ArticlePDF Available

Dynamic Signaling in the Hog1 MAPK Pathway Relies on High Basal Signal Transduction

Authors:

Abstract and Figures

Appropriate regulation of the Hog1 mitogen-activated protein kinase (MAPK) pathway is essential for cells to survive osmotic stress. Here, we show that the two sensing mechanisms upstream of Hog1 display different signaling properties. The Sho1 branch is an inducible nonbasal system, whereas the Sln1 branch shows high basal signaling that is restricted by a MAPK-mediated feedback mechanism. A two-dimensional mathematical model of the Snl1 branch, including high basal signaling and a Hog1-regulated negative feedback, shows that a system with basal signaling exhibits higher efficiency, with faster response times and higher sensitivity to variations in external signals, than would systems without basal signaling. Analysis of two other yeast MAPK pathways, the Fus3 and Kss1 signaling pathways, indicates that high intrinsic basal signaling may be a general property of MAPK pathways allowing rapid and sensitive responses to environmental changes.
Content may be subject to copyright.
(63), ra13. [DOI: 10.1126/scisignal.2000056] 2Science Signaling
Posas (24 March 2009)
Javier Macia, Sergi Regot, Tom Peeters, Núria Conde, Ricard Solé and Francesc
Transduction
Dynamic Signaling in the Hog1 MAPK Pathway Relies on High Basal Signal`
This information is current as of 17 May 2013.
The following resources related to this article are available online at http://stke.sciencemag.org.
Article Tools http://stke.sciencemag.org/cgi/content/full/sigtrans;2/63/ra13
Visit the online version of this article to access the personalization and article tools:
Materials
Supplemental http://stke.sciencemag.org/cgi/content/full/sigtrans;2/63/ra13/DC1
"Supplementary Materials"
Related Content
http://stke.sciencemag.org/cgi/content/full/sci;324/5924/151c http://stke.sciencemag.org/cgi/content/abstract/sigtrans;2/68/ec151
http://stke.sciencemag.org/cgi/content/abstract/sigtrans;2/77/pc12 http://stke.sciencemag.org/cgi/content/abstract/sigtrans;2/81/eg10
http://stke.sciencemag.org/cgi/content/abstract/sigtrans;2/81/ra40 http://stke.sciencemag.org/cgi/content/abstract/sigtrans;2/82/pe48 http://stke.sciencemag.org/cgi/content/abstract/sigtrans;2/96/eg14 http://stke.sciencemag.org/cgi/content/abstract/sigtrans;3/103/pc1 http://stke.sciencemag.org/cgi/content/abstract/sigtrans;4/172/ec137
http://stke.sciencemag.org/cgi/content/abstract/sigtrans;4/192/ra63 http://stke.sciencemag.org/cgi/content/abstract/sigtrans;6/261/ec36
's sites:ScienceThe editors suggest related resources on
References http://stke.sciencemag.org/cgi/content/full/sigtrans;2/63/ra13#BIBL
14 article(s) hosted by HighWire Press; see: cited byThis article has been
http://stke.sciencemag.org/cgi/content/full/sigtrans;2/63/ra13#otherarticles
This article cites 38 articles, 16 of which can be accessed for free:
Glossary http://stke.sciencemag.org/glossary/
Look up definitions for abbreviations and terms found in this article:
Permissions http://www.sciencemag.org/about/permissions.dtl
Obtain information about reproducing this article:
the American Association for the Advancement of Science; all rights reserved. byAssociation for the Advancement of Science, 1200 New York Avenue, NW, Washington, DC 20005. Copyright 2008
(ISSN 1937-9145) is published weekly, except the last week in December, by the AmericanScience Signaling
on May 17, 2013 stke.sciencemag.orgDownloaded from
COMPUTATIONAL BIOLOGY
Dynamic Signaling in the Hog1 MAPK Pathway
Relies on High Basal Signal Transduction
Javier Macia,
1
* Sergi Regot,
2
* Tom Peeters,
2
Núria Conde,
1,2
Ricard Solé,
1,3
Francesc Posas
2
(Published 24 March 2009; Volume 2 Issue 63 ra13)
Appropriate regulation of the Hog1 mitogen-activated protein kinase (MAPK) pathway is essential for
cells to survive osmotic stress. Here, we show that the two sensing mechanisms upstream of Hog1
display different signaling properties. The Sho1 branch is an inducible nonbasal system, whereas the
Sln1 branch shows high basal signaling that is restricted by a MAPK-mediated feedback mechanism.
A two-dimensional mathematical model of the Snl1 branch, including high basal signaling and a Hog1-
regulated negative feedback, shows that a system with basal signaling exhibits higher efficiency, with
faster response times and higher sensitivity to variations in external signals, than would systems without
basal signaling. Analysis of two other yeast MAPK pathways, the Fus3 and Kss1 signaling pathways,
indicates that high intrinsic basal signaling may be a general property of MAPK pathways allowing rapid
and sensitive responses to environmental changes.
INTRODUCTION
Appropriate regulation of signaling through mitogen-activated protein
kinase (MAPK) pathways is necessary to maximize cellular responses
to extracellular stimuli (1,2). Exposure of cells to high osmolarity results
in activation of a conserved family of MAPKsp38 MAPK in mammals
and Hog1 in yeast. In vivo replacement of components of the Hog1
MAPK pathway by their mammalian counterparts showed that there is
a strong functional preservation of these MAPK pathways from yeast to
mammals (3). Exposure of yeast to high osmolarity results in rapid activation
of the Hog1 MAPK signaling pathway, which coordinates the adaptive
program required for cell survival during periods of osmotic stress (3,4).
Activation of the Hog1 MAPK is mediated by two independent up-
stream sensing mechanisms that lead to the activation of either the MAPK
kinase kinase (MAPKKK) Ssk2 (or the functionally redundant Ssk22) or
Ste11 (fig. S1). The Sho1 branch involves the transmembrane protein
Sho1 and the mucin-like transmembrane proteins Hkr1 and Msb2, which
are the potential osmosensors of this branch of the HOG pathway (57).
Although additional components still need to be identified to completely
understand signaling from the Sho1 branch, signaling requires the guano-
sine triphosphatase (GTPase) Cdc42, the adaptor protein Ste50, and the
kinases Ste20 and Cla4, which are members of the PAK (p21-activated
protein kinase) family (711). When stimulated by osmotic stress, the
Sho1 branch activates the MAPKKK Ste11 and, subsequently, the MAPKK
Pbs2 (12). Although the exact mechanism of Pbs2 activation by the Sho1
module remains unclear, Cdc42, Ste50, and Sho1 act as adaptor proteins
that control the flow of the osmotic stress signal from Ste20 and Cla4 to
Ste11, and then on to Pbs2 (13,14).
A second sensing mechanism involves the two-componentosmosensor
composed of Sln1, Ssk1, and Ypd1. The Sln1 transmembrane osmosensor
has intrinsic histidine kinase activity and is a homolog of bacterial two-
component signal transducers. Using a phosphorelay mechanism involving
the Ypd1 and Ssk1 proteins, Sln1 inhibits the activity of Ssk1, which con-
trols the activity of the MAPKKKs Ssk2 and Ssk22 (15,16). Therefore,
under normal osmotic conditions, active Sln1 histidine kinase maintains
Ssk1 in its inactive phosphorylated form, whereas, in response to osmotic
stress, the kinase activity of Sln1 is inhibited and then active Ssk1 induces
Ssk2 activity. Signaling from either of the two branches leads to phospho-
rylation of the MAPKK Pbs2 and activation of the MAPK Hog1.
Kinetic analyses of Hog1 phosphorylation using mutants in each of
the two branches of the pathway revealed that they respond differently to
osmotic stress (17). Band-shift analysis with a microfluidic device
showed that, whereas the Sln1 branch of the pathway is capable of fast
signal integration to repeated stimuli, the Sho1 branch does not, suggest-
ing that the signaling properties of the two branches are different (18).
The origin of such differences, as well as their biological relevance, is
still unclear.
Signaling through the MAPK pathway is controlled not only by activity
of the kinases upstream of the MAPK, but also by the activity of protein
phosphatases and various feedback systems. Inactivation of the MAPK is
specifically controlled by direct dephosphorylation by protein phospha-
tases. The type 2C serine/threonine protein phosphatases and the protein
tyrosine phosphatases (Ptp2 and Ptp3) decrease Hog1 and Pbs2 activities
(19). Closure of the Fps1 glycerol channel has been proposed to act as a
feedback mechanism that limits sensor activation in the HOG pathway
(20). In addition, phosphorylation of Sho1 upon Hog1 activation seems
to be important to decrease signaling through this branch (21). Mathemat-
ical modeling and single-cell analyses have shown that inactivation of the
pathway by a mechanism independent of transcription might be important to
modulate acute responses; whereas transcription-dependent mechanisms
might be important for proper adaptation to future stimuli (20,22,23). There-
fore, although most of the components of the signaling pathway have been
defined, the signaling properties of the pathway are still poorly understood
(24,25).
Here, we use a chemical inhibitor of the MAPK Hog1 and extensive
signal quantification to show that the HOG pathway is controlled by high
basal signaling counteracted by a negative feedback regulatory system.
We developed a two-dimensional mathematical model that provided a
1
ICREA-Complex Systems Laboratory, Universitat Pompeu Fabra, E-08003
Barcelona, Spain.
2
Cell Signaling Unit, Departament de Ciències Experimen-
tals i de la Salut, Universitat Pompeu Fabra, E-08003 Barcelona, Spain.
3
Santa
Fe Institute, Santa Fe, NM 87501, USA.
*These authors contributed equally to this work.
To whom correspondence should be addressed. E-mail: ricard.sole@upf.
edu (R.S.) and francesc.posas@upf.edu (F.P.)
RESEARCH ARTICLE
www.SCIENCESIGNALING.org 24 March 2009 Vol 2 Issue 63 ra13 1
on May 17, 2013 stke.sciencemag.orgDownloaded from
general analytic paradigm based on sensor and target architecture that can
be applied to any signaling pathway and showed that a system displaying
high basal signaling, such as the HOG pathway, exhibits higher efficiency
than systems without basal signaling, in terms of faster response and high-
er sensitivity to small variations in extracellular stimuli. Similar to the
HOG pathway, signaling through the MAPK pathway controlling the
mating response also seems to be controlled by a similar design, indicating
that high basal signaling coupled to negative feedback may be a common
trend in MAPK pathways.
RESULTS
The Sln1 and Sho1 branches of the HOG pathway have
different signaling properties
To understand the underlying properties of the two branches of the HOG
MAPK pathway, we performed several quantitative time-course and dose-
response experiments, monitoring Hog1 phosphorylation in mutants in
which both or only one branch of the pathway was active wild-type yeast,
ssk2/ssk22 double mutant (Sho1 branch active), and ste11 or ste50 mutant
(Sln1 branch active). Cells were subjected to osmotic stress (0.07 to 0.8 M
NaCl), then fixed at various times, and total and phosphorylated Hog1
were detected with specific antibodies by quantitative Western blotting
(see Materials and Methods). By plotting the percentage of phosphoryl-
ated Hog1 relative to the maximum in wild type over time, it is clear that
the two branches contribute differently to Hog1 phosphorylation (Fig. 1).
Cells lacking the Sln1 branch (ssk2 ssk22 mutant) do not respond to low
osmolarity (up to 0.1 M NaCl), display slower responses at each osmolar-
ity, and show a lower maximum response than the responses of the wild
type or ste50 mutant. In contrast, the ste50 mutant cells respond as fast as
wild-type cells, with a maximum response that is similar to that of wild-
type cells, but with a shorter duration (Fig. 1A). At the highest osmolarities
tested, the amplitude of the response of the wild type reached a maximum
and plateaued. At each higher osmolarity, the duration of the maximal
response period was longer. Thus, although both branches seem important
for achieving the wild-type response, the Sln1 branch is critical for setting
the speed and maximum amplitude of the response. Physiologically, these
differences in signaling properties result in mutants lacking signaling through
the Sln1 branch that are more sensitive (reduced growth), both on solid and
in liquid media, to osmotic stress than are mutants lacking signaling through
the Sho1 branch (Fig. 1, B and C).
The fact that the MAPKK Pbs2 receives the signal from two upstream
mechanisms, Sln1 and Sho1, with different dynamic responses to osmotic
stress suggests that Pbs2 may function as a signal integrator involved in
changing the response from one that is dose encoded,when the signal is
generated by the Sln1 branch, to one that is duration encoded,when the
signal is generated by the Sho1 branch. This would suggest that a reduction
in Pbs2 activity should result in a change on both amplitude and response
duration. Through orthogonal targeting, we created a yeast strain carrying
a mutant allele of PBS2 (pbs2as) that is specifically inhibited by the small-
molecule inhibitor 1NM-PP1 and a cell-permeable analog of 1NM-PP1, SPP86.
After incubation of this pbs2as strain with different concentrations of in-
hibitor and exposure to 0.2 M NaCl, analysis of total and phosphorylated
Hog1 in cell extracts showed that inhibition of Pbs2 reduces the maximum
amplitude of the response (fig. S2); thus, Pbs2 is a limiting factor in Hog1
signaling and suggests that Pbs2 not only transmits information to Hog1 but
might be critical to integrate different kinetic responses from the upstream
branches.
ABC
Fig. 1. Inactivation of any of the two branches of the HOG pathway
results in cells with less efficient signaling and increased sensitivity
to stress. (A) Hog1 phosphorylation is specifically affected in mu-
tants of the Sln1 branch. Different salt concentrations were added
to the indicated strains and quantification of total and phosphory-
lated Hog1 (P-Hog1) was assessed by quantitative Western
blotting. The data represent the percent of P-Hog1 relative to the
maximum in wild type (wt) and are presented as the mean ± SD
from three independent experiments. (B) Both branches of the
HOG pathway contribute differentially to growth upon increasing
osmolarity. The ssk2 ssk22 strain is more sensitive to osmotic
stress than is the ste50 strain. Serial dilutions of indicated strains
were spotted onto YPD and salt plates and growth was scored after
4 days. (C) Cell growth was assessed in liquid media containing
different NaCl concentrations.
RESEARCH ARTICLE
www.SCIENCESIGNALING.org 24 March 2009 Vol 2 Issue 63 ra13 2
on May 17, 2013 stke.sciencemag.orgDownloaded from
Inhibition of the activity of Hog1 increases
its phosphorylation in both the absence and the
presence of stress
In the absence of stress, catalytically inactive Hog1 shows higher basal
phosphorylation than does wild-type Hog1, and stress causes a prolonged
increase in its phosphorylation compared to wild-type Hog1 (26,27). In
addition, combined deletion of the phosphatases that down-regulate active
Hog1 is lethal unless HOG1 or PBS2 is deleted (28,29). Both results are com-
patible with constant signaling to Hog1, even in the absence of stress, which
must be down-regulated for the cells to maintain viability. If Hog1 is con-
stantly phosphorylated and its activity is required for its dephosphorylation,
then specific inhibition of the kinase activity of Hog1 should result in
increased phosphorylation of Hog1. To study whether Hog1 receives
constant signaling that is silenced by a feedback regulatory loop that de-
pendsonHog1activity,wecreatedaHog1mutant(hog1as)thatissensitive
to the small-molecule inhibitors 1NM-PP1 and SPP86 (30,31). Wild-type
or hog1as cells carrying an empty vector or a vector containing wild-type
HOG1 or a catalytically inactive hog1 (hog1kn) were grown in selective
medium and the inhibitor was added at time 0. Strikingly, addition of the
inhibitor to nonstressed cells resulted in rapid phosphorylation of Hog1
(Fig. 2A). The addition of the inhibitor to cells previously subjected to
stress (10 min) prevented dephosphorylation of hog1as as reported be-
fore (31) (fig. S3). Addition of the inhibitor, once Hog1 was dephosphoryl-
ated after exposure to stress and adaptation, resulted in rephosphorylation
of the kinase (Fig. 2B). Therefore, our results suggest that inhibition of the
kinase activity of the MAPK results in rapid phosphorylation of kinase
even in the absence of stress. Correspondingly, in the absence of stress,
the phosphorylated form of Hog1 does not appear in wild-type cells upon
addition of the inhibitor or in cells containing a wild-type allele of HOG1
together with the hog1as mutant allele (Fig. 2A). Osmotic stress triggered
the accumulation of hog1as fused to green fluorescent protein (GFP) in the
nucleus, whereas in the presence of the inhibitor, hog1as-GFP did not ac-
cumulate in the nucleus (fig. S4).
The Sln1 branch of the HOG pathway is responsible
for high basal signaling to Hog1
Several potential regulatory feedback loops have been described for the
Hog1 pathway. To test whether the control of Hog1 phosphorylation is
exerted by a fast (transcription-independent) or slow (transcription-
dependent) feedback loop, we analyzed phosphorylation of the kinase
upon blockage of transcription or translation. Phosphorylation of Hog1
upon inhibition of its kinase activity is independent of transcription and
translation, because the presence of cycloheximide (a translation inhibitor)
or thiolutin (a transcription inhibitor) did not affect Hog1 phosphorylation
(fig. S5A); these drugs did block the induction of Stl1 protein production
(fig. S5B). The rapid production of phosphorylated Hog1 when Hog1 ac-
tivity is inhibited does not depend on changes in glycerol concentration,
which is known to regulate the HOG pathway. Both hog1as cells with
wild-type FPS1, which encodes a glycerol channel, and hog1as cells with
fps1D1, which encodes a constitutively open glycerol channel, show accu-
mulation of phosphorylated Hog1 in response to inhibition of Hog1 kinase
activity (Fig. 3A). Furthermore, although the amount of phosphorylated
Hog1 produced in response to the inhibitor, as well as in response to osmotic
A
B
Fig. 2. Inhibition of Hog1 activity results in its phosphorylation. (A) Hog1
catalytic activity is required to prevent accumulation of the phosphorylated
form (P-Hog1) in the absence of stress. The indicated strains (wt and hog1as)
containing a vector control or a wild type or catalytically inactive mutant were
grown in SD medium and 5 mM inhibitor (SPP86) was added at time 0. Phos-
phorylated Hog1 (P-Hog1) was assessed as in Fig. 1A. (B) Inhibition of hog1as
after adaptation to stress also induces accumulation of its phosphorylated
form. Wild-type and hog1as strains were stressed with 0.4 M NaCl and in-
hibitor (5 mM SPP86) was added at time 40 min as indicated by the arrow.
Hog1 was detected by chemoluminescence with specific antibodies.
A
B
Fig. 3. Phosphorylation of inactive Hog1 depends on the Sln1 branch and
does not require closure of the glycerol channel. (A)hog1as cells carrying
the wild-type glycerol channel (FPS1) or a constitutively active mutant
(fps1D1) were grown in SD medium and inhibitor (5 mM SPP86) was added
at time 0. Samples were collected at the indicated times and total and
phosphorylated Hog1 were detected by chemoluminescence with specific
antibodies. (B) The indicated strains were grown in YPD, SPP86 (5 mM)
was added at time 0, and phosphorylated Hog1 (P-Hog1) was assessed
as in Fig. 1A.
RESEARCH ARTICLE
www.SCIENCESIGNALING.org 24 March 2009 Vol 2 Issue 63 ra13 3
on May 17, 2013 stke.sciencemag.orgDownloaded from
stress, was less in cells lacking FPS1, possibly due to increased internal
glycerol content, FPS1 strains still exhibited phosphorylation of hog1as
in the presence of the inhibitor (fig. S6, A and B). Thus, basal signaling
through the HOG pathway is controlled by a fast feedback regulatory mech-
anism that is dependent on the kinase activity of Hog1.
The basal signaling to Hog1 could be produced by either or both of the
upstream sensing mechanisms that converge on Pbs2. To dissect which
branch of the pathway, Sln1 or Sho1 or both, is responsible for high basal
signaling to Hog1, we measured Hog1 phosphorylation after addition of
inhibitor to hog1as cells that also contained ssk2 ssk22 or ste50 mutations.
Phosphorylation of Hog1 upon inhibition of its kinase activity in a ste50
strain showed similar kinetics to those cells in which both branches of the
pathway were active (Fig. 3B). Similar results were obtained in cells
containing sho1 mutation (fig. S7). Thus, loss of signal through the
Sho1 branch does not affect the basal signaling to Hog1. However,
Hog1 phosphorylation in response to inhibition of its kinase activity is
completely absent when the Sln1 branch is inactive (Fig. 3B). Therefore,
the two branches of the HOG pathway display different signaling proper-
ties, with the Sln1 branch mediating high basal signaling to Hog1 and the
Sho1 branch serving as an inducible system.
A mathematical model that includes high basal
signaling and a negative feedback loop describes the
dynamics of the HOG pathway
Previous comprehensive mathematical models for the yeast HOG path-
way included most of the components of the pathway or potential regulatory
loops, but they did not include basal signaling (18,22,23). Therefore, we
developed a new mathematical model that includes a stress-independent,
basal signal for the sensor and the requirement for negative feedback reg-
ulated by the target Hog1, which involves a minimal set of nonlinear in-
teractions (Fig. 4A and Supplementary Materials). The advantage of a model
that involves just the sensor and the target components is that it is amenable
to analytical mathematical analysis without requiring parameter estimation.
Despite its simplicity, this model fully describes the basic mechanisms
behind the observed experimental results (Fig. 4, B to D) and can be studied
analytically by considering only the interactions between the sensor protein
Sln1 (designated A in the model) and the target protein Hog1 (designated B
in the model).
Once cells are subjected to stress, the concentration of signaling-competent
Sln1 (A*) increases as does the active form of Hog1 (B*) (Fig. 4A). Thus,
A*andB* represent the main players immediately after stress, before
cellular adaptation has begun. Assuming that the total concentrations of
the proteins remains constant (22), one can reduce the pathway to a two-
dimensional model, which can be described by a set of two differential
equations, namely, dA*/dt=g(A*, S) (Equation 27 in Supplementary Ma-
terials) and dB*/dt=f(A*,B*) (Equation 28 in Supplementary Materials)
where Srepresents osmotic stress and fand gare nonlinear Hill-like re-
sponse functions. These equations describe the production of active Sln1
in response to stress (designated Sin the model) and the variation in Hog1
phosphorylation that depends on the concentration of active Sln1 and the
Hog1-mediated negative feedback loop (see Supplementary Materials for
a full description of the model and equations).
Fig. 4. A mathematical model of
the Sln1 branch describes the
signaling properties of the path-
way. (A) The minimal backbone
of the model includes the sensor
Sln (A, inactive state; A*, active
state), the MAPK Hog1 (B, in-
active state; B*, active state), os-
molites (G), and a stress-regulated
osmolite channel (C). Activation
of A occurs spontaneously (basal
activation). Deactivation of A* is
stress-dependent, K-a(S), where
Sis osmotic stress. A* stimulates
the formation of B*. B* represses
its own production by either en-
hancing B* deactivation or by
inhibiting B activation (not shown)
and enhances osmolite (G) pro-
duction, which exits through the
stress-regulated channels (C).
(B) The model describes the ex-
perimental data with the dynam-
ics of B* [phosphorylated Hog1
(P-Hog1)] from the model match-
ing the experimental data (colored
dots) (see Supplementary Mate-
rials for parameters) in response
to different Sconditions (concen-
trations of salt). (Cand D) The
model describes that by breaking the B*-dependent feedback, addition of the inhibitor in the experimental condition (arrow) induces phosphorylation of
B after adaptation to stress (C) and in the absence of stress (D).
RESEARCH ARTICLE
www.SCIENCESIGNALING.org 24 March 2009 Vol 2 Issue 63 ra13 4
on May 17, 2013 stke.sciencemag.orgDownloaded from
With this model we explored two hypothetical scenariosone in which
the target of the feedback regulatory loop was Hog1 itself and one in which
the target was a component upstream of Hog1. Both cases are biologically
distinct, but the model displays the same qualitative dynamic response (see
Supplementary Materials for the detailed analysis of each scenario). Thus,
this new mathematical model incorporates the basal signaling and the
Hog1-dependent negative feedback loop and allows analysis of the signal
dynamics independently from the specific target of the feedback loop.
The two-dimensional mathematical model demonstrates
why high basal signaling is essential for the stress response
To understand the properties underlying the pathway with the defined
characteristics we studied the geometrical features of the so-called nullclines
of the system (32), which allow one to determine the maximum and
minimum states of the system. In this system, the nullclines are defined
by the two curves g(A*, S)=f(A*, B*) = 0, which allows us to determine
the presence of equilibrium points and their dynamical behavior. Thus, we
can analyze the relationship between active Hog1 and active Sln1 (Fig. 5A).
These curves determine how A* and B* change in time under stress.
Because g= 0 means that A* does not change, and f= 0 means that B*
does not change, the system will move parallel to the Aaxis when approach-
ing for parallel to the Baxis when approaching g. In our model, g=0isa
vertical line and the location of this line depends on the values of the
stress within a range A*
min
<A*<A*
max
, where A*
min
is the nonstress
(basal) value and A*
max
its upper value. Moreover, independently from
the specific set of Hill factors used, the f= 0 nullcline exhibits, for both
theoretical scenarios, generic geometrical features that determine its shape.
The f= 0 nullcline has a horizontal asymptote located at B
v
, crossing the
axis at A*=0,B* = 0 (Fig. 5A), without local maximum or minimum
values within the interval (0, B
v
). Therefore, when cells are adapted to the
environmental osmolarity, the system is in a stable state (A*
min
,B*
min
).
However, when cells are subjected to hyperosmotic stress, the nullcline
g= 0 is abruptly displaced to higher values of A* (that is, A*
max
) and
the previous stable state becomes unstable. After adaptation to the new
osmotic conditions, the nullcline g= 0 moves again to lower values of
A*, and the system evolves toward a new stable state (A
o
*, B
o
*). It is
worth noting that immediately after stress, cells are not adapted and the
production of active Hog1 (B*) is predominantly determined by the tra-
jectory in the A-B plane that depicts the evolution of A* versus B* com-
prised between (A*
min
,B*
min
)and(A*
max
,B*
max
)(Fig.5A).
Duetotheshapeofthenullcline,asystemstartingfromA*
min
=B*
min
=0
(a system without basal signal) will slowly increase, because the vertical
component fof the field has its lowest value exactly where the horizontal
component ghas its highest value. Therefore, the system will display a
delayed response. In contrast, if the system starts with A*
min
>0(inthe
presence of basal signaling), then the response is faster and the delay is
reduced or avoided. Even in cases were the value of A*
min
is not very high
(relatively low basal signaling), the response will be faster because of the
position of the initial point on the region of the nullcline f= 0 with the
steeper slope (Fig 5B). Simulations with different amounts of basal sig-
naling (A*
min
) produce the same relative increase in Hog1 phosphorylation
(B*), showing that the slope of the curve increases with the basal signal,
thus generating faster responses (Fig. 5B).
These results are consistent with the experimental results obtained for
the basal signaling branch Sln1 and for the nonbasal branch Sho1. The
Sln1 branch does not exhibit a delayed response to hyperosmotic stress,
whereas the Sho1 branch does exhibit a delay (Fig. 6A). As a result of the
same geometrical constraints, systems with very low or no basal signal
exhibit less sensitivity to small changes in the external osmolarity. Small
changes in the external osmolarity lead to small changes in the concen-
tration of the activated sensor A*. If the initial point is at (0,0) or near it,
changes in A*willhaveasmalleffectonB* because in this region the
nullcline f= 0 exhibits a lower slope. However, for systems with basal sig-
nal, the initial point will be located in regions of the nullcline where the
slope is steeper, and thus, small changes in A* will trigger large changes in
B*. Eventually, if the basal signal is too high, the initial point is located in
Fig. 5. Nullcline analysis reveals
the need for basal signaling in
the response to osmotic stress.
(A) Phase space A*versusB*.
Concentrations are determined
bytheequationdA*/dt=g(A*, S)
and dB*/dt=f(A*,B*). Nullcline
g= 0 defines a vertical line whose
location depends on the A*val-
ue without stress (basal signal).
Without basal signaling and no
stress, the nullcline is located at
the origin (blue line). If basal sig-
nal is present, then A* > 0 (green
line). The equilibrium state is de-
termined by g(A*,S)=f(A*,B*) =
0 (circle). After stress, nullcline
g= 0 shifts toward higher A*
values (violet line), and the sys-
tem moves to a new state (square).
The trajectory is confined in the gray region. The slope of f= 0 for small A*
values introduces a slowdown in the B* response through time, but once
A* is large enough (A*>A*
min
), B* grows faster. The inset in (A) shows the
ratio of the vertical and horizontal components of the field (f/g)whena
small increase in the concentration of active sensor A*, with respect to
its stable state, is performed. The curve displays an optimum at some A*
p
.
If A*
min
is close to A*
p
, the stress response is faster. (B)SimulationsofB*
versus time for different amounts of basal signaling. By increasing A*(see
trajectories in the inset) higher slopes are observed. (See Supplementary
Materials for details about nullcline analysis.)
RESEARCH ARTICLE
www.SCIENCESIGNALING.org 24 March 2009 Vol 2 Issue 63 ra13 5
on May 17, 2013 stke.sciencemag.orgDownloaded from
a region of the nullcline f= 0 where the curve is near the horizontal as-
ymptote with a lower slope, and, as a consequence, the pathway will lose
sensitivity. Based on these data, cells adapted to media with different os-
molarities should display different basal signaling. Correspondingly, inhi-
bition of hog1as in different media results in altered phosphorylation
kinetics of Hog1. Thus, Hog1 is phosphorylated faster after inhibition
of Hog1 activity in cells adapted to a medium with higher osmolarity
(Fig.6,BandC).
The overall behavior of the mathematical model is summarized in Fig.
5A, which explains the observed results of reduced sensitivity and slower
response of ssk2 ssk22 cells (Fig. 6A and fig. S8) and reveals that a com-
bination of basal signal, nonlinearity associated with the presence of Hill-
like responses, and a negative feedback provide for the high efficiency
and sensitivity of the pathway and its quick response to even small changes
in osmolarity.
The Fus3 and Kss1 MAPK signaling pathways
are also controlled by a high basal signal
In yeast, other MAPK signaling pathways exist that respond to external
stimuli (2). Both the Fus3 and Kss1 MAPKs are activated by pheromone
(33). We created strains in which only one of the two kinases was present
(fus3 and kss1 strains) and strains in which the remaining kinase could be
inhibited (Fus3as kss1;Kss1as fus3) and tested whether inhibition of the
kinase activity resulted in phosphorylation of the MAPK in the absence
of pheromone. As expected, the addition of pheromone stimulated the phos-
phorylation of wild-type and mutant fus3as, which, in turn, resulted in cell
cyclearrestatG
1
and induction of FUS1 expression (FUS1p:GFP) (Fig. 7).
Interestingly, addition of the inhibitor to the fus3as or kss1as cells triggered
Fus1 or Kss1 phosphorylation (Fig. 7, A and D). However, inhibition of the
kinase activity of the MAPK ( fus3as cells) failed to trigger downstream pro-
cesses, such as arrest in G
1
and FUS1 expression (Fig. 7, B and C). It is
worth noting that, whereas the presence of wild-type Kss1 did not affect the
phosphorylation of fus3as in the presence of inhibitor, the presence of wild-
type Fus3 strongly reduced the phosphorylation of kss1as in the presence of
the inhibitor, indicating that the two kinases might have different input into
the feedback loop(s) in response to pheromone (fig. S9). Both Fus3 and
Kss1 MAPK pathways also seem to include high basal signaling. Thus,
three MAPKs involved in various biological processes, from responding
to osmotic stress to initiating developmental programs, seem to have high
basal signaling that is inhibited by negative feedback regulatory loops con-
trolled by kinase activity of the MAPKs. Intrinsic basal signaling may be a
general property of MAPK pathways, allowing efficient response to envi-
ronmental changes.
DISCUSSION
Signaling through MAPK pathways is essential for cellular response to
extracellular stimuli. Yeast cells activate the Hog1 MAPK to control cel-
lular response and adaptation to osmotic stress and, therefore, precise mod-
ulation of Hog1 activity is necessary to maximize cell survival. Activation
of the MAPK is achieved by two upstream sensing mechanisms with slight-
ly different kinetics of response and sensitivity to osmotic stress [(17)and
our results (Figs. 1 and 6A)], suggesting that both branches may have
different signaling capacity. Partial inhibition of the MAPKK Pbs2 shows
that the kinase not only integrates the signal from the two branches of the
pathway but may also transform the input from a dose-encoded signal to
one that is duration encoded.
The use of a kinase-inhibitable form of Hog1 showed that inactivation
of its kinase activity results in its own rapid and strong phosphorylation,
A
B
C
Fig. 6. Basal signaling is essential for proper dynamic response of the
HOG pathway. (A) Signaling through the Sho1 branch is slower than
signaling through the Sln1 branch. Indicated strains were grown in me-
dium, subjected to 0.8 M NaCl, and phosphorylated Hog1 was assessed
as in Fig. 1A. (B) Reduction of the external osmolarity results in a slower
response. The ste50 strain was grown in YPD medium, cells were centri-
fuged and resuspended in medium or in distilled water as indicated, then
cells were subjected to 0.8 M NaCl stress and Hog1 phosphorylation was
assessed as in (A). (C) The kinetics of phosphorylation of Hog1 after inhi-
bition of Hog1 activity depends on external osmolarity. hog1as cells were
grown in YPD or YPD plus the indicated salt concentrations, spun, and
resuspended in YPD at the same salt concentration, YPD alone, or YPD
diluted in distilled water at the indicated dilution factors. The inhibitor
(5 mM SPP86) was added at time 0 and Hog1 phosphorylation was as-
sessed as described in (A).
RESEARCH ARTICLE
www.SCIENCESIGNALING.org 24 March 2009 Vol 2 Issue 63 ra13 6
on May 17, 2013 stke.sciencemag.orgDownloaded from
even in the absence of stress, indicating that there is a basal signal into
the MAPK. This is consistent with previous studies that, upon stress, a
catalytically inactive kinase remained phosphorylated longer than did the
wild-type MAPK (26). The use of the inhibitable Hog1 together with spe-
cific mutations in upstream components of the pathway showed that the
basal signal to Hog1 comes from only the Sln1 branch of the pathway,
with the Sho1 branch acting as an inducible system without basal sig-
naling. Thus, the Sln1 branch is possibly the key determinant of the signal-
ing properties of the pathway. Cells with mutations in the Sln1 branch of
the pathway are more sensitive to stress than are cells with mutations in
the Sho1 branch. Our results are consistent with previous bandwidth anal-
ysis with a microfluidic device, which showed that only the Sln1 branch
of the pathway was capable of fast signal integration (18). In some yeast
species, the Sho1 branch is either absent or is not connected to Hog1 MAPK
signaling but instead has a role in morphogenesis rather than osmo-
sensing (34).
Thus, in contrast to the typical assumption that the signal originates in
response to the presence of a stimuli, our experimental studies indicate
that the signaling through several yeast MAPK pathways depends on high
basal signal transduction that must be constantly counteracted by a fast-
acting feedback mechanism that is controlled by the kinase activity of the
MAPK. Although we do not know the exact nature of such a negative
feedback, it does not involve transcription or translation and is indepen-
dent of glycerol accumulation. Cells with a constitutively open glycerol
channel responded in the same manner as the wild type. Protein phospha-
tases are good candidates for mediating the negative feedback. Elimination
of phosphatases results in hyperactivation of the Hog1, and studies on
cells exposed to stress pulses followed by frequency determination showed
the presence of a fast-acting negative feedback regulatory loop and sug-
gested that it could involve protein phosphatases (23). For cells grown in
control or osmotic stress conditions, we can estimate the ratio of the phos-
phorylated and dephosphorylated states of Hog1 in the presence of the
inhibitor by assuming that because the catalytic activity of Hog1 is inhib-
ited and the external osmolarity is maintained constant, the rates of phos-
phorylation and dephosphorylation are constant. Under basal conditions,
the ratio K
phosphorylation
/K
dephosphorylation
1.1, whereas under 0.4 M salt
stress conditions, K
phosphorylation
/K
dephosphorylation
13.3. These ratios have
been calculated such that they reproduce the different slopes and the rel-
ative difference of phosphorylated Hog1 in the presence of inhibitor (Fig.
6C). We have also analyzed the model to estimate which of the two com-
ponents of the HOG pathway is the target of the negative feedback. The
model shows that regardless of whether the target of the feedback regula-
tory loop is Hog1 or is upstream of Hog1, the qualitative results are simi-
lar. However, the model also indicates that if the target of the feedback
loop is located upstream of Hog1, the system seems to be more efficient
(see the Mathematical Analysis section of the Supplementary Materials).
Feedback regulatory loops have profound implications and, typically,
systems with negative feedback show higher robustness against external
and stochastic noise (35,36), thereby increasing the efficiency in signal trans-
mission. However, the presence of negative feedbacks introduces several con-
straints on the dynamics of the systems because they introduce delays in the
response and reduce the sensitivity. Thus, the high basal signal may function
to counteract the constraints created by the negative feedback loops.
To explore the properties of a system with high basal signal restricted
by a negative feedback loop, we developed a two-dimensional mathemat-
ical model of the HOG pathway based on the minimal backbone of inter-
actions that reproduces the experimental results and allows analytical
analysis of the dynamic properties of the pathway. The analysis of this
model was simplified by considering that just after stress only the sensor-
target protein interaction and the negative feedback are relevant to explain
the basic dynamical behavior. The analysis of the main interactions showed
that the dynamics of the pathway immediately after stress depend on the
balance between the negative feedback and the basal signaling from the sen-
sor. The feedback tends to delay Hog1 phosphorylation; whereas the basal
signal tends to enhance its phosphorylation. Therefore, the sensitivity to
small changes in the external osmolarity depends on the amount of basal
A
BC
D
Fig. 7. The MAPKs Fus3 and Kss1 receive basal signaling in the
absence of stimuli. (A) Inhibition of Fus3 activity leads to its phos-
phorylation. Mating pheromone (a-factor) or inhibitor (SPP86)
was added to the FUS3 kss1 (wt; left panels) or fus3as kss1
mutant (fus3as; right panels) and phosphorylated or total Fus3
was assessed with specific antibodies. (B) Inhibition of fus3as
does not result in cell cycle arrest at G
1
.a-Factor or inhibitor
was added to the fus3as kss1 mutant and DNA was assessed
by fluorescence-activated cell sorting (FACS) analysis. (C)Inhi-
bition of fus3as does not result in induction of FUS1 expression.
Expression of GFP driven by the FUS1 promotor was assessed
in the same strain as in (B) by FACS analysis and cells expres-
sing GFP were quantified. (D) Inhibition of Kss1 activity leads to
its phosphorylation. a-Factor or inhibitor was added to the
KSS1 fus3 (wt; left panels) or kss1as fus3 mutant (kss1as; right
panels) and Kss1 phosphorylation was assessed with Kss1-
specific antibodies.
RESEARCH ARTICLE
www.SCIENCESIGNALING.org 24 March 2009 Vol 2 Issue 63 ra13 7
on May 17, 2013 stke.sciencemag.orgDownloaded from
signaling, such that this system is much more sensitive than one without
basal signaling. Correspondingly, cells deficient on the Sln1 branch show
reduced sensitivity to small variations in osmolarity, responded more
slowly to osmotic stress, and were more osmosensitive. Because high basal
signaling seems to be a key factor that counteracts the lack of response
and delay caused by the negative feedback, the dynamics of the response
should change upon variations in the basal signal. We showed that cells
that were adapted to a reduced osmolarity exhibited reduced basal signal-
ing, reduced capacity to cope with small variations in stress conditions,
and responded more slowly to osmotic stress (Fig. 6B). Therefore, in this
system with a negative feedback loop, a high basal signal is critical for de-
termining the signaling capacity.
To explore how common basal signaling with negative feedback is,
we extended our studies to the pheromone-responsive Fus3 and Kss1
MAPK pathways. Both kinases became phosphorylated upon inhibition
of their kinase activity, even in the absence of pheromone, suggesting that
these MAPK signaling pathways are also controlled by high basal signal
together with a MAPK-dependent negative feedback. This is consistent
with a study that identified Sst2 as the target of the feedback loop con-
trolledbyFus3(37). Thus, three MAPK pathways appear to have developed
high basal signaling repressed by MAPK-dependent negative feedback loops,
which implies that high intrinsic basal signaling could be a general property
of MAPK pathways, allowing efficient response to environmental changes.
MATERIALS AND METHODS
Quantitative Western blotting
Samples were taken in mid-exponential growth phase and fixed in 20%
trichloroacetic acid for SDSpolyarylamide gel electrophoresis and inmu-
noblotting. Hog1, Fus3, and Pbs2 were detected with antibodies specific
for these proteins (Santa Cruz), phosphorylated Hog1 with an antibody
against phospho p38 (Cell Signaling), and both P-Kss1 and P-Fus3 with
an antibody against phospho p44/42 (Cell Signaling). Quantif ication analysis
was performed by fluorescent detection with the IRDye 800CW donkey
antibody against goat immunoglobulin G (IgG) and the IRDye 680 don-
key antibody against rabbit IgG (LI-COR Biosciences) and the ODISSEY
application software 2.1 (LI-COR Biosciences). All phosphorylated Hog1
values were normalized against the 10-min sample taken from the wild type
stressed with 0.4 M salt.
Flow cytometry
The pheromone pathway was activated with a-factor (2 mg/ml) and DNA
content was assessed by staining with propidium iodide.
Inhibitors
1NM-PP1 was used at 5 mM to inhibit analog-sensitive mutants (38). SPP86,
a cell-permeable adenine analog similar to 1NM-PP1, was used at 5 mM.
Strains and plasmids
W303-1A (wild type) and derivatives ssk2 ssk22,ste50,ste11,sho1,hog1
T100G
(hog1as), ssk2 ssk22 hog1as,ste50 hog1as,fps1 hog1as,bar1 FUS1::GFP,
bar1 kss1 fus3 FUS1::GFP P(FUS3):: fus3
Q93A
(fus3as), bar1 kss1
fus3 FUS1::GFP P(KSS1)::kss1
E94A
(kss1as), bar1 fus3 FUS1::GFP
P(FUS3)::fus3as,bar1 kss1 FUS1::GFP P(KSS1)::kss1as, and pbs2
M435A
(pbs2as) were used in this study. YCPlac111 with HOG1-3HA or hog1
K52S,53N
-
3HA (hog1-KN), YEPlac195 with FPS1 or fps1D13-230 (fps1D1)and
pRS416 with HOG1-GFP or hog1as-GFP plasmids were used in this study.
Growth conditions
Cultures were maintained on YPD (1% yeast extract, 2% Bacto Peptone,
2% glucose) or on the appropriate synthetic dropout (SD) medium [0.17% yeast
nitrogen base without amino acids, 0.5% ammonium sulfate, 2% glucose,
Complete Supplement Mixture quadruple dropout (Qbiogene) (0.6 g/liter),
and histidine, tryptophan, uracil, or leucine combined (40 mg/liter)] to main-
tain the appropriate selection for maintenance of the plasmids. Cells were
grown at 30°C and NaCl was added at various concentrations as indicated.
Solid culture was performed on 1.5% agar plates of YPD or YPD with NaCl
at the indicated concentration. Plates were incubated at 30°C.
SUPPLEMENTARY MATERIALS
www.sciencesignaling.org/cgi/content/full/2/63/ra13/DC1
Mathematical Analysis
Fig. S1. Schematic diagram of the HOG pathway.
Fig. S2. Pbs2 activity is a limiting factor to determine amplitude of Hog1 phosphorylation in
response to osmotic stress.
Fig. S3. Inhibition of Hog1 results in phosphorylation of the MAPK.
Fig. S4. Inhibited Hog1 remains mainly cytoplasmic.
Fig. S5. Inhibition of transcription or translation does not prevent phosphorylation of Hog1.
Fig. S6. Deletion of FPS1 does not abolish hog1as phosphorylation in the presence of inhibitor.
Fig. S7. Neither Sho1 nor Ste50 is essential for phosphorylation of inactive Hog1.
Fig. S8. The Sln1 branch is essential for rapid response to osmotic stress.
Fig. S9. The kinases Fus3 and Kss1 receive basal signaling in the absence of stimuli and
differently control a negative feedback loop.
REFERENCES AND NOTES
1. J. M. Kyriakis, J. Avruch, Mammalian mitogen-activated protein kinase signal transduc-
tion pathways activated by stress and inflammation. Physiol. Rev. 81,807869 (2001).
2. R. E. Chen, J. Thorner, Function and regulation in MAPK signaling pathways: Lessons
learned from the yeast Sacchar omyces cerevisi ae. Biochim. Biop hys. Acta 1773,
13111340 (2007).
3. D. Sheikh-Hamad, M. C. Gustin, MAP kinases and the adaptive response to hyper-
tonicity: Functional preservation from yeast to mammals. Am. J. Physiol. Renal Physiol.
287, F1102F1110 (2004).
4. E. de Nadal, P. M. Alepuz, F. Posas, Dealing with osmostress through MAP kinase
activation. EMBO Rep. 3, 735740 (2002).
5. K. Tatebayashi, K. Tanaka, H.-Y. Yang , K. Yamamoto, Y. Matsushita, T. Tomida, M. Imai,
H. Saito, Transmembrane mucins Hkr1 and M sb2 are putative osmosensors in the SHO1
branch of yeast HOG pathway. EMBO J. 26,35213533 (2007).
6. E. de Nadal, F. X. Real, F. Posas, Mucins, osmosensors in eukaryotic cells? Trends
Cell Biol. 17,571574 (2007).
7. I. Dan, N. M. Watanabe, A. Kusumi, The Ste20 group kinases as regulators of MAP
kinase cascades. Trends Cell Biol. 11, 220230 (2001).
8. V. Reiser, S. M. Salah, G. Ammerer, Polarized localization of yeast Pbs2 depends on
osmostress, the membrane protein Sho1 and Cdc42. Nat. Cell Biol. 2, 620627
(2000).
9. F. Posas, E. A. Witten, H. Saito, Requirement of STE50 for osmostress-induced activation
of the STE11 mitogen-activated protein kinase kinase kinase in the high-osmolarity
glycerol response pathway. Mol. Cell. Biol. 18, 57885796 (1998).
10. D. M. Truckses, J. E. Bloomekatz, J. Thorner, The RA domain of Ste50 adaptor protein
is required for delivery of Ste11 to the plasma membrane in the filamentous growth
signaling pathway of the yeast Saccharomyces cerevisiae.Mol. Cell. Biol. 26,912928
(2006).
11. M. J. Winters, P. M. Pryciak, Interaction with the SH3 domain protein Bem1 regulates
signaling by the Saccharomyces cerevisiae p21-activated kinase Ste20. Mol. Cell.
Biol. 25, 21772190 (2005).
12. F. Posas,H. Saito, Osmotic activation of the HOG MAPK pathway via Ste11p MAPKKK:
Scaffold role of Pbs2p MAPKK. Science 276,17021705 (1997).
13. K. Tatebayashi, K. Yamamoto, K. Tanaka, T. Tomida, T. Maruoka, E. Kasukawa, H. Saito,
Adaptor functions of Cdc42, Ste50, and Sho1 i n the yeast osmoregulatory HOG MAPK
pathway. EMBO J. 25, 30333044 (2006).
14. A. Zarrinpar, R. P. Bhattacharyya, M. P. Nittler, W. A. Lim, Sho1 and Pbs2 act as coscaffolds
linking components in the yeast high osmolarity MAP kinase pathway. Mol. Cell 14,
825832 (2004).
15. F. Posas, S. M. Wurgler-Murphy, T. Maeda, E. A. Witten, T. C. Thai, H. Saito, Yeast
HOG1 MAP kinase cascade is regulated by a multistep phosphorelay mechanism in the
SLN1-YPD1-SSK1 two-componentosmosensor. Cell 86,865875 (1996).
RESEARCH ARTICLE
www.SCIENCESIGNALING.org 24 March 2009 Vol 2 Issue 63 ra13 8
on May 17, 2013 stke.sciencemag.orgDownloaded from
16. F. Posas, H. Saito, Activation of the yeast SSK2 MAP kinase kinase kinase by the
SSK1 two-component response regulator. EMBO J. 17, 13851394 (1998).
17. T. Maeda, M. Takekawa, H. Saito, Activation of yeast PBS2 MAPKK by MAPKKKs or
by binding of an SH3-containing osmosensor. Science 269, 554558 (1995).
18. P. Hersen, M. N. McClean, L. Mahadevan, S. Ramanathan, Signal processing by the
HOG MAP kinase pathway. Proc. Natl. Acad. Sci. U.S.A. 105, 71657170 (2008).
19. H. Saito, K. Tatebayashi, Regulation of the osmoregulatory HOG MAPK cascade in
yeast. J. Biochem. 136, 267272 (2004).
20. S. Hohmann, M. Krantz, B. Nordl ander, Yeast osmoregulation. Methods Enzymol.
428,2945 (2007).
21. N. Hao, M. Behar, S. C. Parnell, M. P. Torres, C. H. Borchers, T. C. Elston, H. G. Dohlman,
A systems-biology analysis of feedback inhibition in the Sho1 osmotic-stress-response
pathway. Curr. Biol. 17,659667 (2007).
22. E. Klipp, B. Nordlander, R. Kruger, P. Gennemark, S. Hohmann, Integrative model of
the response of yeast to osmotic shock. Nat. Biotechnol. 23, 975982 (2005).
23. J. T. Mettetal , D. Muzzey, C. Gomez -Uribe, A. van Oude naarden, The frequency
dependence of osmo-adaptation in Saccharomyces cerevisiae.Science 319, 482484
(2008).
24. S. Hohmann, Osmotic stress signaling and osmoadaptation in yeasts. Microbiol. Mol.
Biol. Rev. 66, 300372 (2002).
25. S. M. ORourke, I. Herskowitz, E. K. OShea, Yeast go the whole HOG for the hyperosmotic
response . Trends Genet. 18,405412 ( 2002).
26. S. M. Wurgler-Murphy, T. Maeda, E. A. Witten, H. Saito, Regulation of the Saccharomyces
cerevisiae HOG1 mitogen-activated protein kinase by the PTP2 and PTP3 protein tyrosine
phosphatases. Mol. Cell. Biol. 17, 12891297 (1997).
27. P. Ferrigno, F. Posas, D. Koepp, H. Saito, P. A. Silver, Regulated nucleo/cytoplasmic
exchange of HOG1 MAPK requires the importin bhomologs NMD5 and XPO1. EMBO
J. 17, 56065614 (1998).
28. T. Jacoby, H. Flanagan, A. Faykin, A. G. Seto, C. Mattison, I. Ota, Two protein-tyrosine
phosphatases inactivate the osmotic stress response pathway in yeast by targeting the
mitogen-activated protein kinase, Hog1. J. Biol. Chem. 272, 1774917755 (1997).
29. M. T akekawa, T. Maeda, H. Saito, Protein phosphatase 2Cainhibits the human
stress-responsive p38 and JNK MAPK pathways. EMBO J. 17, 47444752 (1998).
30. S. Kim, K. Shah, Dissecting yeast Hog1 MAP kinase pathway using a chemical genetic
approach. FEBS Lett. 581,12091216 (2007).
31. P. J. Westfall, J. Thorner, A nalysis of mitogen-activated protein kinase signaling specificity
in response to hyperosmotic stress: Use of an analog-sensitive HOG1 allele. Eukaryot.
Cell 5,12151228 (2006).
32. S. H. Strogatz, Interpreting the human phase response curve to multiple bright-light
exposures. J. Biol. Rhythms 5, 169174 (1990).
33. H. G. Dohlman, J. E. Slessareva, Pheromone signaling pathways in yeast. Sci. STKE
2006, cm6 (2006).
34. M. Krantz, E. Becit, S. Hohmann, Comparative genomics of the HOG-signalling system
in fungi. Curr. Genet. 49,137151 (2006).
35. A. Becskei, L. Serrano, Engineering stability in gene networks by autoregulation. Nature
405,590593 (20 00).
36. T. S. Gardner, J. J. Collins, Neutralizing noise in gene networks. Nature 405,520521 (2000).
37. R.C.Yu,C.G.Pesce,A.Colman-Lerner,L.Lok,D.Pincus,E.Serra,M.Holl,K.Benjamin,
A. Gordon, R. Brent, Negative feedback that improves information transmission in yeast
signalling. Nature 456,755761 (2008).
38. A. C. Bishop, J. A. Ubersax, D. T. Petsch, D. P. Matheos, N. S. Gray, J. Blethrow, E. Shimizu,
J. Z. Tsien, P. G. Schultz, M. D. Rose, J. L. Wood, D. O. Morgan, K. M. Shoka t, A chemical
switch for inhibitor-sensitive alleles of any protein kinase. Nature 407,395401 (2000).
39. We thank E. de Nadal for helpful discussions and support, M. Morillas and A. Vendrell
for strains, and M. Grötli (Sweden) for inhibitor design and supply. This work was
supported by an FPU fellowship to S.R.; grant CSD2007-0015 from Ministerio de Ciencia
y Tecnología, Consolider Ingenio 2010 program; through the European Commission
Directorate General Research, FP6 contract no. ERAS-CT-2003-980409 EURYI (European
Young Investigator Awards) award (www.esf.org/euryi) and QUASI to F.P; and FP6-
2005-NEST-PATH CELLCOMPUT project to F.P. and R.S. The FP laboratory also receives
support from the Fundación Marcelino Botín and Institució Catalana de Recerca i Estudis
Avançats (ICREA) Acadèmia (Generalitat de Catalunya).
Submitted 23 September 2008
Accepted 6 March 2009
Final Publication 24 March 2009
10.1126/scisignal.2000056
Citation: J. Macia, S. Regot, T. Peeters, N. Conde, R. Solé, F. Posas, Dynamic signaling
in the Hog1 MAPK pathway relies on high basal signal transduction. Sci. Signal. 2, ra13
(2009).
RESEARCH ARTICLE
www.SCIENCESIGNALING.org 24 March 2009 Vol 2 Issue 63 ra13 9
on May 17, 2013 stke.sciencemag.orgDownloaded from
... Another signaling branch is activated through the Synthetic, High Osmolarity-sensitive (Sho1) signaling protein osmosensor that activates the STErile Signal transducing MEK kinase (STE11) converging on Pbs2 (Maeda et al., 1994;Posas et al., 1996;Posas and Saito, 1997;Macia et al., 2009;Saito and Posas, 2012). Activated Pbs2 then dually phosphorylates the evolutionarily conserved MAPK Hog1 on threonine residue 174 (T174) and ...
... The multiple and presumably redundant MAPK phosphatases dephosphorylate and inactivate Hog1, which, along with the termination of upstream signaling after adaptation, results in its return to the cytosol. This knowledge of the Hog1 pathway was established through acute osmotic stress concentration increases that induce Hog1 phosphorylation, activation, and translocation to the nucleus ( Figures 3A, B) (Brewster et al., 1993;Ferrigno et al., 1998;Reiser et al., 1999;Hersen et al., 2008;Macia et al., 2009;Muzzey et al., 2009;Pelet et al., 2011;Saito and Posas, 2012;English et al., 2015;Mitchell et al., 2015;Granados et al., 2017). Activated Hog1 controls the regulation of cellular osmoadaptation and survival (Saito and Posas, 2012;Mitchell et al., 2015;Johnson et al., 2021). ...
... In detailed experiments, they showed that Hog1 nuclear localization depends on an osmolyte concentration and a rate threshold. Additional experiments showed that the Hog1 pathway uses an AND logic to integrate the previously determined concentration threshold (Macia et al., 2009) and the rate threshold. Both thresholds are required to activate Hog1 nuclear localization. ...
Article
Full-text available
All cells employ signal transduction pathways to respond to physiologically relevant extracellular cytokines, stressors, nutrient levels, hormones, morphogens, and other stimuli that vary in concentration and rate in healthy and diseased states. A central unsolved fundamental question in cell signaling is whether and how cells sense and integrate information conveyed by changes in the rate of extracellular stimuli concentrations, in addition to the absolute difference in concentration. We propose that different environmental changes over time influence cell behavior in addition to different signaling molecules or different genetic backgrounds. However, most current biomedical research focuses on acute environmental changes and does not consider how cells respond to environments that change slowly over time. As an example of such environmental change, we review cell sensitivity to environmental rate changes, including the novel mechanism of rate threshold. A rate threshold is defined as a threshold in the rate of change in the environment in which a rate value below the threshold does not activate signaling and a rate value above the threshold leads to signal activation. We reviewed p38/Hog1 osmotic stress signaling in yeast, chemotaxis and stress response in bacteria, cyclic adenosine monophosphate signaling in Amoebae, growth factors signaling in mammalian cells, morphogen dynamics during development, temporal dynamics of glucose and insulin signaling, and spatio-temproral stressors in the kidney. These reviewed examples from the literature indicate that rate thresholds are widespread and an underappreciated fundamental property of cell signaling. Finally, by studying cells in non-linear environments, we outline future directions to understand cell physiology better in normal and pathophysiological conditions.
... In S. cerevisiae, the HOG pathway is proceeded by consecutive phosphorylation with two upstream branches, Sln1 and Sho1 (O' Rourke and Herskowitz, 2004). The Sln1 branch is controlled by a two-component signaling pathway composed of a dimeric sensor kinase (Sln1), the histidinecontaining phosphotransfer protein (Ypd1), and a response regulator (Ssk1) (Macia et al., 2009). Sln1 is a membranelocalized histidine kinase and acts as a negative regulator of the HOG pathway (Hohmann, 2002). ...
Article
During the infection process, plant pathogenic fungi encounter plant-derived oxidative stress, and an appropriate response to this stress is crucial to their survival and establishment of the disease. Plant pathogenic fungi have evolved several mechanisms to eliminate oxidants from the external environment and maintain cellular redox homeostasis. When oxidative stress is perceived, various signaling transduction pathways are triggered and activate the downstream genes responsible for the oxidative stress response. Despite extensive research on antioxidant systems and their regulatory mechanisms in plant pathogenic fungi, the specific functions of individual antioxidants and their impacts on pathogenicity have not recently been systematically summarized. Therefore, our objective is to consolidate previous research on the antioxidant systems of plant pathogenic fungi. In this review, we explore the plant immune responses during fungal infection, with a focus on the generation and function of reactive oxygen species. Furthermore, we delve into the three antioxidant systems, summarizing their functions and regulatory mechanisms involved in oxidative stress response. This comprehensive review provides an integrated overview of the antioxidant mechanisms within plant pathogenic fungi, revealing how the oxidative stress response contributes to their pathogenicity.
... This points to an important role of yl-Hog1 basal activity. The basal activity of MAPK cascades means that a small fraction of these proteins is constantly active to maintain a balance between different signaling pathways, but also to accelerate the activation of the stress response when needed [25]. The fact that both strains hog1∆ HOG1-49 and hog1∆ HOG1-171/173 show a very similar form of growth to hog1∆, enables two conclusions to be drawn about the yl-Hog1 basal activity. ...
Article
Full-text available
Background: Yarrowia lipolytica is a dimorphic fungus, which switches from yeast to filament form in response to environmental conditions. For industrial purposes it is important to lock cells in the yeast or filamentous form depending on the fermentation process. yl-Hog1 kinase is a key component of the HOG signaling pathway, responsible for activating the osmotic stress response. Additionally, deletion of yl-Hog1 leads to increased filamentation in Yarrowia lipolytica, but causes significant sensitivity to osmotic stress induced by a high concentration of a carbon source. Results: In this study, we tested the effect of point mutations on the function of yl-Hog1 protein kinase. The targets of modification were the phosphorylation sites (T171A-Y173A) and the active center (K49R). Introduction of the variant HOG1-49 into the hog1∆ strain partially improved growth under osmotic stress, but did not recover the yeast-like shape of the cells. The HOG1-171/173 variant was not functional, and its introduction further weakened the growth of hog1∆ strains in hyperosmotic conditions. To verify a genetic modification in filament form, we developed a new system based on green fluorescent protein (GFP) for easier screening of proper mutants. Conclusions: These results provide new insights into the functions of yl-Hog1 protein in dimorphic transition and constitute a good starting point for further genetic modification of Y. lipolytica in filament form.
... Changes in expression are affected greatly by external signaling pathways and transcription factors, which can modulate expression for a short or long term [96]. The large shifts in gene expression with regard to 30-mM-adapted cells were unexpected; however, when comparing unadapted to 60-mM-adapted cells, some of the expression changes seen in 30-mM-adapted cells are likely attributed to gene expression tuning in response to the lactate stress [97]. Responding to a temporary stress can cause shifts in complex gene regulation pathways; adaptation to chronic stresses can cause the drastic rebalancing of regulatory cascades [96]. ...
Article
Full-text available
The accumulation of metabolic wastes in cell cultures can diminish product quality, reduce productivity, and trigger apoptosis. The limitation or removal of unintended waste products from Chinese hamster ovary (CHO) cell cultures has been attempted through multiple process and genetic engineering avenues with varied levels of success. One study demonstrated a simple method to reduce lactate and ammonia production in CHO cells with adaptation to extracellular lactate; however, the mechanism behind adaptation was not certain. To address this profound gap, this study characterizes the phenotype of a recombinant CHO K-1 cell line that was gradually adapted to moderate and high levels of extracellular lactate and examines the genomic content and role of extrachromosomal circular DNA (eccDNA) and gene expression on the adaptation process. More than 500 genes were observed on eccDNAs. Notably, more than 1000 genes were observed to be differentially expressed at different levels of lactate adaptation, while only 137 genes were found to be differentially expressed between unadapted cells and cells adapted to grow in high levels of lactate; this suggests stochastic switching as a potential stress adaptation mechanism in CHO cells. Further, these data suggest alanine biosynthesis as a potential stress-mitigation mechanism for excess lactate in CHO cells.
... This channel remains closed upon hyperosmotic conditions preventing glycerol from exiting the cell [46,47]. The other mechanism involves the bona fide sensors, Sln1p and Sho1p, that control the HOG-MAPK (High Osmolarity Glycerol-Mitogen Activated Protein Kinase) pathway [48,49]. The MAPK of this pathway, Hog1, acts on cytoplasmic and nuclear targets to modify cellular metabolism to increase glycerol synthesis [50][51][52]. ...
Preprint
Full-text available
The yeast Saccharomyces cerevisiae is widely used in food and nonfood industries. During industrial fermentations yeast are exposed to fluctuations in oxygen concentration, osmotic pressure, pH, ethanol concentration, nutrient availability and temperature. Fermentation performance depends on the ability of different yeast strains to adapt to these changes. Suboptimal growth conditions trigger responses to these external stimuli to allow cellular homeostasis to be maintained. Stress-specific signaling pathways are activated to coordinate changes in tran-scription, translation, protein function, and metabolic fluxes while a transient arrest of growth and cell cycle progression occur. cAMP-PKA, HOG-MAPK and CWI signalling pathways are signal transduction pathways turned on during stress re-sponse. Comprehension of the mechanisms involved in the responses and in the adaptation to these stresses during fermentation is key to improving this industrial process. The scope of this review is to outline the advancement of knowledge about the cAMP-PKA signalling and the crosstalk of this pathway with the CWI and HOG-MAPK cascades in response to the environmental challenges heat and hy-perosmotic stress.
... This points to an important role of yl-Hog1 basal activity. The basal activity of MAPK cascades means that a small fraction of these proteins is constantly active to maintain a balance between different signaling pathways, but also to accelerate the activation of the stress response when needed [24]. The fact that both strains hog1∆ HOG1-49 and hog1∆ HOG1-171/173 show a very similar form of growth to hog1∆, enables two conclusions to be drawn about the yl-Hog1 basal activity. ...
Preprint
Full-text available
Background Yarrowia lipolytica is a dimorphic fungus, which switches from yeast to yeast-to-filament form in response to environmental conditions. For industrial purposes it is important to lock cells in the yeast or filamentous form depending on the fermentation process. yl-Hog1 kinase is a key component of the HOG signaling pathway, responsible for activating the osmotic stress response. Additionally, deletion of yl-Hog1 leads to increased filamentation in Yarrowia lipolytica, but causes significant sensitivity to osmotic stress induced by a high concentration of a carbon source. Results In this study, we tested the effect of point mutations on the function of yl-Hog1 protein kinase. The targets of modification were the phosphorylation sites (T171A-Y173A) and the active center (K49R). Introduction of the variant HOG1-49 into the hog1∆ strain partially improved growth under osmotic stress, but did not recover the yeast-like shape of the cells. The HOG1-171/173 variant was completely inactive, and its introduction further weakened the hog1∆ strains. To verify a genetic modification in filament form, we developed a new system based on green fluorescent protein (GFP) for easier screening of proper mutants. Conclusions These results provide new insights into the functions of yl-Hog1 protein in dimorphic transition and constitute a good starting point for further genetic modification of Y. lipolytica in filament form.
... For example, mass spectrometry measurements inform about relative changes in double phosphorylated Hog1 within the first 60 s of the signaling response as well as at later time points 29,30 . These data were complemented by western blot measurements with antibodies recognizing doubly phosphorylated Hog1 with conditions including strains lacking different upstream components or the Ptp2 and Ptp3 phosphatases, as well as inhibition of Hog1 activity by small molecule inhibitors 20,26,36 . Moreover, Hog1 activity correlates with its nuclear translocation, which can be quantified by fluorescence microscopy in single cells 27 . ...
Article
Full-text available
Cellular decision making often builds on ultrasensitive MAPK pathways. The phosphorylation mechanism of MAP kinase has so far been described as either distributive or processive, with distributive mechanisms generating ultrasensitivity in theoretical analyses. However, the in vivo mechanism of MAP kinase phosphorylation and its activation dynamics remain unclear. Here, we characterize the regulation of the MAP kinase Hog1 in Saccharomyces cerevisiae via topologically different ODE models, parameterized on multimodal activation data. Interestingly, our best fitting model switches between distributive and processive phosphorylation behavior regulated via a positive feedback loop composed of an affinity and a catalytic component targeting the MAP kinase-kinase Pbs2. Indeed, we show that Hog1 directly phosphorylates Pbs2 on serine 248 (S248), that cells expressing a non-phosphorylatable (S248A) or phosphomimetic (S248E) mutant show behavior that is consistent with simulations of disrupted or constitutively active affinity feedback and that Pbs2-S248E shows significantly increased affinity to Hog1 in vitro. Simulations further suggest that this mixed Hog1 activation mechanism is required for full sensitivity to stimuli and to ensure robustness to different perturbations.
Article
Full-text available
The yeast Saccharomyces cerevisiae is widely used in food and non-food industries. During industrial fermentation yeast strains are exposed to fluctuations in oxygen concentration, osmot-ic pressure, pH, ethanol concentration, nutrient availability and temperature. Fermentation performance depends on the ability of the yeast strains to adapt to these changes. Suboptimal conditions trigger responses to the external stimuli to allow homeostasis to be maintained. Stress-specific signalling pathways are activated to co-ordinate changes in transcription, translation, protein function, and metabolic fluxes while a transient arrest of growth and cell cycle progression occur. cAMP-PKA, HOG-MAPK and CWI signalling path-ways are turned on during stress response. Comprehension of the mechanisms involved in the responses and in the adaptation to these stresses during fermentation is key to improving this industri-al process. The scope of this review is to outline the advancement of knowledge about the cAMP-PKA signalling and the crosstalk of this pathway with the CWI and HOG-MAPK cascades in response to the environmental challenges heat and hyperosmotic stress.
Preprint
Full-text available
How cells respond to dynamic environmental changes is crucial for understanding fundamental biological processes and cell physiology. In this study, we developed an experimental and quantitative analytical framework to explore how dynamic stress gradients that change over time regulate cellular volume, signaling activation, and growth phenotypes. Our findings reveal that gradual stress conditions substantially enhance cell growth compared to conventional acute stress. This growth advantage correlates with a minimal reduction in cell volume dependent on the dynamic of stress. We explain the growth phenotype with our finding of a logarithmic signal transduction mechanism in the yeast Mitogen-Activated Protein Kinase (MAPK) osmotic stress response pathway. These insights into the interplay between gradual environments, cell volume change, dynamic cell signaling, and growth, advance our understanding of fundamental cellular processes in gradual stress environments.
Article
To cope with an increased external osmolarity, the budding yeast Saccharomyces cerevisiae activates the Hog1 mitogen-activated kinase (MAPK) through the High-Osmolarity Glycerol (HOG) pathway, which governs adaptive responses to osmostress. In the HOG pathway, two apparently redundant upstream branches, termed SLN1 and SHO1, activate cognate MAP3Ks Ssk2/22 and Ste11, respectively. These MAP3Ks, when activated, phosphorylate and thus activate the Pbs2 MAP2K, which in turn phosphorylates and activates Hog1. Previous studies have shown that protein tyrosine phosphatases (PTP) and the serine/threonine protein phosphatases type 2C (PP2C) negatively regulate the HOG pathway to prevent its excessive and inappropriate activation, which is detrimental to cell growth. The tyrosine phosphatases Ptp2 and Ptp3 dephosphorylate Hog1 at Tyr-176, whereas the PP2Cs Ptc1 and Ptc2 dephosphorylate Hog1 at Thr-174. In contrast, the identities of phosphatases that dephosphorylate Pbs2 remained less clear. Here, we examined the phosphorylation status of Pbs2 at the activating phosphorylation sites Ser-514 and Thr-518 (S514 and T518) in various mutants, both in the unstimulated and osmostressed conditions. Thus, we found that Ptc1-Ptc4 collectively regulate Pbs2 negatively, but each Ptc acts differently to the two phosphorylation sites in Pbs2. T518 is predominantly dephosphorylated by Ptc1, whereas the effect of Ptc2-Ptc4 could be seen only when Ptc1 is absent. Conversely, S514 can be dephosphorylated by any of Ptc1-4 to an appreciable extent. We also show that Pbs2 dephosphorylation by Ptc1 requires the adaptor protein Nbp2 that recruits Ptc1 to Pbs2, thus highlighting the complex processes involved in regulating adaptive responses to osmostress.
Article
Full-text available
Haploid Saccharomyces cerevisiae yeast cells use a prototypic cell signalling system to transmit information about the extracellular concentration of mating pheromone secreted by potential mating partners. The ability of cells to respond distinguishably to different pheromone concentrations depends on how much information about pheromone concentration the system can transmit. Here we show that the mitogen-activated protein kinase Fus3 mediates fast-acting negative feedback that adjusts the dose response of the downstream system response to match the dose response of receptor-ligand binding. This 'dose-response alignment', defined by a linear relationship between receptor occupancy and downstream response, can improve the fidelity of information transmission by making downstream responses corresponding to different receptor occupancies more distinguishable and reducing amplification of stochastic noise during signal transmission. We also show that one target of the feedback is a previously uncharacterized signal-promoting function of the regulator of G-protein signalling protein Sst2. Our work suggests that negative feedback is a general mechanism used in signalling systems to align dose responses and thereby increase the fidelity of information transmission.
Article
Full-text available
Protein phosphatases inactivate mitogen-activated protein kinase (MAPK) signaling pathways by dephosphorylating components of the MAPK cascade. Two genes encoding protein-tyrosine phosphatases, PTP2, and a new phosphatase, PTP3, have been isolated in a genetic selection for negative regulators of an osmotic stress response pathway called HOG, for high osmolarity glycerol, in budding yeast. PTP2 and PTP3 were isolated as multicopy suppressors of a severe growth defect due to hyperactivation of the HOG pathway. Phosphatase activity is required for suppression since mutation of the catalytic Cys residue in Ptp2 and Ptp3, destroys suppressor function and biochemical activity. The substrate of these phosphatases is likely to be the MAPK, Hog1. Catalytically inactive Ptp2 and Ptp3 coprecipitate with Hog1 from yeast extracts. In addition, strains lacking PTP2 and PTP3 do not dephosphorylate Hog1-phosphotyrosine as well as wild type. The latter suggests that PTP2 and PTP3 play a role in adaptation. Consistent with this role, osmotic stress induces expression of PTP2 and PTP3 transcripts in a Hog1-dependent manner. Thus Ptp2 and Ptp3 likely act in a negative feedback loop to inactivate Hog1.
Article
Czeisler and his colleagues have recently reported that bright light can induce strong (Type 0) resetting of the human circadian pacemaker. This surprising result shows that the human clock is more responsive to light than has been previously thought. The interpretation of their results is subtle, however, because of an unconventional aspect of their experimental protocol: They measured the phase shift after three, cycles of the bright-light stimulus, rather than after the usual single pulse. A natural question is whether the apparent Type 0 response could reflect the summation of three weaker Type 1 responses to each of the daily light pulses. In this paper I show mathematically that repeated Type 1 resetting cannot account for the observed Type 0 response. This finding corroborates the strong resetting reported by Czeisler et al., and supports their claim that bright light induces strong resetting by crushing the amplitude of the circadian pacemaker. Furthermore, the results indicate that back-to-back light pulses can have a cooperative effect different from that obtained by simple iteration of a phase response curve (PRC). In this sense the resetting response of humans is similar to that of Drosophila, Kalanchoe, and Culex, and is more complex than that predicted by conventional PRC theory. To describe the way in which light resets the human circadian pacemaker, one needs a theory that includes amplitude resetting, as pioneered by Winfree and developed for humans by Kronauer.
Article
The role of mitogen-activated protein (MAP) kinase cascades in integrating distinct upstream signals was studied in yeast. Mutants that were not able to activate PBS2 MAP kinase kinase (MAPKK; Pbs2p) at high osmolarity were characterized. Pbs2p was activated by two independent signals that emanated from distinct cell-surface osmosensors. Pbs2p was activated by MAP kinase kinase kinases (MAPKKKs) Ssk2p and Ssk22p that are under the control of the SLN1-SSK1 two-component osmosensor. Alternatively, Pbs2p was activated by a mechanism that involves the binding of its amino terminal proline-rich motif to the Src homology 3 (SH3) domain of a putative transmembrane osmosensor Sho1p.
Article
An osmosensing mechanism in the budding yeast (Saccharomyces cerevisiae) involves both a two-component signal transducer (Sln1p, Ypd1p and Ssk1p) and a MAP kinase cascade (Ssk2p/Ssk22p, Pbs2p, and Hog1p). The transmembrane protein Sln1p contains an extracellular sensor domain and cytoplasmic histidine kinase and receiver domains, whereas the cytoplasmic protein Ssk1p contains a receiver domain. Ypd1p binds to both Sln1p and Ssk1p and mediates the multistep phosphotransfer reaction (phosphorelay). This phosphorelay system is initiated by the autophosphorylation of Sln1p at His576. This phosphate is then sequentially transferred to Sln1p-Asp-1144, then to Ypd1p-His64, and finally to Ssk1p-Asp554. We propose that the multistep phosphorelay mechanism is a universal signal transduction apparatus utilized both in prokaryotes and eukaryotes.
Article
In response to increases in extracellular osmolarity, Saccharomyces cerevisiae activates the HOG1 mitogen-activated protein kinase (MAPK) cascade, which is composed of a pair of redundant MAPK kinase kinases, namely, Ssk2p and Ssk22p, the MAPK kinase Pbs2p, and the MAPK Hog1p. Hog1p is activated by Pbs2p through phosphorylation of specific threonine and tyrosine residues. Activated Hog1p is essential for survival of yeast cells at high osmolarity. However, expression of constitutively active mutant kinases, such as those encoded by SSK2deltaN and PBS2(DD), is toxic and results in a lethal level of Hog1p activation. Overexpression of the protein tyrosine phosphatase Ptp2p suppresses the lethality of these mutations by dephosphorylating Hog1p. A catalytically inactive Cys-to-Ser Ptp2p mutant (Ptp2(C/S)p) is tightly bound to tyrosine-phosphorylated Hog1p in vivo. Disruption of PTP2 leads to elevated levels of tyrosine-phosphorylated Hog1p following exposure of cells to high osmolarity. Disruption of both PTP2 and another protein tyrosine phosphatase gene, PTP3, results in constitutive Hog1p tyrosine phosphorylation even in the absence of increased osmolarity. Thus, Ptp2p and Ptp3p are the major phosphatases responsible for the tyrosine dephosphorylation of Hog1p. When catalytically inactive Hog1(K/N)p is expressed in hog1delta cells, it is constitutively tyrosine phosphorylated. In contrast, Hog1(K/N)p, expressed together with wild-type Hog1p, is tyrosine phosphorylated only when cells are exposed to high osmolarity. Thus, the kinase activity of Hog1p is required for its own tyrosine dephosphorylation. Northern blot analyses suggest that Hog1p regulates Ptp2p and/or Ptp3p activity at the posttranscriptional level.
Article
Exposure of the yeast Saccharomyces cerevisiae to high extracellular osmolarity induces the Sln1p-Ypd1p-Ssk1p two-component osmosensor to activate a mitogen-activated protein (MAP) kinase cascade composed of the Ssk2p and Ssk22p MAP kinase kinase kinases (MAPKKKs), the Pbs2p MAPKK, and the Hog1p MAPK. A second osmosensor, Sho1p, also activated Pbs2p and Hog1p, but did so through the Ste11p MAPKKK. Although Ste11p also participates in the mating pheromone-responsive MAPK cascade, there was no detectable cross talk between these two pathways. The MAPKK Pbs2p bound to the Sho1p osmosensor, the MAPKKK Ste11p, and the MAPK Hog1p. Thus, Pbs2p may serve as a scaffold protein.
Article
Exposure of yeast cells to increased extracellular osmolarity induces the HOG1 mitogen-activated protein kinase (MAPK) cascade, which is composed of SSK2, SSK22 and STE11 MAPKKKs, PBS2 MAPKK and HOG1 MAPK. The SSK2/SSK22 MAPKKKs are activated by a 'two-component' osmosensor composed of SLN1, YPD1 and SSK1. The SSK1 C-terminal receiver domain interacts with an N-terminal segment of SSK2. Upon hyperosmotic treatment, SSK2 is autophosphorylated rapidly, and this reaction requires the interaction of SSK1 with SSK2. Autophosphorylation of SSK2 is an intramolecular reaction, suggesting similarity to the mammalian MEKK1 kinase. Dephosphorylation of SSK2 renders the kinase inactive, but it can be re-activated by addition of SSK1 in vitro. A conserved threonine residue (Thr1460) in the activation loop of SSK2 is important for kinase activity. Based on these observations, we propose the following two-step activation mechanism of SSK2 MAPKKK. In the first step, the binding of SSK1 to the SSK1-binding site in the N-terminal domain of SSK2 causes a conformational change in SSK2 and induces its latent kinase activity. In the second step, autophosphorylation of SSK2 renders its activity independent of the presence of SSK1. A similar mechanism might be applicable to other MAPKKKs from both yeast and higher eukaryotes.
Article
MAPK (mitogen-activated protein kinase) cascades are common eukaryotic signaling modules that consist of a MAPK, a MAPK kinase (MAPKK) and a MAPKK kinase (MAPKKK). Because phosphorylation is essential for the activation of both MAPKKs and MAPKs, protein phosphatases are likely to be important regulators of signaling through MAPK cascades. To identify protein phosphatases that negatively regulate the stress-responsive p38 and JNK MAPK cascades, we screened human cDNA libraries for genes that down-regulated the yeast HOG1 MAPK pathway, which shares similarities with the p38 and JNK pathways, using a hyperactivating yeast mutant. In this screen, the human protein phosphatase type 2Calpha (PP2Calpha) was found to negatively regulate the HOG1 pathway in yeast. Moreover, when expressed in mammalian cells, PP2Calpha inhibited the activation of the p38 and JNK cascades induced by environmental stresses. Both in vivo and in vitro observations indicated that PP2Calpha dephosphorylated and inactivated MAPKKs (MKK6 and SEK1) and a MAPK (p38) in the stress-responsive MAPK cascades. Furthermore, a direct interaction of PP2Calpha and p38 was demonstrated by a co-immunoprecipitation assay. This interaction was observed only when cells were stimulated with stresses or when a catalytically inactive PP2Calpha mutant was used, suggesting that only the phosphorylated form of p38 interacts with PP2Calpha.