplitude of the discrepancy between the reduced and complete models
increases as value increases, whereas the length of the transient
period increases with decreasing value.
|
|
Although the transient behavior of EFSH is
under the control of
, its steady-state
E
is not. The steady state corresponding
to E0 > 0 is characterized by
|
(18)
|
with X
a solution of
|
(19)
|
By use of simple geometric reasoning, it is possible to prove that
Eq. 19 always admits a unique, positive real root (see APPENDIX, Existence and Uniqueness of Strictly
Positive Roots of Eq. 19). This root defines the unique
equilibrium state of the system. The level of intracellular cAMP at
this steady state is an increasing function of the following
parameters: FSH input, size of receptor pool
RT and constants k+,
ki kr, and
. Conversely it is a decreasing function of k
,
kPDE, and
(see proof in
APPENDIX, Control of the Steady-State Level
cAMP*). The influence of the slope parameter
is not univocal:
cAMP steady-state level is either an increasing function of
if
or a decreasing one in the opposite case.
Stability of the Steady State
Linear stability of systems of ordinary differential equations
such as those arising in this paper is determined by the roots of a
polynomial. The stability analysis involves the linearization of
system 8-12 in the form
where q is the vector of the time-dependent
concentrations (RFSH,
XFSH, XpFSH,
EFSH, cAMP), and Mj is
the matrix of the linearized nonlinear terms evaluated at the steady
state, which is defined as the Jacobian matrix and is given by
using for simplicity the notation
Solutions are obtained by setting
where q0 is the constant vector of initial
values, and the eigenvalues
are the roots of a characteristic
polynomial | Mj
I | = 0, with I the identity matrix.
The steady state is stable if all roots
have a negative real part.
Because formal calculation did not allow us to carry through the study
of the real part signs, we made use of the Hurwitz criterion
(5), which derives necessary and sufficient conditions for
negativity. In the case where steady-state
,
*, is saturated and
can be approximated by the constant value
, the Hurwitz criterion shows that the eigenvalues of Mj have strictly negative
real parts, so that the steady state is asymptotically stable (see
details in APPENDIX, Hurwitz Criterion for Linear
Stability Analysis). Application of the criterion is not so
straightforward when the dependence of
* on cAMP* is
taken into account, so that linear stability analysis in the general
case remains to be studied. However, note that, because the equilibrium
in the case of a constant
is hyperbolic, it is locally structurally
stable; hence, it will also be asymptotically stable whenever the
dependence of
on cAMP is weak (16).
Controllability Analysis
Roughly speaking, a dynamic system is said to be controllable if,
starting from given initial conditions, one can find an admissible
control variable (here FSH) such that there exists a time for which the
state variables will be steered to prescribed values. Controllability
is an important feature of the model, because if the system were not
controllable, the equilibrium values would be reached independently of
FSH, which would be unsatisfactory for a model that we want to use for
control purposes.
The study of the controllability matrix associated with the linearized
system at steady state is more easily tractable (24) than
that of the Jacobian matrix and allows us to conclude that the
nonlinear system 8-12 is locally strongly accessible
everywhere except if E0 = 0 (details in
APPENDIX, Local Control of the System). The
controllability analysis does not require the assumption of a constant
FSH input, so it leads to quite general results regarding FSH input shape.
 |
CONTROL OF FSH-INDUCED cAMP LEVELS |
Dimension of Model Variables and Parameters
To handle the model equations from a numerical viewpoint, we need
to know the dimensions and ranges of both variables and parameters so
as to confine cAMP output values within physiological limits. As far as
variables are concerned (Table 1),
granulosa-specific information is available. During terminal follicular
development, there are ~103-104 FSH binding
sites per cell (17, 27). Experimental measurements of
cAMP concentration in granulosa cells (1, 13, 19, 20) under different conditions lie in a range from 0.1 to 10 pmol/106 cells, roughly corresponding to 0.2 to 20 × 104 molecules per cell (molecular mass of cAMP is 327 Da).
As far as parameters are concerned (Table
2), FSH binds its receptors with high
affinity; the equilibrium dissociation constant
Kd = k
/k+ is on the order of
10
10 M (23). The other kinetic constants are
assigned ranges of values consistent with published biochemical models
in other cell types (14, 32). We used physiological FSH
plasma concentrations as inputs, lying in the range from 1 to 10 ng/ml,
with 3 ng/ml corresponding to 10
10 M on the basis of an
average FSH molecular mass of 3 × 104 Da
(35). The lowest FSH values correspond to tonic secretion, whereas the highest rather correspond to the surge secretion or the
level used in stimulation protocols or in vitro experiments.
Physiological Meaning of Variations in Parameter Values
Variations in the model parameter values correspond to
physiological or pathological alterations in the different steps of FSH
signal processing by granulosa cells. Binding equilibrium parameters
(k+, k
) might vary
among species in relation to species-specific genetic differences or
even intraspecies as a result of functional mutations or receptor
polymorphism affecting the extracellular domain of the FSH receptor
that contains the binding site. The number of FSH receptors available
for binding can be experimentally altered. For instance, the treatment
of granulosa cells with neuramidase, which catalyzes the removal of
cell surface sialic acid, increases specific FSH binding
(25). In the model, such a treatment would result in an
increase in the size of the global receptor pool
(RT). The level of adenylyl cyclase activation
might differ according to genetic differences affecting the FSH
receptor domain(s) responsible for G protein activation, the degree of
G
s-intrinsic GTPase activity, or the use of adenylyl
cyclase activators such as forskolin. Variations in the amplification
parameter (
) may partly account for such processes, as this
parameter is related to the average number of adenylyl cyclase
molecules activated by one FSH-bound receptor during its lifetime as an
active form. Different values of the cAMP synthesis parameter (
)
could correspond to different types of adenylyl cyclase, as
several of them have been identified (33). Signal
extinction in the model is ensured by the hydrolysis of cAMP and the phosphorylation-induced desensitization of bound receptors. Variations in the hydrolysis rate
(kPDE) can be experimentally achieved through
chronic infusion with a PDE inhibitor such as isobutylmethylxanthine
(IBMX) or, conversely, through constrained overexpression of PDE in
cultured cells. Similarly, infusion of kinase inhibitors such as
staurosporine alters the balance between kinase and phosphatase
activities and can be related to variations in the parameters of the
phosphorylation rate, especially its saturation value (
). The rate
of renewal of free FSH receptors results from a dynamic equilibrium
between the processes of internalization, degradation, recycling, and
synthesis. In the model, renewal is dependent on both the
internalization (ki,) and the recycling (kr) rates. Finally, the time scale parameter
(
) measures the speed of amplification of FSH signal in granulosa
cells. It integrates the role of cross talks with different signaling
pathways, notably paracrine and autocrine signaling through growth
factors and steroids.
For a given combination of the model parameters, variations in FSH
input help to determine the range of values where the model is the most
sensitive to changes in FSH levels. Besides, increasing the level of
the constant FSH input illustrates how the cell is protected against an
overflow in intracellular cAMP.
Control of cAMP Steady-State and Transient Levels
Study of cAMP steady-state levels.
Given a fixed value of FSH input, every parameter except the time
scale parameter (
) affects the value of the cAMP steady-state level
cAMP*. This value is an increasing function of FSH input, the size of receptor pool RT, the rate constants
k+, ki, and
kr, and the half-saturating cAMP concentration
. Conversely it is a decreasing function of the unbinding rate
k
, the hydrolysis rate
kPDE, and the saturation value
. The way
cAMP* is influenced by a parameter is analyzed formally in
APPENDIX (Control of the Steady-State Level
cAMP*). Interestingly, the
-parameter, which rules the rate of
increase in the phosphorylation rate (the slope of the Hill function),
has a nonunivocal influence on cAMP*, depending on the value
of RT compared with a threshold value given by
For values of RT lower than
RThresh, cAMP* is an increasing function of
,
whereas it is a decreasing function for values >RThresh.
When RT = RThresh, altering the
value of
simply has no effect. This means that, if the receptor
pool is small, the cAMP steady-state level rather benefits from an
almost all-or-nothing effect of cAMP level on the phosphorylation
process than from a progressive, smoother effect.
Beyond this qualitative study, quantitative dose-effect-like curves can
be constructed from Eq. 19, which amounts, in term of
cAMP*, to
These curves are displayed in Figs.
2 and 3.
The range of variations for some parameters has been deliberately
exaggerated beyond physiological values so as to examine a large range
of cAMP steady-state-reachable levels for a given parameter.

View larger version (18K):
[in this window]
[in a new window]
|
Fig. 2.
Influence of follicle-stimulating hormone (FSH), size of
receptor pool (RT), FSH binding rate
(k+), FSH unbinding rate
(k ), amplification parameter ( ), cAMP
synthesis rate ( ), cAMP hydrolysis rate by phosphodiesterase
(kPDE), phosphorylated receptor internalization
rate (ki), and internalized receptor recycling
rate (kr) on cAMP steady-state
(cAMP*) level. Panels illustrate the influence of the FSH
input and other parameter values from left to
right and top to bottom on
cAMP*.
|
|

View larger version (16K):
[in this window]
[in a new window]
|
Fig. 3.
Influence of saturation value of phosphorylation rate ( ), slope parameter ( ), and half-saturating cAMP concentration
( ) (depending on RT value) on
cAMP*. Panels illustrate the influence of the parameters of
the phosphorylation rate on cAMP*. Top:
influence of ( ) (left) and (right);
bottom: influence of when either
|
(left)
|
or
|
(right)
|
|
|
Study of cAMP transient levels.
To retrace the history of FSH-induced cAMP production, starting from a
quiescent initial state, we also need to understand the transient
behavior of the system before its reaching steady state. To do so, we
performed a series of numerical simulations of the model using a
computer program written in C language. The differential equations were
integrated by means of a Runge-Kutta method of order 4 (28), with a step of 0.01 s. cAMP transient levels
depend in a complicated manner on the values of the model parameters
and FSH input. The same steady state can in particular be achieved in
different ways depending on the value of the time scale parameter
.
This parameter is the only one in the model that does not affect the
steady state but instead exerts a substantial influence on the
transient behavior, especially of the coupling variable
EFSH.
Initial values.
The start of the simulation was assumed to be the point in time when
FSH receptors become efficiently coupled to G proteins and start
influencing intracellular cAMP production, which corresponds to the
follicle's entering the FSH-responsive stage. Before this point, it is
assumed that there is a minor basal level of activated adenylyl
cyclase, resulting in a low basal level of cAMP production uncoupled
with FSH input. Initial conditions for the differential equations
Eqs. 8-12 are set to
These correspond to the binding equilibrium between
RFSH and XFSH and steady
state for Eq. 11 with activated cyclase level E0 decoupled from receptor stimulation.
Influence of the receptor pool size.
Figure 4 (top) illustrates the
effects of varying the size RT within the
physiological range of 0.5-2 × 104
receptors/cell. As expected, decreasing RT leads
to a lower cAMP steady-state level. With the smallest
RT value (0.5 × 104), this
cAMP level corresponds to a steady-state value of the desensitization
rate
being much lower (0.02/s) than the saturation value
(0.06/s).

View larger version (33K):
[in this window]
[in a new window]
|
Fig. 4.
Influence of RT, , and
kPDE. Top: influence of
RT on changes in cAMP levels (left)
and on changes in (right); inset: nos. are in
molecules/cell. Middle: influence of on changes in
activated adenylyl cyclase levels (left) and on changes in
cAMP levels (right); inset: nos. are
dimensionless. Bottom: influence of
kPDE on changes in cAMP levels (left)
and on changes in the phosphorylation rate (right);
inset: nos. are in kPDE (/s). In this
figure and in Figs. 5-9, the various panels give the time
evolution of the various concentrations that constitute the variables
in the model. These are expressed in units of 104
molecules/cell (see Table 1). The phosphorylation rate is shown as
a function of time. Figures 5-8, top left, show the
pattern of applied FSH input (10 10 M; this is constant in
Figs. 5-7). Differently styled lines on each plot of the same
figure correspond to different values of the model parameter under
study (displayed in insets), the other parameters being kept
unchanged. The basic set of common parameter values is summarized in
Table 2, and common initial values are RT = 2.0, basal level of adenylyl cyclase (E0) = 0.01, and
cAMP = 0.25 (104 molecules/cell), with constant
FSH input of 3 × 10 10 M. Time units are 102
s, so the simulations correspond to periods of between ~8 h and 5 days.
|
|
Influence of the binding dissociation constant.
The increase in the dissociation constant,
Kd = k
/k+, leads to an
increase in the free receptor concentration together with a decrease in
the concentration of bound receptors and its derived (phosphorylated
and internalized) forms. This again affects the steady-state values
cAMP* and
*, with the highest value of
Kd corresponding to the lowest cAMP level.
Influence of the amplification parameter.
Figure 4 (middle) illustrates the role of the amplification
parameter
. The patterns of changes in EFSH
and cAMP are almost superimposed. The scale of the cAMP
value range is dramatically increased as
increases.
Influence of the hydrolysis parameter.
A nonzero value of kPDE is necessary for the
cAMP concentration to reach a steady-state value. If
kPDE = 0, as can be seen on the solid line
of Fig. 4 (bottom), signal turn-off is mediated only
by the phosphorylation function
, which quickly reaches its
saturation value (
) and cannot control the exponential increase in
cAMP concentration. Conversely, increasing the value of
kPDE affects cAMP levels so as to stabilize
at a value far below saturation.
Influence of the phosphorylation saturation parameter.
Changes in the saturation capacity of the desensitization function
affect not only the steady-state level of cAMP but also the different
forms of FSH receptors, as can be seen in Fig.
5. Increasing the value of
reduces
the number of FSH receptors in the bound active state
XFSH in favor of the phosphorylated state
XpFSH. The associated increase in internalized
receptors cannot compensate for this imbalance even if the
internalization process is at the source of free receptor renewal and
hence, indirectly, of active bound receptors.

View larger version (21K):
[in this window]
[in a new window]
|
Fig. 5.
Influence of the desensitization saturation capacity (/s).
Top left: changes in the levels of active bound FSH
receptors (XFSH); top right: changes
in cAMP levels; bottom left: changes in the
phosphorylation rate; bottom right: changes in the levels of
phosphorylated bound FSH receptors (XpFSH).
|
|
Influence of receptor renewal.
Increasing either the internalization rate ki or
the recycling rate kr allows for a quicker
renewal of free FSH receptors from phosphorylated bound receptors, thus
enhancing the FSH signal.
Influence of the time scale parameter.
Low
values (
1) lead to a marked contrast between the
dynamics of fast (RFSH,
XFSH, cAMP, and
XpFSH) and slow (EFSH) variables, whereas high values (i.e., not much lower than 1) tend to
homogenize the time scales of all the variables. As
increases from
a small value toward 1, the time required to reach equilibrium is
significantly decreased, as can be seen in Fig.
6. Thus, for
close to 1, the
steady state is reached in a few minutes, whereas, for small values (as
low as 10
3 in this instance), it can take several days.
The maximal value reached by EFSH and
cAMP can significantly overshoot its steady-state value, and
this effect also becomes more pronounced as
increases toward 1. The
transient response is sensitive to even small variations in the value
of
, especially for given parameter combinations. This is
illustrated in Fig. 7, where the
hydrolysis rate is about twice what its value is in other figures
(0.09/s). The time derivative of EFSH
changes signs, and thus crosses its steady-state value, one or more
times depending on the precise value of
, subsequently leading to a
variety of cAMP-transient patterns.

View larger version (25K):
[in this window]
[in a new window]
|
Fig. 6.
Influence of the time scale parameter (large
variations), showing from left to right and from
top to bottom the correspondence of each panel to
the next: FSH input, changes in the levels of free FSH receptors, bound
FSH receptors, activated adenylyl cyclase, cAMP, phosphorylated
receptors, internalized receptors, phosphorylation rate. Inset: nos.
are dimensionless.
|
|

View larger version (29K):
[in this window]
[in a new window]
|
Fig. 7.
Influence of (small variations), showing from
left to right and from top to
bottom the correspondence of each panel to the next: FSH
input, changes in the levels of free FSH receptors, bound FSH
receptors, activated adenylyl cyclase, cAMP, phosphorylated receptors,
internalized receptors, phosphorylation rate. Inset: nos.
are dimensionless.
|
|
Level and pattern of FSH input.
In the physiological range from 0.3 to 3.0 × 10
10
M, a 10-fold variation in FSH concentration (Fig.
8, top) results in a twofold variation in cAMP level. Beyond a given level of FSH (depending on the
values of the other parameters), increasing the FSH level will have
almost no effect on the cAMP response. We also investigated the pattern
of cAMP response to nonconstant FSH inputs (results not shown). In the
case of a square-shaped FSH input, the system switches from one steady
state to another as FSH switches between its high and low values. The
reaction of the system to the changes in FSH input is again under the
control of
. The smaller
is, the smoother the changes in the
system variables, to the extent that the effects of the variation in
FSH input may be perceptible only in the behavior of the receptor
species (RFSH, XFSH,
XpFSH, Ri). Other
simulations with exponentially decreasing or sinusoidal FSH input
yielded qualitatively similar conclusions. We show in Fig.
9 an example of the model behavior in
response to real FSH data taken from Adams et al. (2; Fig. 1B,
p. 631). The changes in FSH input are mirrored in those of free
FSH receptor concentration, whereas they are quite tightly tracked by
those in bound FSH receptors. The changes in phosphorylated and
internalized receptors are nearly similar and follow the
phosphorylation rate, which starts rising only when significant cAMP
levels have been reached. The changes in activated adenylyl cyclase and
cAMP are smoother. After increasing in a continuous way, they end up
oscillating in a dampened manner around a steady-state value.

View larger version (28K):
[in this window]
[in a new window]
|
Fig. 8.
Response to different levels of FSH stimulation, showing
from left to right and from top to
bottom the correspondence of each panel to the next: FSH
input, changes in the levels of free FSH receptors, bound FSH
receptors, activated adenylyl cyclase, cAMP, phosphorylated receptors,
internalized receptors, phosphorylation rate. Inset: nos. are in moles
per liter.
|
|

View larger version (21K):
[in this window]
[in a new window]
|
Fig. 9.
Response to nonconstant FSH input (real data), showing
from left to right and from top to
bottom the correspondence of each panel to the next: FSH
input, changes in the levels of free FSH receptors, bound FSH
receptors, activated adenylyl cyclase, cAMP, phosphorylated receptors,
internalized receptors, phosphorylation rate.
|
|
 |
DISCUSSION |
Our model is concerned with the cAMP dynamics resulting from FSH
signal transduction in average granulosa cells of maturing ovarian
follicles. We have chosen to focus on the dynamics of coupling between
FSH receptor stimulation and adenylyl cyclase activation and have
assumed that, on average, the cell has a constant pool of receptors.
This is consistent with experimental observations that, in the first
part of terminal follicular development, the increased response of cAMP
production to FSH occurs in the absence of significant changes in the
number of FSH-binding sites per granulosa cell (29).
Instead, the changes in signal transduction associated with follicular
development appear to affect the adenylyl cyclase enzyme system. This
is corroborated by investigations on cell lines expressing the FSH
receptor, in which the FSH-dependent accumulation of cAMP is highly
variable but not correlated with the receptor density (reviewed in Ref.
31) and may be due to the different coupling efficiency in
the different cell lines.
At the scale of a single cell, the assumption of a constant pool of
receptors implies that a newborn daughter cell doubles its inherited
pool of receptors (which can be considered roughly one-half of the pool
of its mother cell) during the first part of G1 after
completion of mitosis. At the scale of the whole follicle, the number
of receptors increases proportionally with the increase in granulosa
cell number.
The notion of an average cell follows from the experimental means of
investigating cAMP levels throughout follicular development. The common
way of measuring cAMP is to dissect follicles and pool granulosa cells
so that, even if the total number of granulosa cells increases with
follicular maturation, the results are expressed as average
concentrations per given number of cells (usually 105 or
106) (1, 13, 19, 20). Such an average
viewpoint also takes into account smoothing interactions between
granulosa cells such as the exchange of cAMP molecules through gap
junctions. Because the cycles of granulosa cells appear to be fully
desynchronized, the average description does not need to consider the
various phases of the cell cycle and moreover allows us to take into
account the heterogeneous features of cell states, including the
nondividing state.
To understand the nature of the most appropriate data for the model, we
recall here what it really performs. The model retraces the long-term
behavior of cAMP in granulosa cells during terminal follicular
development in response to FSH alone. It is interested in follicles
from their entering the FSH-responsive stage. For instance, in the ewe,
this stage corresponds to a 1-mm diameter, compared with the 7.5-mm
diameter of ovulatory follicles (monoovulating breeds). The output of
the model is the cAMP level as a function of the follicle's age in
response to a given pattern of FSH input. Detailed analysis has been
made for constant input, but real FSH data can also be handled.
The most appropriate data would consist of repeated measurements of
intracellular cAMP throughout the development of dynamically monitored
follicles. Concomitantly, FSH levels should be measured. Such data
could be directly exploited from the FSH-responsive to the
LH-responsive stage. Once LH receptors appear on granulosa cells, cAMP
production is a mixed response to both FSH and LH stimulation. To track
cAMP production until the ovulatory stage, one needs to be placed in
controlled situations. In physiological situations, the luteal
phase in ruminant species would be the most appropriate window on the
ovarian cycle to harvest data. In such species, there is no estradiol
secretion from the corpus luteum, so follicular growth proceeds
normally until the preovulatory size; yet LH pulsatility is low, due to
high progesterone levels, so the follicles are prevented from
ovulating. Hence, terminal follicular development during the luteal
phase is mainly FSH dependent and thus fulfills the requirements for
investigating FSH-induced cAMP production throughout terminal
development. In pharmacological situations, appropriate conditions can
be reproduced artificially by means of either previous desensitization
with gonadotropin-releasing hormone (GnRH) agonists, or use of GnRH
antagonists and administration of recombinant FSH with known
bioactivity. The most limiting point in both situations is the need for
dynamic, noninvasive measurements of intracellular cAMP. This might be
achievable in the future through repeated ultrasound-guided follicular
cell pickup as follicular development progresses. In domestic animals,
follicular fluid pickup is already running well, and technical progress
in devices may render direct cell pickup feasible in the medium term.
Some data on the long-term evolution of cAMP are, nevertheless, already
available. The most interesting ones (13) provide information about the trends in cAMP production throughout follicular development. Unfortunately, they are not straightforward enough to
handle on quantitative grounds, because cAMP concentrations are
expressed against the follicular diameter. Besides, they need ovariectomy and dissection of the follicles, whose granulosa cells are
pooled, so they do not allow individual or dynamic study of the
follicles. cAMP levels may also be overestimated because of the use of IBMX. Despite this lack of appropriate data, we constrained the numerical application of the model as much as we could, as is
detailed in Dimension of model variables and paramenters.
In summary, the simulations presented in Control of FSH-Induced
cAMP Levels have investigated the possible alterations in the
capacity of the granulosa cell for FSH signal 1) detection, 2) relay and amplification, and 3) overflow
control. The capacity for signal detection depends on the size of the
receptor pool and on the dissociation rate, which is presumably
constant within a single individual but could be subject to inter- or
even intraspecies differences due to the existence of gene mutations or
polymorphisms. The capacity for signal relay and amplification is
mediated mainly through the coupling variable
EFSH for activated adenylyl cyclase and the
parameter. At this stage, our formulation of the dynamics of
EFSH is a "black box," subject to biological
constraints. From available knowledge, one can only speculate on the
underlying mechanisms. They could imply, for instance, a modulation in
the intrinsic GTPase activity of different splice variants of
G
s subunits (9) or a shift in the balance
between FSH receptor coupling with G
s (activating
adenylyl cyclase) and FSH receptor coupling with G
i
(inhibiting adenylyl cyclase). The design of new experiments
would help to answer the question. They could consist, for instance, of
measuring FSH-induced cAMP responses in cultured granulosa cells
derived from small compared with large follicles after adding either
cholera (inhibiting GTPase activity) or pertussis (inhibiting
G
i subunits) toxins in the culture medium. If a decrease
in G
i or GTPase activity occurs during terminal follicular development, one would expect to observe a greater rise in
the cAMP response of treated granulosa cells, compared with control
cells, from small follicles than in the response of treated cells from
large follicles. The capacity for signal overflow control is exerted by
means of both receptor desensitization and cAMP hydrolysis. The balance
between cAMP synthesis and hydrolysis (corresponding to
/kPDE) could be related to different isoforms of adenylyl cyclase and PDE enzymes. The processes of desensitization and hydrolysis act as protection mechanisms of the granulosa cells against overstimulation and cAMP overflow, resulting in the control of
the maximal reachable value of the intracellular cAMP level as well as
of the speed in reaching a prescribed value. Early elevated levels of
intracellular cAMP are known to have deleterious effects such as
precocious luteinization (34). Although mathematically the
system remains at this equilibrium for all time, in reality this is not
the case, because the cAMP-induced expression of LH receptors around
the time of selection will again modify the pattern of cAMP production.
We intend to incorporate this effect in our model in the future.
The model's result that cAMP behaves as an increasing, saturating
function of FSH is compatible with the observation that FSH-induced
estradiol production drops for large FSH concentrations after the
ovulatory surge (22). First, this drop may concern only
that part of the cascade downstream of cAMP. In particular, differential regulation of protein kinase A regulatory subunits (18) could be involved, which would control the expression
level of aromatase (but would not prevent LH from maintaining
steroidogenesis, because it has direct, not only genic, effects such as
enhancing the entry of cholesterol into the mitochondria). Second,
cross talks among the various cellular signaling pathways may be of greater importance after the follicle has acquired LH receptors and
affect estradiol response to FSH. Finally, the fully cAMP-dependent desensitization process of the model may be inadequate to describe what
happens in the presence of large FSH concentrations. At high agonist
concentration, G protein-coupled receptor kinases could also be
implicated, together with the proteins of the arrestin family, in the
phosphorylation of FSH receptors, as has been established for the
2-adrenergic receptor (3).
Control of the dynamics of cAMP production in granulosa cells is
a key point in the regulation of terminal follicular development. Yet,
as far as we are aware, manipulation of the different steps of the cAMP
cascade is not used as a way of controlling ovarian function either in
domestic animals or in humans. A realistic model would thus be of great
help in developing new strategies for the control of follicle
maturation and understanding pathophysiological situations such as
those encountered in polyovulating models. The Booroola Merino is a
breed of sheep carrying a major gene that influences its ovulation
rate. Homozygous (F/F) and heterozygous (F/+)
carriers and noncarriers (+/+) of the gene have ovulation rates of
5, 3 or 4, and 1 or 2, respectively. Comparative
studies (19) have shown that the F gene induces
specific differences in follicular development because of the granulosa
cells from F/F and F+ ewes being more
responsive to FSH and/or LH than granulosa cells from +/+
ewes with respect to cAMP synthesis. Similar observations have been
made in polyovulating Romanov breeds compared with monoovulating Ile de
France breeds (1). Furthermore, FSH-induced cAMP response was clearly greater in Romanov ewes, although the number of FSH receptors was similar, suggesting a more efficient coupling between FSH
receptors and adenylyl cyclase in this breed. Interestingly, deregulation of cAMP production also seems to occur in women suffering from polycystic ovary syndrome. Premature generation of preovulatory concentrations of cAMP in granulosa cells could be at the source of
anovulation (37).
The model combines biochemical and physiological angles of FSH
action on granulosa cells. It helps us gain a better understanding of
the dynamic control of cAMP synthesis in granulosa cells during terminal follicular development. It allows us to investigate how FSH
concentrations will affect the responsiveness of follicles in terms of
cAMP production. Variations in the model parameter values correspond to
physiological or pathological alterations in the different steps of FSH
signal processing by granulosa cells. The resulting cAMP dynamics, in
turn, control the commitment of granulosa cells to proliferation,
differentiation, or apoptosis. Thus understanding FSH-induced
cAMP dynamics is a first step in understanding how FSH controls
granulosa cell behavior on the scale of the whole follicle.
The main improvement to the present model would consist of giving a
fully biochemically based formulation to the equation describing
changes in coupling efficiency. For the moment, this remains beyond
reach, because little is known about the biochemical mechanisms
underlying the increase in the cAMP response of granulosa cells to FSH
stimulation. Obviously, the balance between activation and inactivation
of G proteins is implicated in this mechanism, but further experimental
investigation is needed before a more realistic model can be built.
Incorporating LH signaling is also an important challenge, because
synergistic signaling by FSH and LH seems to be the basis for the
selection of the ovulatory follicle(s).
Whereas inadequate response of granulosa cells to gonadotropin signals
may have major repercussions on follicular development and may even
lead to infertility, a realistic model characterizing both
physiological and pathological signal transduction would be very useful
for simulating the development of new therapeutic strategies.
 |
APPENDIX |
Upper Bounds of Solutions
From Eqs. 1-6, one can see that, if any one of
the variables RFSH, XFSH,
cAMP, XpFSH,
EFSH, and Ri is zero and
the other variables are nonnegative, then that variable that is zero is
nondecreasing. It immediately follows that, if the system starts in the
physiologically relevant region with all variables nonnegative, then it
remains there for all time. With the conservation equation Eq. 7, this implies that RFSH
RT, XFSH
RT, XpFSH
RT, and Ri
RT. From XFSH
RT and Eq. 10, it follows that
FSH
(
RT
EFSH)EFSH. Hence, if
EFSH(0)
RT, then
EFSH(t)
RT for all t
0, whereas if
EFSH(0)
RT, then, for any
> 0, there exists
a T
0 such that
EFSH(t)
RT +
for all t
T. Substituting this into Eq. 11, we see that as
long as kPDE > 0, then for any
> 0 there exists a T
0 such that
for all t
T.
Existence and Uniqueness of Strictly Positive Roots of Eq. 19
We first notice that x = 0 cannot be a root of
Eq. 19. For x > 0, let
and
The roots of Eq. 19 are the points of intersection
between the curves of f and g. It can be seen
clearly that f(x) is a strictly decreasing
function of x that tends toward zero as x tends
toward infinity and tends toward infinity as x tends toward
zero. On the other hand, g(x) is an increasing
and bounded positive function of x. It follows that there
can be only one intersection point for x > 0 and that
Eq. 19 admits one and only one strictly positive root. This
geometric reasoning is illustrated in Fig.
10.

View larger version (9K):
[in this window]
[in a new window]
|
Fig. 10.
Illustration of the search for the intersection point S between
the hyperbola passing through point H (1, RT)
and the increasing sigmoid curve bounded by (A, A + B), which
corresponds to the root of Eq. 19. In Figs. 10 and 11, A and
B are given by
|
|
Control of Steady-State Level cAMP*
Influence of FSH and model parameters except
.
The influence of FSH and the model parameters on cAMP* is
effected through their influence on
X*FSH. Let
X*
and
X*
be the corresponding steady-state values on
XFSH when applying respective input
FSH1 and FSH2, with FSH1 < FSH2. Suppose that
X*
X*
. Because
FSH1 < FSH2, we have
Because
* is an increasing function of
X*FSH
Substituting into Eq. 19 we obtain
which is a contradiction. Hence, we must have
X*
> X*
whenever FSH1 < FSH2, so that
The same reasoning applied to the parameters leads to
and
Influence of
.
Because X*FSH > 0, let
y = 1/ X*FSH.
Equation 19 can then be written as
Furthermore, let
then
|
(A1)
|
Let z*1 and
z*2 be the roots of Eq. A1
respectively corresponding to
1 and
2,
with
1 <
2. Both are given through
Eq. 20 as the points of intersection of the straight line
whose slope is the left-hand term in front of z, with the
curve representing the right-hand term function of z. The
search for this intersection point is illustrated in Fig.
11.

View larger version (10K):
[in this window]
[in a new window]
|
Fig. 11.
Illustration of the search for the intersection point between the
straight line of slope
 RT/( kPDE) with
the decreasing sigmoid curve bounded by (A, A + B), which corresponds
to the root of Eq. 20. The straight line passing through
point J, whose coordinates are (1, A + B/2), delimits 2 distinct areas.
For any intersection point whose abscissa is <1, such as I, the
steady-state level of cAMP, cAMP*, decreases as
the value of increases. On the other hand, for any intersection
point whose abscissa is >1, such as K, this level increases as the
value of increases.
|
|
If
both roots are <1. Hence, z
1 > z
2, so that
It follows that z*1 < z*2. Recalling that z =
kPDE/(
X*FSH),
we can conclude that X
> X
and that
On the other hand, if
then
Hurwitz Criterion for Linear Stability Analysis
The dependence of
* on cAMP* can be neglected when
the value of
* is close to that of
. In this case, the 
*
term vanishes, and the Jacobian matrix Mj can be rewritten
as
The eigenvalues of Mj are given by the roots of the
characteristic equation
We can build the following sequence of determinants associated
with the ais
The signs of these determinants can be determined using a
symbolic manipulation package such as Maple. This allows us to show
that all of the determinants are positive, which, by the Hurwitz
criterion (5), is a necessary and sufficient condition for
all the eigenvalues of Mj to have strictly negative real
parts. This, in turn, implies that the corresponding equilibrium is
asymptotically stable.
Local Controllability of the System
In this section, we consider FSH a control variable. The
linearization of system 8-12 about the steady state can
be written after separation of the state and control variables as
|
(A2)
|
where q = (RFSH,
XFSH, EFSH,
cAMP, XpFSH)T,
u = FSH, the Jacobian matrix Mj defines the
drift vector field, and B = (
k+R
, k+R
, 0, 0, 0)T is the input vector field. The
controllability matrix associated with Eq. A2 is the square
matrix whose columns are given by
Formal calculation shows that the determinant of C is nonzero,
which is a necessary and sufficient condition for the linearized system
around the steady state to be controllable. It follows that nonlinear
system 8-12 is locally strongly accessible, which is
confirmed by the study of the strong accessibility Lie algebra (24).
A similar analysis applied to the zero steady state characterized by
EFSH = 0 concludes that neither the
linearized form of the system nor the nonlinear one is controllable;
thus EFSH = 0 is a strongly degenerate
point for the system.
 |
FOOTNOTES |
Address for reprint requests and other correspondence: F. Clément, Institut National de Recherche en Informatique et
Automatique, Unité de Recherche de Rocquencourt, Domaine de
Voluceau, Rocquencourt, BP 105, 78153 Le Chesnay Cedex, France (E-mail:
Frederique.Clement{at}inria.fr).
The costs of publication of this
article were defrayed in part by the
payment of page charges. The article
must therefore be hereby marked
"advertisement"
in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.
Received 30 June 2000; accepted in final form 14 February 2001.
 |
REFERENCES |
1.
Abdennebi, L,
Monget P,
Pisselet C,
Remy JJ,
Salesse R,
and
Monniaux D.
Comparative expression of luteinizing hormone and follicle-stimulating hormone receptors in ovarian follicles from high and low prolific sheep breeds.
Biol Reprod
60:
845-854,
1999[Abstract/Free Full Text].
2.
Adams, GP,
Matteri RL,
and
Ginther OJ.
Effect of progesterone on ovarian follicles, emergence of follicular waves and circulating follicle-stimulating hormone in heifers.
J Reprod Fertil
95:
627-640,
1992.
3.
Aittomaki, K,
Lucena JL,
Pakarinen P,
Sistonen P,
Tapaneinen P,
Gromoll J,
Kaskikari R,
Sankila EM,
Lehvaslaiho H,
Engel AR,
Nieschlag E,
Huhtaniemi I,
and
de la Chapelle A.
Mutation in the follicle-stimulating hormone receptor gene causes hereditary hypergonadotropic ovarian failure.
Cell
82:
959-968,
1995[ISI][Medline].
4.
Ascoli, M.
Functional consequences of the phosphorylation of the gonadotropin receptors.
Biochem Pharmacol
52:
1647-1655,
1996[ISI][Medline].
5.
Carson, RS,
Findlay JK,
Burger HG,
and
Trounson AO.
Gonadotropin receptors of the ovine ovarian follicle during follicular growth and atresia.
Biol Reprod
21:
75-87,
1979[ISI][Medline].
6.
Cesari, L.
Asymptotic Behaviour and Stability Problems in Ordinary Differential Equations. Berlin: Springer-Verlag, 1971.
7.
Clément, F.
Optimal control of the cell dynamics in the granulosa of ovulatory follicles.
Math Biosci
152:
123-142,
1998[ISI][Medline].
8.
Clément, F,
Gruet MA,
Monget P,
Terqui M,
Jolivet E,
and
Monniaux D.
Growth kinetics of the granulosa cell population in ovarian follicles: an approach by mathematical modelling.
Cell Prolif
30:
255-270,
1997[ISI][Medline].
9.
Clément, F,
Thalabard JC,
and
Claude D.
Le système de reproduction vu sous l'angle de la commande: modélisation et commande de la fonction ovarienne.
In: Modélisation et Commande de Régulations Biologiques. Paris: Journée Thématique CNRS, 1999.
10.
DiMeglio, LA,
and
Pescovitz OH.
Disorders of puberty: inactivating and activating molecular mutations.
J Pediatr
131:
S8-S12,
1997[ISI][Medline].
11.
Europe-Finner, GN,
Cartwright E,
Bellinger J,
Mardon HJ,
Barlow DH,
and
López Bernal A.
Identification of G
s messenger ribonucleic acid splice variants in human granulosa cells.
J Mol Endocrinol
18:
27-35,
1997[Abstract].
12.
Goodman, RL.
Neuroendocrine control of the ovine estrous cycle.
In: The Physiology of Reproduction, edited by Knobil E,
and Neill JD.. New York: Raven, 1994, p. 659-709.
13.
Graña, X,
and
Reddy EP.
Cell cycle control in mammalian cells: role of cyclins, cyclin dependent kinases (CDKs), growth suppressor gene and cyclin-dependent kinase inhibitors (CKIs).
Oncogene
11:
211-219,
1995[ISI][Medline].
14.
Greenwald, G,
and
Roy S.
Follicular development and its control.
In: The Physiology of Reproduction, edited by Knobil E,
and Neill JD.. New York: Raven, 1994, p. 629-724.
15.
Henderson, KM,
Kieboom LE,
Mac Natty KP,
Lun S,
and
Heath D.
Gonadotrophin-stimulated cyclic AMP production by granulosa cells from Booroola × Romney ewes with and without a fecundity gene.
J Reprod Fertil
75:
111-120,
1985[Abstract].
16.
Hoffman, JF,
Linderman JJ,
and
Omann GM.
Receptor up-regulation, internalization and interconverting receptor states. Critical components of a quantitative description of N-formyl peptide-receptor dynamics in the neutrophil.
Biol Chem
271:
18394-18404,
1996[Abstract/Free Full Text].
17.
Jolly, PD,
Tisdall DJ,
De'ath G,
Heath DA,
Lun S,
Hudson NL,
and
MacNatty KP.
Granulosa cell apoptosis, aromatase activity, cyclic adenosine 3',5'-monophosphate response to gonadotrophins, and follicular fluid steroid levels during spontaneous and induced follicular atresia in ewes.
Biol Reprod
56:
830-836,
1997[Abstract].
18.
Katok, A,
and
Hasselblatt B.
Introduction to the Modern Theory of Dynamical Systems. Cambridge, UK: Cambridge Univ. Press, 1997.
19.
Knecht, M,
Ranta T,
and
Catt KJ.
Granulosa cell differentiation in vitro: induction and maintenance of follicle-stimulating hormone receptors by adenosine 3',5'-monophosphate.
Endocrinology
113:
949-956,
1983[Abstract].
20.
Kwast-Welfeld, J,
and
Jungmann RA.
Hormonal regulation of nuclear cyclic AMP-dependent protein kinase subunit levels in rat ovaries.
J Biol Chem
263:
14343-14350,
1988[Abstract/Free Full Text].
21.
MacNatty, KP,
and
Henderson KM.
Gonadotrophins, fecundity genes and ovarian follicular function.
J Steroid Biochem
27:
365-373,
1987[ISI][Medline].
22.
MacNatty, KP,
Kieboom LE,
MacDiarmid J,
Heath DA,
and
Lun S.
Adenosine cyclic 3',5'-monophosphate and steroid production by small ovarian follicles from Booroola ewes with and without a fecundity gene.
J Reprod Fertil
76:
471-480,
1986[Abstract].
23.
Mason, HD,
Mannaerts B,
de Leeuw R,
Willis DS,
and
Franks S.
Effects of recombinant human follicle-stimulating hormone on cultured human granulosa cells: comparison with urinary gonadotrophins and actions in preovulatory follicles.
Hum Reprod
8:
1823-1827,
1993[Abstract].
24.
Nakamura, K,
Hipkin RW,
and
Ascoli M.
The agonist-induced phosphorylation of the rat follitropin receptor maps to the first and third intracellular loops.
Mol Endocrinol
12:
580-591,
1998[Abstract/Free Full Text].
25.
Nijmeijer, H,
and
van der Schatf AJ.
Nonlinear Dynamical Control Systems. Berlin: Springer-Verlag, 1990.
26.
Peluso, JJ,
Pappalardo A,
and
White BA.
Control of rat cell mitosis by phorbol ester-, cyclic AMP- and estradiol-17
-dependent pathways.
Biol Reprod
49:
416-422,
1993[Abstract].
27.
Perrotin, F,
Royere D,
Roussie M,
Combarnous Y,
Lansac J,
and
Muh JP.
Characterization of human follicle-stimulating hormone binding to human granulosa cells by an immunoenzymological method.
Anal Biochem
202:
71-75,
1992[ISI][Medline].
28.
Press, WH,
Flannery BP,
Teukolsky SA,
and
Vetterling WT.
Numerical Recipes in C, The Art of Scientific Computing. Cambridge, UK: Cambridge Univ. Press, 1990.
29.
Richards, JS.
Maturation of ovarian follicles: actions and interactions of pituitary and ovarian hormones on follicular cell differentiation.
Physiol Rev
60:
51-89,
1980[Free Full Text].
30.
Richards, JS,
and
Hedin L.
Molecular aspects of hormone action in ovarian follicular development, ovulation and luteinization.
Annu Rev Physiol
50:
441-463,
1988[ISI][Medline].
31.
Simoni, M,
Gromoll J,
and
Nieschlag E.
The follicle-stimulating hormone receptor: biochemistry, molecular biology, physiology and pathophysiology.
Endocr Rev
18:
739-773,
1997[Abstract/Free Full Text].
32.
Starbuck, C,
and
Lauffenburger DA.
Mathematical model for the effects of epidermal growth factor receptor trafficking dynamics on fibroblast proliferation responses.
Biotechnol Prog
8:
132-143,
1992[ISI][Medline].
33.
Taussig, R,
and
Gilman AG.
Mammalian membrane-bound adenylyl cyclases.
J Biol Chem
270:
1-4,
1995[Free Full Text].
34.
Themmen, AP,
Martens JW,
and
Brunner HG.
Gonadotropin receptor mutations.
J Endocrinol
153:
179-183,
1997[Free Full Text].
35.
Ubaldi, F,
Camus M,
Bennink HC,
Van Steirteghem A,
and
Devroey P.
Premature luteinization in in vitro fertilization cycles using gonadotrophin-releasing hormone agonist (GnRH-a) and recombinant follicle-stimulating hormone (FSH) and GnRH-a and urinary FSH.
Fertil Steril
66:
275-280,
1996[ISI][Medline].
36.
Ward, DN,
Bousfield GR,
and
Moore KH.
Gonadotropins.
In: Reproduction in Domestic Animals (4th ed.), edited by Cupps PT.. London: Academic, 1991, p. 25-80.
37.
Webb, R,
and
England BG.
Identification of the ovulatory follicle in the ewe: associated changes in follicular size, thecal and granulosa cell luteinizing hormone receptors, antral fluid steroids, and circulating hormones during the preovulatory period.
Endocrinology
110:
873-881,
1982[ISI][Medline].
38.
Willis, DS,
Watson H,
Mason HD,
Galea R,
Brincat M,
and
Franks S.
Premature response to luteinizing hormone of granulosa cells from anovulatory women with polycystic ovary syndrome: relevance to mechanism of ovulation.
J Clin Endocrinol Metab
83:
3984-3991,
1998[Abstract/Free Full Text].
Am J Physiol Endocrinol Metab 281(1):E35-E53
0193-1849/01 $5.00
Copyright © 2001 the American Physiological Society