Main menu User menu Search Journal of Experimental BiologySUMMARYIntroductionRegulation of signaling by membrane recruitment of cytosolic proteinsSignal transduction through MAPK cascades can require endocytic trafficking and/or active molecular transportLong-range signalingConcluding remarksAbbreviationsACKNOWLEDGEMENTSReferencesCitation Manager FormatsArticle navigationOther journals from The Company of BiologistsEditors’ choice - Mitochondrial plasticity in the cerebellum of two anoxia-tolerant sharks: contrasting responses to anoxia/re-oxygenationFeatured article – Newt chewing is all in the tongueCommentary - Loss-of-function approaches in comparative physiology: is there a future for knockdown experiments in the era of genome editing?News - So long Raul and welcome TrishScientific Meeting GrantsArticlesAbout usFor AuthorsJournal InfoContact
Hochachka, 1999Hollenbeck, 2001Verhey and Rapoport, 2001Chang and Karin, 2001Kholodenko, 2002Lewis et al., 1998Kholodenko, 2002Mochly-Rosen, 1995Egan et al., 1993Fig.1Schlessinger, 2000Schlessinger and Bar-Sagi,
1994Buday and Downward,
1993Sastry et al., 1995Adam and Delbruck, 1968Axelrod and Wang, 1994Berg and Purcell, 1977Kholodenko
et al., 2000bKholodenko et al., 2000bCherry, 1979Kusumi et al., 1993Bray, 1998Ferrell, 19981997Axelrod and Wang, 1994Rohwer et al., 1998Kholodenko et al., 2000bKholodenko et al., 1999Moehren et al., 2002Buday and Downward,
1993Corbalan-Garcia et al.,
1998Lenzen et al.,
1998Bray, 1998Haugh and
Lauffenburger, 1997Drugan et al., 2000Ugi et al., 2002Bollag
and McCormick, 1992Kouhara et al., 1997Kouhara et al.,
1997Chang and Karin,
2001Lewis et al.,
1998Garrington
and Johnson, 1999Gutkind,
1998van Biesen et al.,
1996Fig.1Pierce et al., 2001Prenzel et al., 1999Andreev et al., 2001Miller and Lefkowitz,
2001Mason et al., 1999Kyriakis et al.,
1992Strack et al.,
1997Brown
and Kholodenko, 1999Kholodenko et al., 2000aFig. 2Kholodenko,
2002Kholodenko et al., 2000aFig.2Arrio-Dupont et al., 2000Dayel et al., 1999Gershon et al., 1985Jacobson and Wojcieszyn,
1984Haugh and Lauffenburger, 1998Kholodenko et al., 2000aZhao and Zhang, 20012002Di Guglielmo et
al., 1994Koenig and Edwardson, 1997Vieira et al., 1996Ceresa et al., 1998Daaka et al., 1998Kranenburg et al., 1999Maudsley et al., 2000Rizzo et al., 2001Vieira et al., 1996Clague and Urbe,
2001Di Fiore and De Camilli,
2001Haugh et al.,
1999Ceresa and Schmid,
2000Kranenburg et al.,
1999Pierce et al.,
2000Leof, 2000Kholodenko, 2002Kranenburg et al., 1999Kranenburg et al.,
1999Agutter et al., 1995Garrington and Johnson, 19992000Miller and Lefkowitz, 2001Pierce et al., 2001Verhey and Rapoport,
2001Ginty and Segal, 2002Reynolds et al., 1998Yano et al., 2001Reynolds et al., 1998MacInnis and Campenot, 2002Ginty and Segal, 2002Verveer et al., 2000Thomas et al., 2003Murray, 1993Sawano et al.,
2002Sawano et al., 2002Thomas et al.,
2003Bhalla and Iyengar, 1999Ferrell and Machleder, 1998Gardner et al., 200019971999Tyson and Novak, 2001Bagowski and Ferrell, 2001Ferrell, 1999Bhalla et
al., 2002Ferrell, 2002Castiglione et al., 2002Christoph et al., 1999Keener and Sneyd, 1998Shvartsman et al., 2002Zhabotinsky and Zaikin, 1973Download PDFDevelopmentJournal of Cell ScienceDisease Models & MechanismsBiology Openshowresearch articleknockdown and knockout approachesannounceapply© 2019 The Company of Biologists Ltd
Skip to main content
var googletag = googletag || ;
googletag.cmd = googletag.cmd || [];
googletag.cmd.push(function()
googletag.defineSlot('/155913954/JEB', [728, 90], 'div-gpt-ad-1475847916783-0').addService(googletag.pubads());
googletag.pubads().enableSingleRequest();
googletag.enableServices();
);
googletag.cmd.push(function()
googletag.display("div-gpt-ad-1475847916783-0");
);
(adsbygoogle = window.adsbygoogle || []).push(
google_ad_client: "ca-pub-4431737545806467",
enable_page_level_ads: true
);
Main menu
COB- About The Company of Biologists
- Development
- Journal of Cell Science
- Journal of Experimental Biology
- Disease Models & Mechanisms
- Biology Open
User menu
- Log in
Search
- Advanced search
c.showpopup();
COB- About The Company of Biologists
- Development
- Journal of Cell Science
- Journal of Experimental Biology
- Disease Models & Mechanisms
- Biology Open
Journal of Experimental Biology
- Log in
- Home
Articles- Accepted manuscripts
- Issue in progress
- Latest complete issue
- Issue archive
- Archive by article type
- Special issues
- Subject collections
- Interviews
- Sign up for alerts
About us- About JEB
- Editors and Board
- Editor biographies
- Travelling Fellowships
- Grants and funding
- Journal Meetings
- Workshops
- The Company of Biologists
- Journal news
For authors- Submit a manuscript
- Aims and scope
- Presubmission enquiries
- Article types
- Manuscript preparation
- Cover suggestions
- Editorial process
- Promoting your paper
- Open Access
- Outstanding paper prize
- Biology Open transfer
Journal info- Journal policies
- Rights and permissions
- Media policies
- Reviewer guide
- Sign up for alerts
Contacts- Contact JEB
- Subscriptions
- Advertising
- Feedback
Journal of Experimental Biology 2003 206: 2073-2082; doi: 10.1242/jeb.00298
- Find this author on Google Scholar
- Find this author on PubMed
- Search for this author on this site
Article
Figures & tables
Info & metrics
PDF
SUMMARY
Extracellular signals received by membrane receptors are processed, encoded
and transferred to the nucleus via phosphorylation and spatial
relocation of protein members of multiple component pathways, such as mitogen
activated protein kinase (MAPK) cascades. The receptor-induced membrane
recruitment of the cytoplasmic protein SOS results in the activation of the
Ras/MAPK cascade. It has been suggested that the membrane recruitment of
signaling proteins causes an increase in the diffusion-limited rates. We have
recently shown that this increase is too small to be responsible for enhanced
signal transduction. Instead we demonstrate that the function of membrane
localization is to increase the number (or average lifetime) of complexes
between signaling partners. A hallmark of signaling pathways is the spatial
separation of activation and deactivation mechanisms; e.g. a protein can be
phosphorylated at the cell surface by a membrane-bound kinase and
dephosphorylated in the cytosol by a cytosolic phosphatase. Given the measured
values of protein diffusion coefficients and of phosphatase and kinase
activities, the spatial separation is shown to result in precipitous
phospho-protein gradients. When information transfer is hampered by slow
protein diffusion and rapid dephosphorylation, phospho-protein trafficking
within endocytic vesicles may be an efficient way to deliver messages to
physiologically relevant locations. The proposed mechanism explains recent
observations that various inhibitors of endocytosis can inhibit MAPK
activation. Additional mechanisms facilitating the relay of signals from
cell-surface receptors to the nucleus can involve the assembly of protein
kinases on a scaffolding protein and active transport of signaling complexes
by molecular motors. We also discuss long-range signaling within a cell, such
as survival signaling in neurons. We hypothesize that ligand-independent waves
of receptor activation or/and traveling waves of phosphorylated kinases emerge
to spread the signals over long distances.
- signal transduction
- protein kinase
- diffusion
- endocytosis
- molecular motor
- mitogen activated protein kinase (MAPK)
- traveling wave
Introduction
The deciphering of the genome of several organisms including humans has
generated a list of the macromolecular parts of living cells. A challenge of
current biology is to understand how this list of `genetics parts' gives rise
to a four-dimensional (i.e. space- and time-varying) behavior, governed by
intracellular regulatory networks. Activation of signal transduction networks
by extracellular stimuli is encoded into complex temporal and spatial patterns
of activation and relocation of numerous proteins, leading to important
cellular decisions ranging from cell survival, growth and proliferation, to
growth arrest, differentiation or apoptosis. During the past decade, there has
been increasing realisation that cytoskeletal structures and intracellular
movement, driven by myosin, kinesin and dyenin motors, play crucial roles in
the regulation of metabolism and signal transduction (see, for example,
reviews by Hochachka, 1999;
Hollenbeck, 2001;
Verhey and Rapoport, 2001).
Our current understanding of the spatio–temporal organization of
signaling processes and its control by cellular topology, diffusion and
intracellular movement, however, is far from complete.
Signaling through a plethora of cell-surface receptors, such as G-protein
coupled receptors (GPCRs), receptor tyrosine kinases (RTKs) and cytokine
receptors, activate mitogen activated protein kinase (MAPK) cascades, which
function as central integration modules in information processing
(Chang and Karin, 2001;
Kholodenko, 2002;
Lewis et al., 1998). MAPK
cascades relay extracellular stimuli from the plasma membrane to crucial
cellular targets distant from the membrane, e.g. transcription factors.
Elucidation of the spatio–temporal organization and regulation of MAPK
signaling is becoming increasingly important for understanding signaling
specificity and the diverse nature of cellular responses. We have recently
shown that simple diffusion of activated kinases may be insufficient for the
effective propagation of phosphorylation signals through MAPK cascades
(Kholodenko, 2002). In this
paper, we analyze the functional implications of membrane targeting of
cytosolic proteins, examine the consequences of the spatial separation of
certain kinases and phosphatases in MAPK pathways and discuss multiple
mechanisms used by a cell to propagate the signals over long distances.
Regulation of signaling by membrane recruitment of cytosolic
proteins
There are numerous examples of the membrane localization of cytoplasmic
proteins following receptor activation
(Mochly-Rosen, 1995). For
instance, stimulation of RTKs is linked to the activation of the extracellular
signal regulated kinase (ERK) Ras/MAPK cascade through the recruitment of a
cytoplasmic protein SOS to the plasma membrane
(Egan et al., 1993). SOS (a
homolog of the Drosophila melanogaster Son of sevenless protein) is
the guanine nucleotide exchange factor for the small GTP-binding protein Ras,
anchored to the cell membrane
(Fig.1). SOS-Ras interactions
convert inactive Ras-GDP to active Ras-GTP. SOS binds to RTKs, such as the
epidermal growth factor receptor (EGFR), not directly, but through the adaptor
protein Grb2 (growth factor receptor binding protein 2). Grb2, in turn, can
bind to the activated receptor either directly or through another adaptor
protein, Shc (src homology and collagen domain protein)
(Schlessinger, 2000;
Schlessinger and Bar-Sagi,
1994). Importantly, RTKs do not phosphorylate SOS, and SOS
catalytic activity towards Ras does not change upon its binding to the
receptor (Buday and Downward,
1993). Note also, that the Grb2-SOS complexes can exist in the
cytoplasm of unstimulated cells due to the high affinity of Src homology (SH)
3 domains of Grb2 to proline-rich regions of SOS
(Sastry et al., 1995).
- Download figure
- Open in new tab
- Download powerpoint
This interaction pattern raises a number of questions about the role of the
membrane recruitment in signal transfer. For instance, why should SOS relocate
to the membrane if its catalytic activity is not activated by RTKs? What
prevents direct interaction of cytosolic SOS or Grb2-SOS complexes with the
membrane-bound Ras from catalyzing GDP/GTP exchange on Ras? To clarify the
effect of membrane localization, we consider two extreme cases. When two
signaling molecules form a productive complex, transducing the signal after
each diffusive encounter, the signal-transduction process is called
`diffusion-limited'. If only a small fraction of the collisions leads to
binding that lasts long enough to transfer the information, the signal
transduction is called `reaction-limited'. In the latter case, the two protein
molecules associate and dissociate a number of times before signal
transduction takes place.
Membrane localization does not significantly enhance
diffusion-limited signal transfer
It has earlier been suggested that localization of signaling proteins close
to the cell membrane causes an increase in the first-encounter rate
(Adam and Delbruck, 1968); the
solutes do not `get lost' by wondering off into the bulk phase
(Axelrod and Wang, 1994;
Berg and Purcell, 1977).
However, such an increase appears to be too small to account for significantly
enhanced signal transduction (Kholodenko
et al., 2000b). The diffusion-limited rates in the membrane and in
the cytosol can be compared in terms of the ratio (h) of the
corresponding association rate constants in two dimensions and in three
dimensions. For diffusion-limited association of two proteins located in the
plasma membrane of a spherical cell or delocalized over the cytosol volume,
the ratio h was estimated to be
(0.02–0.05)×(Dmembr/Dcyt)×(rcell/rprot)
(Kholodenko et al., 2000b).
For instance, if the cell radius (rcell) is 10 μm and
the sum (rprot) of typical protein radii is approximately
10-2 μm, then
h≈(20–50)×Dmembr/Dcyt.
Because the diffusion coefficients of proteins in the membrane
(Dmembr) are about two orders of magnitude lower than in
the cytosol (Dcyt)
(Cherry, 1979;
Kusumi et al., 1993), we
conclude that the function of the membrane recruitment is unlikely to be an
enhancement of the encounter rates. In fact, it has already been suggested
that the fastest route to diffuse is through the cytosol, not through the
membrane, because of two orders of magnitude difference in the diffusion
coefficients (Bray, 1998).
Gain in the number of signaling complexes is more critical than an
increase in the association rate
In the reaction-limited extreme, the first-encounter rate is much faster
than the signal transduction rate, which can be estimated as the fraction of
molecules in the associated state (as if this step were at equilibrium)
multiplied by the reaction rate constant. An increase in the effective local
concentration of a cytosolic protein due to membrane relocation brings about
an increase in the apparent affinity to a binding partner in the membrane
(Ferrell, 1998). Haugh and
Lauffenburger (1997) estimated
that an increase in the reaction-limited protein association rate could be as
high as 102–103. In the general case, however, the
rate enhancement depends on many kinetic and molecular details including the
probability of success per diffusion encounter, the concentration depletion
zones near the targets, the reversibility of the binding and other factors
(Axelrod and Wang, 1994). The
binding reversibility suggests that the dissociation rate constant, in
addition to the association rate, may change upon the membrane relocation, as
in fact was observed for affinity enhancement due to `macromolecular crowding'
(Rohwer et al., 1998).
Therefore, although the speed of signal transduction increases by the
translocation of signaling proteins, it should not be assumed that any such
enhancement is due to an increase in the rate of formation of protein
complexes. Instead, the underlying mechanism can be an increase in the number
(or average lifetime) of signaling complexes, which act as catalysts
activating downstream processes. We submit that membrane localization serves
to enhance the extent of complex formation of signal-transduction proteins,
and hence increase the intensity of the signal that is being transduced.
Receptor-mediated membrane recruitment significantly increases the
number of complexes formed by a cytoplasmic protein and a membrane-anchored
protein
A gain in the number of signaling complexes involving a cytosolic protein X
and a membrane-bound protein Y can be brought about by a two-step, `piggyback'
mechanism. The first step is binding of X to an activated membrane receptor R
(`piggy'), which confines X to a membrane shell, where Y is located. The
second step is a `piggyback riding' of X until it meets membrane-anchored Y. X
then forms a complex with Y, while continuing to ride piggyback on R. The
quantitative analysis shows that this piggyback riding leads to a strong
reduction of the apparent dissociation constant, Kd,app,
which can be expressed through the equilibrium dissociation constant
Kd of cytosolic X and membrane-bound Y, as follows
(Kholodenko et al., 2000b):
1
Vc is the cytosol volume and Vm is the
volume of a water layer adjacent to the membrane where protein Y is confined.
The value of the dimensionless factor k is determined by the ratio of the
dissociation constant Kd,R for the binding of cytosolic X
to the activated receptor R and the concentration of the latter
(CR). Experimental data show that the
Kd,R values appear to be in the range of 1–100 nmol
l-1, whereas the total RTK concentrations (based on the whole
cytoplasmic volume) are in the range of 100–1000 nmol l-1 and
approximately 20–50% of the total amount is activated by the stimuli
(see Kholodenko et al., 1999;
Moehren et al., 2002, and the
references therein). Therefore, for membrane recruitment by RTKs, the value of
k does not exceed 1. We conclude that when the activated receptor is present
in excess of cytosolic X, the apparent dissociation constant decreases by a
factor as high as Vm/Vc, i.e. by 2 or
3 orders of magnitude in comparison with the actual Kd of
binding X and Y.
An alternative mechanism of an enhancement in signal transfer rate may be
an increase in the efficiency of the reaction between cytosolic X and
membrane-bound Y due to a conformational change of X upon binding to a
membrane anchor, such as receptor R. However, for SOS-Ras interactions,
structural and kinetic data on the catalytic mechanism and activity rule out
this possibility (Buday and Downward,
1993; Corbalan-Garcia et al.,
1998; Lenzen et al.,
1998).
Recruitment to a scaffold
Our results apply not only to the case of membrane translocation, but to
translocation into any subcellular compartment. Scaffolds of various sorts
should be mentioned here. They act as templates, bringing together signaling
proteins, organizing and coordinating the function of entire signaling
cascades (Bray, 1998).
Importantly, our results suggest that the number of signaling complexes will
increase only if these complexes do not dissociate from a scaffold. Even if
the interacting proteins were brought to close vicinity on a scaffold, the
dissociation of the protein complex from the scaffold will result in further
dissociation of the complex, which will be in thermodynamic equilibrium with
its components.
Specifics of SOS-Ras and GAP-Ras interactions
The calculations demonstrate that activation of Ras by SOS from the cytosol
is two or three orders of magnitude less effective than the catalysis,
mediated by the RTK-bound SOS, for instance, mediated by the EGFR-Grb2-SOS and
EGFR-Shc-Grb2-SOS complexes (Haugh and
Lauffenburger, 1997). After the dissociation of phosphorylated Src
homology and collagen domain protein (Shc) from EGFR, the Shc-Grb-SOS
complexes may bring about an additional route for Ras activation. However,
quasi-equilibrium association of these complexes with Ras may be insignificant
compared to receptor-mediated association unless the Shc-Grb-SOS complexes are
also targeted to the plasma membrane through specific domains, such as
phosphotyrosine-binding and pleckstrin homology domains
(Drugan et al., 2000;
Ugi et al., 2002).
Quantification of relative contributions of SOS complexes with the receptor
and with phosphorylated Shc to Ras activation is awaiting experimental
verification.
Signaling of activated Ras is turned off by the activation of
GTPase-activating proteins (GAPs) (Bollag
and McCormick, 1992). Similar to SOS, p120 GAP is a cytoplasmic
protein. As in the case of SOS signals, the membrane recruitment of GAP is
necessary for an appreciable increase in the rate of GTP hydrolysis on
Ras.
A variation on the topic of Ras activation by RTKs has been described for
fibroblast growth factor receptors (FGFRs)
(Kouhara et al., 1997).
Stimulation of FGFRs induces cell proliferation, differentiation and migration
by activation of the Ras/MAPK signaling pathway. However, unlike other RTKs,
activated FGFR cannot bind Grb2 directly. Recently, a novel membrane-anchored
protein, phosphorylated by activated FGFR, has been discovered and is known as
FRS2 (FGFR substrate 2) (Kouhara et al.,
1997). Tyrosine-phosphorylated FRS2 is able to bind Grb2 and,
therefore, the Grb2-SOS complex. In a familiar twist, the juxtaposition of the
FRS2-Grb2-Sos complex on the membrane may facilitate the Ras activation by
SOS, providing a feasible mechanism for linking FGFRs to the Grb2/SOS/Ras/MAPK
pathway.
Signal transduction through MAPK cascades can require endocytic
trafficking and/or active molecular transport
Our results demonstrate that the membrane recruitment of specific cytosolic
proteins can enhance receptor-induced activation of a membrane-anchored
target, such as Ras protein or a membrane-bound kinase, by a 1000-fold; but
how do signals emanating from membrane-bound proteins spread into the cell
interior and reach the nucleus? We will show here that simple diffusion of
activated proteins may be insufficient for the signal transfer through MAPK
pathways.
MAPK activation by RTKs and GPCRs
MAPK cascades are widely involved in eukaryotic signal transduction, and
these pathways are evolutionarily conserved in cells from yeast to mammals
(reviewed in Chang and Karin,
2001; Lewis et al.,
1998). Mammalian cells express at least four different MAPK
families, including the ERK, the c-Jun N-terminal kinase (JNK) and p38 MAPK
cascades. A three-tiered cascade comprises MAPK, MAPK kinase (MKK) and MKK
kinase (MKKK). A kinase at the first cascade level, MKKK, is activated at the
cell membrane. A kinase at the bottom level, MAPK, is activated in the cytosol
and phosphorylates target proteins both in the cytosol and nucleus. Mitogenic
signaling by RTKs and GPCRs is associated with Ras-dependent stimulation of
the ERK cascade. The classical paradigm for GPCR signaling involves the
agonist-dependent interaction of GPCRs with heterotrimeric G proteins that
stimulates the exchange of bound GDP for GTP, resulting in dissociation of G
proteins into α and βγ subunits. Subsequent interaction ofα
and βγ subunits with effector enzymes or ion channels
regulates the generation of soluble second messengers or ionic conductance
changes. However, this paradigm should be extended to include the
proliferative and differentiative effects of GPCRs. It is presently known that
in a multitude of cell types, GPCRs stimulate the MAPK cascades, such as the
ERK, JNK and p38 MAPK cascades (Garrington
and Johnson, 1999; Gutkind,
1998; van Biesen et al.,
1996). Interestingly, the pathways of GPCR and RTK-mediated MAPK
activation often converge. Recently, evidence has emerged that the overlap of
GPCR and RTK signaling pathways is accounted for in part by GPCR-mediated
`transactivation' of RTKs, such as EGFR
(Fig.1), platelet-derived
growth factor receptor and insulin-like growth factor-1 receptor. For
instance, EGFR activation by GPCR was shown to be mediated by signaling
pathways involving non-receptor tyrosine kinases, Src and Pyk2, and through
activation of the metalloprotease leading to the release of heparin binding
EGF (Pierce et al., 2001;
Prenzel et al., 1999). It is,
however, important to realize that although RTK transactivation is now a
recognized mechanism for GPCR-induced activation of the ERK cascade, other
signaling pathways can also contribute to this link
(Andreev et al., 2001;
Miller and Lefkowitz,
2001).
Heterogeneous spatial distribution of activated components in MAPK
cascades
Multiple cellular proteins are phosphorylated and dephosphorylated at
distinct cellular locations. In the Ras-ERK pathway, inactive Raf-1 (MKKK)
resides in the cytosol. Upon stimulation of cell-surface receptors, Raf-1 is
translocated from the cytosol to the plasma membrane by a high-affinity
binding to GTP-loaded Ras. At the membrane, Raf-1 undergoes a series of
activation steps involving interaction with 14-3-3 proteins and
phosphorylation on specific tyrosine and serine residues
(Mason et al., 1999). Although
the mechanism of activation is not yet completely understood, the association
of Raf-1 with membranes appears to be essential for its activation. The Raf-1
kinase phosphorylates the cytosolic kinase MEK (MKK) at the plasma membrane,
whereas soluble serine/threonine phosphatases dephosphorylate the activated
MEK in the bulk phase (Kyriakis et al.,
1992; Strack et al.,
1997). In the cytosol, active MEK kinase phosphorylates ERK (MAPK)
and specific ERK phosphatases are localized to the cytosol and nucleus.
Spatial separation of protein kinases and phosphatases suggests that there may
be cellular gradients of phosphorylated (active) forms of MEK and ERK, with
high concentrations in the periplasmic region near the phosphorylation
location and low concentrations at a distance from the plasma membrane. Since
activated ERK phosphorylates multiple cytoplasmic and nuclear targets, large
spatial gradients of this kinase may have very important implications for cell
signaling.
Indeed, the quantitative analysis demonstrates that the spatial separation
of kinases and phosphatases potentially results in precipitous gradients of
phospho-proteins, given measured values of protein diffusion coefficients and
phosphatase and kinase activities (Brown
and Kholodenko, 1999;
Kholodenko et al., 2000a). We
illustrate this analysis by calculating the spatial gradients for a single
level cascade, in which the kinase is located on the plasma membrane and the
phosphatase is distributed homogeneously in the cytoplasm (These calculations
and Fig. 2, which illustrates
them, are reproduced from Kholodenko,
2002, with permission from Elsevier Science.) For a spherical
cell, the following equation describes radial diffusion of the phospho-protein
(p) from the cell membrane, where p is produced at the rate,
vkin, to the cell interior, where p is
dephosphorylated at the rate, vp,
2
Here p is the phospho-protein concentration, D is the
protein diffusion coefficient in the cytosol (assumed to be equal for the
phosphorylated and unphosphorylated forms), x is a dimensionless
coordinate, equal to the distance from the cell center divided by the cell
radius, L [the distance d from the cell membrane is
expressed in terms of x as
d=(1–x)L], vkin and
vp are the rates of the kinase and phosphatase,
respectively. If we assume that the phosphatase is not saturated by a target
phospho-protein (a reasonable assumption for most cytosolic phosphatases), the
phosphatase rate vp can be described as
vp=Kp×p, where
Kp is the observed first-order rate constant (equal to the
ratio of the maximal activity to the Michaelis constant,
Kp=Vmax,p/Km,p).
Then, the steady-state solution to Equation 1 can be found readily
(Kholodenko et al., 2000a),
and the ratio of the phospho-protein concentrations at the distance x
from the cell center and at the plasma membrane,
p(x)/p(1), is given by:
3
Therefore, the phospho-protein concentration decreases toward the cell
interior approximately exponentially, as illustrated in
Fig.2. The phospho-protein
concentration profile, p(x)/p(1), depends on the
dimensionless parameter α (recognizable as the square root of the
Damkohler number), which compares the phosphatase and the diffusive time
scales. Therefore, all we need to know are the observed first-order rate
constant of the phosphatase reaction Kp, the diffusion
coefficient D and the cell radius. A typical value for a hepatocyte
radius is 10 μm, D is estimated to be of the order of
10-8 cm2 s-1
(Arrio-Dupont et al., 2000;
Dayel et al., 1999;
Gershon et al., 1985;
Jacobson and Wojcieszyn,
1984), and values for Kp were found to range
from roughly 0.1 to 10 s-1 (see
Haugh and Lauffenburger, 1998;
Kholodenko et al., 2000a;
Zhao and Zhang, 2001, and
references therein). With Kp values of 0.25 or 1
s-1, the distance from the plasma membrane, at which the
phosphorylation signal is attenuated by a factor of 10, is equal to 6.8 or 2.6μ
m, respectively. Importantly, the exponential character of the rapid
decrease in the phosphorylation signal with the distance from the cell
membrane, d=L(1–x), does not depend on the
specific activity and kinetics of the membrane kinase, provided that the
phosphatase is far from saturation.
- Download figure
- Open in new tab
- Download powerpoint
This analysis suggests that unless there is an additional mechanism
(besides slow protein diffusion) for signal propagation through the MAPK
cascade, the levels of activated MEK and ERK will drop precipitously in the
cell interior. At distances larger than several μm from the plasma
membrane, the phosphorylation signal should decrease to subthreshold levels,
provided that the cytosolic phosphatase activity is sufficiently high. Since
eukaryotic cell radii vary from 5 to 50 μm, we conclude that propagating a
message from the plasma membrane to the nucleus can require an additional
vehicle besides diffusion in the cytoplasm.
A novel role of endocytosis in activation of MAPK signaling
Upon ligand binding and activation, many GPCRs and RTKs internalize
via clathrin-coated pits. For instance, in hepatocytes over 50% of
phosphorylated EGFR is transferred to early endosomes during the first 10 min
after EGF stimulation (Di Guglielmo et
al., 1994). Internalization takes receptor–ligand complexes
and other signaling proteins from the plasma membrane and brings them inside
the cell. Molecules that were not recycled back to the cell membrane are
degraded in lysosomes. Although internalized GPCRs continuously recycle back
to the cell surface after dephosphorylation in endosomes, a significant
proportion of receptors are located internally
(Koenig and Edwardson, 1997).
Therefore, traditionally, clathrin-mediated endocytosis has been implicated in
downregulation of signaling by plasma membrane receptors. A novel role of
endocytosis in `turning on' activation of the ERK cascade by cell surface
receptors was first reported for the EGF receptor
(Vieira et al., 1996). A
conditional defect in endocytosis can be imposed by the regulated expression
of a mutant form of dynamin (Dyn1-K44A), a GTPase that is required for
clathrin-coated vesicle formation. In HeLa cells, this expression led to a
marked decrease in EGF-induced ERK activation, whereas Shc phosphorylation was
enhanced in endocytosis-defective cells. Subsequent studies have demonstrated
that both GPCR- and EGFR-mediated activation of ERK is sensitive to various
distinct inhibitors of clathrin-mediated endocytosis, including
monodansylcadaverine, depletion of intracellular K+ or cholesterol,
cytochalasin D and a mutant dynamin
(Ceresa et al., 1998;
Daaka et al., 1998;
Kranenburg et al., 1999;
Maudsley et al., 2000;
Rizzo et al., 2001;
Vieira et al., 1996).
Therefore, a possible mechanism of control over signal transduction may engage
receptor endocytosis (Clague and Urbe,
2001; Di Fiore and De Camilli,
2001; Haugh et al.,
1999). However, whereas experimental evidence points to an
essential role of receptor endocytosis in the activation of MAPK cascades, the
reason for the involvement of the endocytic machinery remains poorly
understood (Ceresa and Schmid,
2000; Kranenburg et al.,
1999; Pierce et al.,
2000). Interestingly, in some cellular systems endocytosis was not
required to activate ERK (for a review, see
Leof, 2000).
The relationship between receptor internalization and ERK activation allows
us to suggest that trafficking of signaling intermediates within endocytic
vesicles may be an efficient way of propagating the signal
(Kholodenko, 2002). Indeed, it
was reported that the engagement of the endocytic machinery is essential for
ERK activation by MEK, but not for activation of Ras
(Kranenburg et al., 1999).
Interestingly, the data on the subcellular distribution of activated MEK
demonstrated that biphosphorylated MEK is detectable only at the plasma
membrane and in intracellular vesicles, whereas the total MEK pool is
cytosolic (Kranenburg et al.,
1999). Endocytic trafficking of active MEK can help to avoid the
formation of steep spatial gradients of phosphorylated MEK and ERK, since this
mechanism overcomes the spatial separation of kinases and phosphatases within
the MAPK cascade. Therefore, the endocytosis of phosphorylated MEK (or a
protein complex containing activated MEK) rather than of activated receptors
appears to be critical for ERK activation.
Scaffolding, cytoplasmic streaming and active transport as mechanisms
facilitating signal propagation
Other mechanisms besides endocytosis can help to spread the phosphorylation
signals from the plasma membrane further into the cell by preventing the
formation of steep spatial gradients of phosphorylated ERK. For instance,
cytoplasmic streaming, i.e. solvent fluxes brought about by the movement of
cytoplasmic organelles along actin cables, can contribute to intracellular
transport of activated MAPK kinases
(Agutter et al., 1995). Recent
evidence indicates that the MAPK cascade components can bind to scaffolding
proteins, e.g. MP1 and JIP-1 in mammalian cells (for a review, see
Garrington and Johnson, 1999).
Dephosphorylation of the MAP kinases in scaffold complexes may be decreased or
even precluded because of sterical obstructions, as was suggested by Levchenko
et al. (2000).
Scaffolding may also help to deliver an entire signaling complex containing
the MAP kinases to endocytic vesicles. Novel mechanisms have been discovered
that link GPCRs to MAPK activation through use of β-arrestin as a
scaffold for the ERK and JNK cascades
(Miller and Lefkowitz, 2001;
Pierce et al., 2001). Besides
its role in GPCR desensitization, β-arrestin has been shown to promote
the targeting of the receptor to clathrin-coated pits. As β-arrestins can
also recruit and activate Src, it is likely that the entire ERK and JNK
cascades can be activated and recruited for clathrin-mediated
internalization.
Recent data suggest that molecular motors can be involved in transport of
signaling complexes. In fact, in nerve cells, a cargo for the molecular motor
kinesin has been identified as scaffolding proteins for the JNK pathway, known
as JIPs (Verhey and Rapoport,
2001). Endocytic vesicles and signaling complexes, loaded on
molecular motors, can be transported along microtubules to remote cellular
locations.
Long-range signaling
Retrograde transport of signaling complexes
One of the most interesting puzzles facing neurobiologists is concerned
with unraveling the molecular mechanisms used by neurons to transfer signals
over long distances. The survival and function of developing neurons depends
on growth factors, the neurotrophins, such as nerve growth factor (NGF) and
its receptor, TrkA. NGF, which is made by peripheral tissues, binds to TrkA
receptors on distal axons, located up to 1 m away from the neuronal soma. How
does the survival signal propagate over a remarkably long distance to the cell
body in a physiologically relevant span of time? Diffusion is ruled out as a
mechanism of the retrograde signaling, since it would be prohibitively slow.
Indeed, we have seen above that diffusion may be insufficient even for
spreading the phosphorylation signals within the cell body from the plasma
membrane to the nucleus.
Several models were suggested to explain the retrograde transduction of
neuronal survival signals (for a review, see
Ginty and Segal, 2002).
According to a widely accepted model, shortly after NGF binding to TrkA at
nerve terminals, the NGF-TrkA complexes are internalized into endosomes by
means of clathrin-mediated endocytosis. Signaling endosomes containing
activated TrkA with associated ligand are retrogradely transported to the cell
bodies. Data on blocking the NGF-TrkA transport by the dynein ATPase inhibitor
(EHNA) suggest that vesicular transport can be carried out by molecular motors
(Reynolds et al., 1998). In
fact, Trk neurotrophin receptors were found to bind directly to a dynein light
chain (Yano et al., 2001). The
retrograde axonal transport was also shown to be inhibited by the
phosphatidylinositol 3-kinase (PI3K) inhibitor wortmannin, by cytochalasin D,
a disruptor of actin filaments, and by the microtubule inhibitor colchicine
(Reynolds et al., 1998). Taken
together, these studies support a model where the retrograde transport of the
NGF-TrkA complexes along axon structural elements (e.g. microtubules) can
provide survival signals to neurons.
Ligand-independent waves of receptor activation
Cells have developed multiple mechanisms to overcome the problem of
long-distance signaling. It was recently discovered that survival signals
might not depend on retrograde transport of NGF
(MacInnis and Campenot, 2002).
Application of NGF, which was covalently cross-linked to beads to prevent
internalization, showed that survival signals were able to reach the cell
bodies unaccompanied by the ligand that initiated TrkA activation. One
possibility is that activated TrkA can be internalized and retrogradely
transported without NGF, implying that the active state of TrkA can be
maintained in the absence of the ligand. Another possibility is that
downstream TrkA targets, such as PI3K or Akt, can mediate the retrograde
signal. An intriguing possibility is that activation of TrkA receptors in
nerve terminals brings about an NGF-independent wave of TrkA activation, which
rapidly propagates through the neuronal axon
(Ginty and Segal, 2002).
Indeed, ligand-independent waves of receptor (EGFR) activation, emanated from
a point EGF source, were recently reported
(Verveer et al., 2000).
How can waves of receptor activation emerge? A possible scenario is that an
activated receptor, monomer or dimer, is capable of transphosphorylating and
activating an inactive receptor molecule upon diffusive encounter, potentially
resulting in lateral spread of receptor activation. In fact, spontaneous
activation and signaling by overexpressed EGFR has been reported recently
(Thomas et al., 2003). Let us
sketch the functional implications of this local amplification assumption for
the dynamics of the receptor system on the cell surface. For simplicity, a
one-dimensional diffusion is analyzed, and only monomer or dimer forms of the
receptor are taken into account (the model can be generalized to include
monomer-dimer association/dissociation).
Given that the total receptor concentration (or surface density) is
c and the concentration of activated receptors at a distance
x from a ligand source is a(x), the amount of newly
activated receptors per unit time and volume/area will be
k×a(x)×[c–a(x)],
where the constant k is proportional to the probability of
phosphorylation per diffusive encounter. Active receptors are dephosphorylated
by the phosphatases at the rate vp(a). Under
these assumptions, the spatial propagation of activated receptor (from a
source at x=0) is described by the following equation:
4
where D is the diffusion coefficient. When the phosphatase activity
is low, Equation 4 reduces to the Fisher-like equation, the first partial
differential equation known to generate traveling waves
(Murray, 1993). A traveling
wave propagates without change of shape, i.e. the shape and speed of
propagation of the front is constant over time. From Equation 4, it follows
that an increase in total receptor amount (c) and a decrease in the
phosphatase activity (vp) may trigger global RTK
activation. The situation in which a cell can be `switched on' by a localized
stimulus is in contrast to the current paradigm of `restricted' signaling by
RTKs activated by local growth factor stimulation reported recently for EGF
signaling in COS cells (Sawano et al.,
2002). Remarkably, overexpression of EGF receptors in these same
cells resulted in global EGFR activation over the entire cell surface
following a localized stimulation with EGF
(Sawano et al., 2002). We
conclude that an initial localized stimulation can be responsible for
triggering global RTK activation in cells. Such a mechanism may have
implications for the pathogenesis of cancer, where RTKs are overexpressed due
to mutations (Thomas et al.,
2003). Studies of the dependence of traveling wave solutions on
molecular mechanisms of RTK activation and kinetic parameters may help
facilitate our understanding of crucial control points in tumor
development.
Traveling waves of activated protein kinases
Another possible scenario of how a survival signal can spread over the axon
includes traveling waves that may arise in a bistable protein kinase cascade.
Bistability is a common theme in cell signaling cascades that contain a
positive feedback loop, or a double-negative feedback loop
(Bhalla and Iyengar, 1999;
Ferrell and Machleder, 1998;
Gardner et al., 2000; Thron,
1997,
1999;
Tyson and Novak, 2001). A
bistable protein kinase cascade can switch between two different stable
states, one corresponding to a low and another to a high activity, but cannot
exist in an intermediate unstable state. When the signal strength is below a
given threshold, a kinase cascade, such as the Mos-MEK-p42 MAPK cascade and
the JNK cascade in Xenopus oocytes
(Bagowski and Ferrell, 2001;
Ferrell, 1999), and the ERK
cascade in mouse NIH-3T3 cells (Bhalla et
al., 2002), remains in a low activity state. An increase in the
signal above the threshold switches the cascade to a high activity state.
Importantly, the cascade will remain in this high activity state even when the
initial signal decreases to subthreshold values.
It was proposed that cells utilize bistable signaling circuits that would
enable them to be capable of `all-or-none' switching and to `remember' a
transient differentiation stimulus long after the stimulus was removed
(Ferrell, 2002). Here we
suggest that yet another role for bistable protein kinase cascades may be to
support traveling waves of activated phospho-proteins, propagating signals to
remote cellular locations. Bistability produces local amplification of the
spreading signal, and the combination of local amplification and diffusion may
generate traveling waves. In fact, traveling waves in bistable systems have
been extensively studied in physics, chemistry and biology
(Castiglione et al., 2002;
Christoph et al., 1999;
Keener and Sneyd, 1998;
Shvartsman et al., 2002;
Zhabotinsky and Zaikin, 1973).
MAPK cascades and the PI3K/PDK1/Akt cascade coupled with the PIP3 synthesis
pathway might be candidates for bistable protein kinase cascades that are
capable of propagating traveling waves.
Importantly, all of the models of long-range signaling considered are not
mutually exclusive. We hypothesize that multiple mechanisms of information
transfer have evolved in neurons to transmit signals over long distances.
Future experimental work will show if waves of activated kinases emerge in
cells.
Concluding remarks
Spatio-temporal organization of mitogenic pathways analyzed here is central
for understanding the control over intracellular signal transfer. A picture is
emerging, in which simple diffusion has a limited role in intracellular
transport of signaling complexes. Endocytosis, scaffolding, molecular motors
and traveling waves of phospho-proteins appear to be involved in the
propagation of signals to different cellular locations. These mechanisms
control cellular decisions that determine cell fate.
Abbreviations
EGF, epidermal growth factor
EGFR, EGF receptor
ERK, extracellular signal regulated kinase
GAP, GTPase-activating protein
GDP/GTP, guanosine di/triphosphate
GPCR, G-protein coupled receptor
Grb2, growth factor receptor binding protein 2
JNK, c-Jun N-terminal kinase
MAPK, mitogen activated protein kinase
MKK, MAPK kinase
MKKK, MKK kinase
NGF, nerve growth factor
PI3K, phosphatidylinositol 3-kinase
RTK, receptor tyrosine kinase
SH, Src homology domain
Shc, Src homology and collagen domain protein
SOS, Son of Sevenless homolog protein
TrkA, the NGF receptor
ACKNOWLEDGEMENTS
I thank Drs A. Davies, J. Pastorino and S. Shvartsman for critical reading
of the manuscript. This work was supported by Grants GM59570 and AA08714 from
the National Institute of Health.
- © The Company of Biologists Limited
2003
References
↵
Adam, G. and Delbruck, M. (1968). Reduction of
dimensionality in biological diffusion processes. In Structural
Chemistry and Molecular Biology (ed. A. Rich and N. Davidson),
pp. 198-215.San Francisco: W. H. Freeman and
Co.
↵
Agutter, P. S., Malone, P. C. and Wheatley, D. N.
(1995). Intracellular transport mechanisms: a critique of
diffusion theory. J. Theor. Biol.
176,261
-272.
OpenUrlCrossRefPubMed
↵
Andreev, J., Galisteo, M. L., Kranenburg, O., Logan, S. K.,
Chiu, E. S., Okigaki, M., Cary, L. A., Moolenaar, W. H. and
Schlessinger, J. (2001). Src and Pyk2 mediate
G-protein-coupled receptor activation of epidermal growth factor receptor
(EGFR) but are not required for coupling to the mitogen-activated protein
(MAP) kinase signaling cascade. J. Biol. Chem.
276,20130
-20135.
OpenUrlAbstract/FREE Full Text
↵
Arrio-Dupont, M., Foucault, G., Vacher, M., Devaux, P. F. and
Cribier, S. (2000). Translational diffusion of
globular proteins in the cytoplasm of cultured muscle cells.
Biophys. J. 78,901
-907.
OpenUrlCrossRefPubMedWeb of Science
↵
Axelrod, D. and Wang, M. D. (1994).
Reduction-of-dimensionality kinetics at reaction-limited cell surface
receptors. Biophys. J.
66,588
-600.
OpenUrlCrossRefPubMedWeb of Science
↵
Bagowski, C. P. and Ferrell, J. E., Jr (2001).
Bistability in the JNK cascade. Curr. Biol.
11,1176
-1182.
OpenUrlCrossRefPubMedWeb of Science
↵
Berg, H. C. and Purcell, E. M. (1977). Physics
of chemoreception. Biophys. J.
20,193
-219.
OpenUrlCrossRefPubMedWeb of Science
↵
Bhalla, U. S. and Iyengar, R. (1999). Emergent
properties of networks of biological signaling pathways [see comments].
Science 283,381
-387.
OpenUrlAbstract/FREE Full Text
↵
Bhalla, U. S., Ram, P. T. and Iyengar, R.
(2002). MAP kinase phosphatase as a locus of flexibility in a
mitogen-activated protein kinase signaling network.
Science 297,1018
-10123.
OpenUrlAbstract/FREE Full Text
↵
Bollag, G. and McCormick, F. (1992). GTPase
activating proteins. Sem. Cancer Biol.
3, 199-208.
↵
Bray, D. (1998). Signaling complexes:
biophysical constraints on intracellular communication. Annu. Rev.
Biophys. Biomol. Struct. 27,59
-75.
OpenUrlCrossRefPubMedWeb of Science
↵
Brown, G. C. and Kholodenko, B. N. (1999).
Spatial gradients of cellular phospho-proteins. FEBS
Lett. 457,452
-454.
OpenUrlCrossRefPubMedWeb of Science
↵
Buday, L. and Downward, J. (1993). Epidermal
growth factor regulates p21ras through the formation of a complex of receptor,
Grb2 adapter protein, and Sos nucleotide exchange factor.
Cell 73,611
-620.
OpenUrlCrossRefPubMedWeb of Science
↵
Castiglione, F., Bernaschi, M., Succi, S., Heinrich, R. and
Kirschner, M. W. (2002). Intracellular signal
propagation in a two-dimensional autocatalytic reaction model.
Phys. Rev. E Stat. Nonlin. Soft Matter Phys.
66,1
-031905-10.OpenUrl
↵
Ceresa, B. P., Kao, A. W., Santeler, S. R. and Pessin, J. E.
(1998). Inhibition of clathrin-mediated endocytosis selectively
attenuates specific insulin receptor signal transduction pathways.
Mol. Cell. Biol. 18,3862
-3870.
OpenUrlAbstract/FREE Full Text
↵
Ceresa, B. P. and Schmid, S. L. (2000).
Regulation of signal transduction by endocytosis. Curr. Opin. Cell
Biol. 12,204
-210.
OpenUrlCrossRefPubMedWeb of Science
↵
Chang, L. and Karin, M. (2001). Mammalian MAP
kinase signalling cascades. Nature
410, 37-40.
OpenUrlCrossRefPubMedWeb of Science
↵
Cherry, R. J. (1979). Rotational and lateral
diffusion of membrane proteins. Biochim. Biophys. Acta
559,289
-327.
OpenUrlPubMed
↵
Christoph, J., Strasser, P., Eiswirth, M. and Ertl, G.
(1999). Remote triggering of waves in an electrochemical system.
Science 284,291
-293.
OpenUrlAbstract/FREE Full Text
↵
Clague, M. J. and Urbe, S. (2001). The
interface of receptor trafficking and signalling. J. Cell
Sci. 114,3075
-3081.
OpenUrlAbstract/FREE Full Text
↵
Corbalan-Garcia, S., Margarit, S. M., Galron, D., Yang, S. S.
and Bar- Sagi, D. (1998). Regulation of Sos activity
by intramolecular interactions. Mol. Cell Biol.
18,880
-886.
OpenUrlAbstract/FREE Full Text
↵
Daaka, Y., Luttrell, L. M., Ahn, S., Della Rocca, G. J.,
Ferguson, S. S., Caron, M. G. and Lefkowitz, R. J.
(1998). Essential role for G protein-coupled receptor endocytosis
in the activation of mitogen-activated protein kinase. J. Biol.
Chem. 273,685
-688.
OpenUrlAbstract/FREE Full Text
↵
Dayel, M. J., Hom, E. F. and Verkman, A. S.
(1999). Diffusion of green fluorescent protein in the
aqueous-phase lumen of endoplasmic reticulum. Biophys.
J. 76,2843
-2851.
OpenUrlCrossRefPubMedWeb of Science
↵
Di Fiore, P. P. and De Camilli, P. (2001).
Endocytosis and signaling. an inseparable partnership.
Cell 106,1
-4.
OpenUrlCrossRefPubMedWeb of Science
↵
Di Guglielmo, G. M., Baass, P. C., Ou, W. J., Posner, B. I. and
Bergeron, J. J. (1994). Compartmentalization of SHC,
GRB2 and mSOS, and hyperphosphorylation of Raf-1 by EGF but not insulin in
liver parenchyma. EMBO J.
13,4269
-4277.
OpenUrlPubMedWeb of Science
↵
Drugan, J. K., Rogers-Graham, K., Gilmer, T., Campbell, S. and
Clark, G. J. (2000). The Ras/p120 GTPase-activating
protein (GAP) interaction is regulated by the p120 GAP pleckstrin homology
domain. J. Biol. Chem.
275,35021
-35027.
OpenUrlAbstract/FREE Full Text
↵
Egan, S. E., Giddings, B. W., Brooks, M. W., Buday, L.,
Sizeland, A. M. and Weinberg, R. A. (1993). Association of
Sos Ras exchange protein with Grb2 is implicated in tyrosine kinase signal
transduction and transformation. Nature
363, 45-51.
OpenUrlCrossRefPubMedWeb of Science
↵
Ferrell, J. E., Jr (1998). How regulated
protein translocation can produce switch-like responses. Trends
Biochem. Sci. 23,461
-465.
OpenUrlCrossRefPubMedWeb of Science
↵
Ferrell, J. E., Jr (1999). Building a cellular
switch: more lessons from a good egg. BioEssays
21,866
-870.
OpenUrlCrossRefPubMedWeb of Science
↵
Ferrell, J. E., Jr (2002). Self-perpetuating
states in signal transduction: positive feedback, double-negative feedback and
bistability. Curr. Opin. Cell Biol.
14,140
-148.
OpenUrlCrossRefPubMedWeb of Science
↵
Ferrell, J. E., Jr and Machleder, E. M. (1998).
The biochemical basis of an all-or-none cell fate switch in Xenopus oocytes.
Science 280,895
-898.
OpenUrlAbstract/FREE Full Text
↵
Gardner, T. S., Cantor, C. R. and Collins, J. J.
(2000). Construction of a genetic toggle switch in
Escherichia coli. Nature
403, 339-342.
OpenUrlCrossRefPubMedWeb of Science
↵
Garrington, T. P. and Johnson, G. L. (1999).
Organization and regulation of mitogen-activated protein kinase signaling
pathways. Curr. Opin. Cell Biol.
11,211
-218.
OpenUrlCrossRefPubMedWeb of Science
↵
Gershon, N. D., Porter, K. R. and Trus, B. L.
(1985). The cytoplasmic matrix: its volume and surface area and
the diffusion of molecules through it. Proc. Natl. Acad. Sci.
USA 82,5030
-5034.
OpenUrlAbstract/FREE Full Text
↵
Ginty, D. D. and Segal, R. A. (2002).
Retrograde neurotrophin signaling: Trking along the axon. Curr.
Opin. Neurobiol. 12,268
-274.
OpenUrlCrossRefPubMedWeb of Science
↵
Gutkind, J. S. (1998). The pathways connecting
G protein-coupled receptors to the nucleus through divergent mitogen-activated
protein kinase cascades. J. Biol. Chem.
273,1839
-1842.
OpenUrlFREE Full Text
↵
Haugh, J. M., Huang, A. C., Wiley, H. S., Wells, A. and
Lauffenburger, D. A. (1999). Internalized epidermal
growth factor receptors participate in the activation of p21(ras) in
fibroblasts. J. Biol. Chem.
274,34350
-34360.
OpenUrlAbstract/FREE Full Text
↵
Haugh, J. M. and Lauffenburger, D. A. (1997).
Physical modulation of intracellular signaling processes by locational
regulation. Biophys. J.
72,2014
-20131.
OpenUrlCrossRefPubMed
↵
Haugh, J. M. and Lauffenburger, D. A. (1998).
Analysis of receptor internalization as a mechanism for modulating signal
transduction. J. Theor. Biol.
195,187
-218.
OpenUrlCrossRefPubMed
↵
Hochachka, P. W. (1999). The metabolic
implications of intracellular circulation. Proc. Natl. Acad. Sci.
USA 96,12233
-12239.
OpenUrlAbstract/FREE Full Text
↵
Hollenbeck, P. (2001). Cytoskeleton:
Microtubules get the signal. Curr. Biol.
11,R820
-R823.
OpenUrlCrossRefPubMedWeb of Science
↵
Jacobson, K. and Wojcieszyn, J. (1984). The
translational mobility of substances within the cytoplasmic matrix.
Proc. Natl. Acad. Sci. USA
81,6747
-6751.
OpenUrlAbstract/FREE Full Text
↵
Keener, J. and Sneyd, J. (1998).
Mathematical Physiology. New York, Berlin, Heidelberg:
Springer.
↵
Kholodenko, B. N. (2002). MAP kinase cascade
signaling and endocytic trafficking: a marriage of convenience?
Trends Cell Biol. 12,173
-177.
OpenUrlCrossRefPubMedWeb of Science
↵
Kholodenko, B. N., Brown, G. C. and Hoek, J. B.
(2000a). Diffusion control of protein phosphorylation in signal
transduction pathways. Biochem. J.
350,901
-907.
↵
Kholodenko, B. N., Demin, O. V., Moehren, G. and Hoek, J. B.
(1999). Quantification of short term signaling by the epidermal
growth factor receptor. J. Biol. Chem.
274,30169
-30181.
OpenUrlAbstract/FREE Full Text
↵
Kholodenko, B. N., Hoek, J. B. and Westerhoff, H. V.
(2000b). Why cytoplasmic signalling proteins should be recruited
to cell membranes. Trends Cell Biol.
10,173
-178.
OpenUrlCrossRefPubMedWeb of Science
↵
Koenig, J. A. and Edwardson, J. M. (1997).
Endocytosis and recycling of G protein-coupled receptors. Trends
Pharmacol. Sci. 18,276
-287.
OpenUrlCrossRefPubMed
↵
Kouhara, H., Hadari, Y. R., Spivak-Kroizman, T., Schilling, J.,
Bar-Sagi, D., Lax, I. and Schlessinger, J. (1997). A
lipid-anchored Grb2-binding protein that links FGF-receptor activation to the
Ras/MAPK signaling pathway. Cell
89,693
-702.
OpenUrlCrossRefPubMedWeb of Science
↵
Kranenburg, O., Verlaan, I. and Moolenaar, W. H.
(1999). Dynamin is required for the activation of
mitogen-activated protein (MAP) kinase by MAP kinase kinase. J.
Biol. Chem. 274,35301
-35304.
OpenUrlAbstract/FREE Full Text
↵
Kusumi, A., Sako, Y. and Yamamoto, M. (1993).
Confined lateral diffusion of membrane receptors as studied by single particle
tracking (nanovid microscopy). Effects of calcium-induced differentiation in
cultured epithelial cells. Biophys. J.
65,2021
-2040.
OpenUrlCrossRefPubMedWeb of Science
↵
Kyriakis, J. M., App, H., Zhang, X. F., Banerjee, P., Brautigan,
D. L., Rapp, U. R. and Avruch, J. (1992). Raf-1
activates MAP kinase-kinase. Nature
358,417
-421.
OpenUrlCrossRefPubMedWeb of Science
↵
Lenzen, C., Cool, R. H., Prinz, H., Kuhlmann, J. and
Wittinghofer, A. (1998). Kinetic analysis by fluorescence of
the interaction between Ras and the catalytic domain of the guanine nucleotide
exchange factor Cdc25Mm. Biochemistry
37,7420
-7430.
OpenUrlCrossRefPubMed
↵
Leof, E. B. (2000). Growth factor receptor
signalling: location, location, location. Trends Cell
Biol. 10,343
-348.
OpenUrlCrossRefPubMedWeb of Science
↵
Levchenko, A., Bruck, J. and Sternberg, P. W.
(2000). Scaffold proteins may biphasically affect the levels of
mitogen-activated protein kinase signaling and reduce its threshold
properties. Proc. Natl. Acad. Sci. USA
97,5818
-5823.
OpenUrlAbstract/FREE Full Text
↵
Lewis, T. S., Shapiro, P. S. and Ahn, N. G.
(1998). Signal transduction through MAP kinase cascades.
Adv. Cancer Res. 74,49
-139.
OpenUrlCrossRefPubMedWeb of Science
↵
MacInnis, B. L. and Campenot, R. B. (2002).
Retrograde support of neuronal survival without retrograde transport of nerve
growth factor. Science
295,1536
-1539.
OpenUrlAbstract/FREE Full Text
↵
Mason, C. S., Springer, C. J., Cooper, R. G., Superti-Furga,
G., Marshall, C. J. and Marais, R. (1999). Serine and
tyrosine phosphorylations cooperate in Raf-1, but not B-Raf activation.
EMBO J. 18,2137
-2148.
OpenUrlAbstract
↵
Maudsley, S., Pierce, K. L., Zamah, A. M., Miller, W. E., Ahn,
S., Daaka, Y., Lefkowitz, R. J. and Luttrell, L. M.
(2000). The beta(2)-adrenergic receptor mediates extracellular
signal-regulated kinase activation via assembly of a multi-receptor complex
with the epidermal growth factor receptor. J. Biol.
Chem. 275,9572
-9580.
OpenUrlAbstract/FREE Full Text
↵
Miller, W. E. and Lefkowitz, R. J. (2001).
Expanding roles for beta-arrestins as scaffolds and adapters in GPCR signaling
and trafficking. Curr. Opin. Cell Biol.
13,139
-145.
OpenUrlCrossRefPubMedWeb of Science
↵
Mochly-Rosen, D. (1995). Localization of
protein kinases by anchoring proteins: a theme in signal transduction.
Science 268,247
-251.
OpenUrlAbstract/FREE Full Text
↵
Moehren, G., Markevich, N., Demin, O., Kiyatkin, A., Goryanin,
I., Hoek, J. B. and Kholodenko, B. N. (2002).
Temperature Dependence of the epidermal growth factor receptor signaling
network can be accounted for by a kinetic model.
Biochemistry 41,306
-320.
OpenUrlCrossRefPubMed
↵
Murray, J. D. (1993). Mathematical
Biology. Berlin, Heidelberg: Springer.
↵
Pierce, K. L., Luttrell, L. M. and Lefkowitz, R. J.
(2001). New mechanisms in heptahelical receptor signaling to
mitogen activated protein kinase cascades. Oncogene
20,1532
-1539.
OpenUrlCrossRefPubMedWeb of Science
↵
Pierce, K. L., Maudsley, S., Daaka, Y., Luttrell, L. M. and
Lefkowitz, R. J. (2000). Role of endocytosis in the
activation of the extracellular signal-regulated kinase cascade by
sequestering and nonsequestering G protein-coupled receptors. Proc.
Natl. Acad. Sci. USA 97,1489
-1494.
OpenUrlAbstract/FREE Full Text
↵
Prenzel, N., Zwick, E., Daub, H., Leserer, M., Abraham, R.,
Wallasch, C. and Ullrich, A. (1999). EGF receptor
transactivation by G-protein-coupled receptors requires metalloproteinase
cleavage of proHB-EGF. Nature
402,884
-888.
OpenUrlPubMedWeb of Science
↵
Reynolds, A. J., Bartlett, S. E. and Hendry, I. A.
(1998). Signalling events regulating the retrograde axonal
transport of 125I-beta nerve growth factor in vivo.
Brain Res. 798,67
-74.
OpenUrlCrossRefPubMedWeb of Science
↵
Rizzo, M. A., Kraft, C. A., Watkins, S. C., Levitan, E. S. and
Romero, G. (2001). Agonist-dependent traffic of
raft-associated Ras and Raf-1 is required for activation of the
mitogen-activated protein kinase cascade. J. Biol.
Chem. 276,34928
-34933.
OpenUrlAbstract/FREE Full Text
↵
Rohwer, J. M., Postma, P. W., Kholodenko, B. N. and Westerhoff,
H. V. (1998). Implications of macromolecular crowding for
signal transduction and metabolite channeling. Proc. Natl. Acad.
Sci. USA 95,10547
-10552.
OpenUrlAbstract/FREE Full Text
↵
Sastry, L., Lin, W., Wong, W. T., Di Fiore, P. P., Scoppa, C. A.
and King, C. R. (1995). Quantitative analysis of
Grb2-Sos1 interaction: the N-terminal SH3 domain of Grb2 mediates affinity.
Oncogene 11,1107
-1112.
OpenUrlPubMedWeb of Science
↵
Sawano, A., Takayama, S., Matsuda, M. and Miyawaki, A.
(2002). Lateral propagation of EGF signaling after local
stimulation is dependent on receptor density. Dev.
Cell 3,245
-257.
OpenUrlCrossRefPubMedWeb of Science
↵
Schlessinger, J. (2000). Cell signaling by
receptor tyrosine kinases. Cell
103,211
-225.
OpenUrlCrossRefPubMedWeb of Science
↵
Schlessinger, J. and Bar-Sagi, D. (1994).
Activation of Ras and other signaling pathways by receptor tyrosine kinases.
Cold Spring Harbor Symp. Quant. Biol.
59,173
-179.
OpenUrlAbstract/FREE Full Text
↵
Shvartsman, S. Y., Hagan, M. P., Yacoub, A., Dent, P., Wiley, H.
S. and Lauffenburger, D. A. (2002). Autocrine loops
with positive feedback enable context-dependent cell signaling. Am.
J. Physiol. Cell Physiol. 282,C545
-C559.
OpenUrlAbstract/FREE Full Text
↵
Strack, S., Westphal, R. S., Colbran, R. J., Ebner, F. F. and
Wadzinski, B. E. (1997). Protein serine/threonine
phosphatase 1 and 2A associate with and dephosphorylate neurofilaments.
Brain Res. Mol. Brain Res.
49, 15-28.
OpenUrlCrossRefPubMed
↵
Thomas, C. Y., Chouinard, M., Cox, M., Parsons, S.,
Stallings-Mann, M., Garcia, R., Jove, R. and Wharen, R.
(2003). Spontaneous activation and signaling by overexpressed
epidermal growth factor receptors in glioblastoma cells. Int. J.
Cancer 104,19
-27.
OpenUrlCrossRefPubMedWeb of Science
↵
Thron, C. D. (1997). Bistable biochemical
switching and the control of the events of the cell cycle.
Oncogene 15,317
-325.
OpenUrlCrossRefPubMedWeb of Science
↵
Thron, C. D. (1999). Mathematical analysis of
binary activation of a cell cycle kinase which down-regulates its own
inhibitor. Biophys. Chem.
79, 95-106.
OpenUrlCrossRefPubMedWeb of Science
↵
Tyson, J. J. and Novak, B. (2001). Regulation
of the eukaryotic cell cycle: molecular antagonism, hysteresis, and
irreversible transitions. J. Theor. Biol.
210,249
-263.
OpenUrlCrossRefPubMedWeb of Science
↵
Ugi, S., Sharma, P. M., Ricketts, W., Imamura, T. and Olefsky,
J. M. (2002). Phosphatidylinositol 3-kinase is required for
insulin-stimulated tyrosine phosphorylation of Shc in 3T3-L1 adipocytes.
J. Biol. Chem. 277,18592
-18597.
OpenUrlAbstract/FREE Full Text
↵
van Biesen, T., Luttrell, L. M., Hawes, B. E. and Lefkowitz, R.
J. (1996). Mitogenic signaling via G protein-coupled
receptors. Endocrinol. Rev.
17,698
-714.
OpenUrlCrossRefPubMedWeb of Science
↵
Verhey, K. J. and Rapoport, T. A. (2001).
Kinesin carries the signal. Trends Biochem. Sci.
26,545
-550.
OpenUrlCrossRefPubMed
↵
Verveer, P. J., Wouters, F. S., Reynolds, A. R. and Bastiaens,
P. I. (2000). Quantitative imaging of lateral ErbB1 receptor
signal propagation in the plasma membrane. Science
290,1567
-1570.
OpenUrlAbstract/FREE Full Text
↵
Vieira, A. V., Lamaze, C. and Schmid, S. L.
(1996). Control of EGF receptor signaling by clathrin-mediated
endocytosis. Science
274,2086
-2089.
OpenUrlAbstract/FREE Full Text
↵
Yano, H., Lee, F. S., Kong, H., Chuang, J., Arevalo, J., Perez,
P., Sung, C. and Chao, M. V. (2001). Association of
Trk neurotrophin receptors with components of the cytoplasmic dynein motor.
J. Neurosci. 21,RC125
.
↵
Zhabotinsky, A. M. and Zaikin, A. N. (1973).
Autowave processes in a distributed chemical system. J. Theor.
Biol. 40,45
-61.
OpenUrlCrossRefPubMedWeb of Science
↵
Zhao, Y. and Zhang, Z. Y. (2001). The mechanism
of dephosphorylation of extracellular signal-regulated kinase 2 by
mitogen-activated protein kinase phosphatase 3. J. Biol.
Chem. 276,32382
-32391.
OpenUrlAbstract/FREE Full Text
View Abstract
Previous ArticleNext Article
Back to top
Previous ArticleNext Article
This Issue
Download PDF
Thank you for your interest in spreading the word on Journal of Experimental Biology.
NOTE: We only request your email address so that the person you are recommending the page to knows that you wanted them to see it, and that it is not junk mail. We do not capture any email address.
Four-dimensional organization of protein kinase signaling cascades: the roles of diffusion, endocytosis and molecular motors
(Your Name) has sent you a message from Journal of Experimental Biology
(Your Name) thought you would like to see the Journal of Experimental Biology web site.
Share
Journal of Experimental Biology 2003 206: 2073-2082; doi: 10.1242/jeb.00298
Citation Tools
Journal of Experimental Biology 2003 206: 2073-2082; doi: 10.1242/jeb.00298
Citation Manager Formats
- BibTeX
- Bookends
- EasyBib
- EndNote (tagged)
- EndNote 8 (xml)
- Medlars
- Mendeley
- Papers
- RefWorks Tagged
- Ref Manager
- RIS
- Zotero
Alerts
Please log in to add an alert for this article.
Article navigation
- Top
Article- SUMMARY
- Introduction
- Regulation of signaling by membrane recruitment of cytosolic
proteins - Signal transduction through MAPK cascades can require endocytic
trafficking and/or active molecular transport - Long-range signaling
- Concluding remarks
- Abbreviations
- ACKNOWLEDGEMENTS
- References
- Figures & tables
- Info & metrics
PDF
Related articles
Cited by...
More in this TOC section
Modular structure of sodium-coupled bicarbonate transporters
NhaA crystal structure: functional–structural insights
Aquaporins: translating bench research to human disease
Similar articles
Other journals from The Company of Biologists
Development
Journal of Cell Science
Disease Models & Mechanisms
Biology Open
var googletag = googletag || ;
googletag.cmd = googletag.cmd || [];
googletag.cmd.push(function()
googletag.defineSlot('/155913954/JEB_2', [336, 280], 'div-gpt-ad-1505473406834-0').addService(googletag.pubads());
googletag.pubads().enableSingleRequest();
googletag.enableServices();
);
googletag.cmd.push(function()
googletag.display("div-gpt-ad-1505473406834-0");
);
Editors’ choice - Mitochondrial plasticity in the cerebellum of two anoxia-tolerant sharks: contrasting responses to anoxia/re-oxygenation
A grey carpet shark (Chiloscyllium punctatum). Photo credit: Jules Devaux
Epaulette sharks and grey carpet sharks frequently experience oxygen deprivation, the equivalent of humans suffering a stroke or heart attack, yet suffer no brain damage. Jules Devaux, Anthony Hickey and Gillian Renshaw show how mitochondria in the brains of these fish have adapted to protect them from the ill effects of anoxia.
Featured article – Newt chewing is all in the tongue
An Italian crested newt (Triturus carnifex). Photo credit: Daniel Schwarz
Most amphibians swallow food unchewed, but not Italian crested newts. They chew by rasping food against teeth in the roof of the mouth with the tongue. Read the research article from Egon Heiss, Daniel Schwarz and Nicolai Konow.
Commentary - Loss-of-function approaches in comparative physiology: is there a future for knockdown experiments in the era of genome editing?
Alex Zimmer et al. discuss the merits and pitfalls of current knockdown and knockout approaches and advocate for the continued use of knockdown experiments in order to maintain progress in physiological research.
News - So long Raul and welcome Trish
We are sad to announce that after 14 years as an Editor of Journal of Experimental Biology Raul Suarez is stepping down, but we are happy to welcome Trish Schulte to the team of Editors.
Trish says that she is looking forward to seeing “the super-cool work that is being submitted to JEB. I love the diversity of science that appears in the journal.”
Scientific Meeting Grants
The Company of Biologists provide grants to fund scientific meetings, workshops and conferences in the fields covered by our journals. Typically, meetings with fewer than 100 people may be granted up to £2,000, increasing up to £6,000 for about 400 people. The next deadline to apply is 27 May 2019.