ArticlePDF Available

Lee FS, Chao MV. Activation of Trk neurotrophin receptors in the absence of neurotrophins. Proc Natl Acad Sci USA 98: 3555-3560

Authors:

Abstract and Figures

Neurotrophins regulate neuronal cell survival and synaptic plasticity through activation of Trk receptor tyrosine kinases. Binding of neurotrophins to Trk receptors results in receptor autophosphorylation and downstream phosphorylation cascades. Here, we describe an approach to use small molecule agonists to transactivate Trk neurotrophin receptors. Activation of TrkA receptors in PC12 cells and TrkB in hippocampal neurons was observed after treatment with adenosine, a neuromodulator that acts through G protein-coupled receptors. These effects were reproduced by using the adenosine agonist CGS 21680 and were counteracted with the antagonist ZM 241385, indicating that this transactivation event by adenosine involves adenosine 2A receptors. The increase in Trk activity could be inhibited by the use of the Src family-specific inhibitor, PP1, or K252a, an inhibitor of Trk receptors. In contrast to other G protein-coupled receptor transactivation events, adenosine used Trk receptor signaling with a longer time course. Moreover, adenosine activated phosphatidylinositol 3-kinase/Akt through a Trk-dependent mechanism that resulted in increased cell survival after nerve growth factor or brain-derived neurotrophic factor withdrawal. Therefore, adenosine acting through the A(2A) receptors exerts a trophic effect through the engagement of Trk receptors. These results provide an explanation for neuroprotective actions of adenosine through a unique signaling mechanism and raise the possibility that small molecules may be used to elicit neurotrophic effects for the treatment of neurodegenerative diseases.
Content may be subject to copyright.
Activation of Trk neurotrophin receptors in the
absence of neurotrophins
Francis S. Lee* and Moses V. Chao
†‡
*Department of Psychiatry, Weill Medical College of Cornell University Medical College, New York, NY 10021; and Molecular Neurobiology Program,
Skirball Institute of Biomolecular Medicine, Departments of Cell Biology and Physiology and Neuroscience, New York University School of Medicine,
New York, NY 10016
Communicated by Hans Thoenen, Max Planck Institute of Neurobiology, Martinsried, Germany, January 11, 2001 (received for review November 2, 2000)
Neurotrophins regulate neuronal cell survival and synaptic plas-
ticity through activation of Trk receptor tyrosine kinases. Binding
of neurotrophins to Trk receptors results in receptor autophos-
phorylation and downstream phosphorylation cascades. Here, we
describe an approach to use small molecule agonists to transacti-
vate Trk neurotrophin receptors. Activation of TrkA receptors in
PC12 cells and TrkB in hippocampal neurons was observed after
treatment with adenosine, a neuromodulator that acts through G
protein-coupled receptors. These effects were reproduced by using
the adenosine agonist CGS 21680 and were counteracted with the
antagonist ZM 241385, indicating that this transactivation event by
adenosine involves adenosine 2A receptors. The increase in Trk
activity could be inhibited by the use of the Src family-specific
inhibitor, PP1, or K252a, an inhibitor of Trk receptors. In contrast to
other G protein-coupled receptor transactivation events, adeno-
sine used Trk receptor signaling with a longer time course. More-
over, adenosine activated phosphatidylinositol 3-kinaseAkt
through a Trk-dependent mechanism that resulted in increased cell
survival after nerve growth factor or brain-derived neurotrophic
factor withdrawal. Therefore, adenosine acting through the A
2A
receptors exerts a trophic effect through the engagement of
Trk receptors. These results provide an explanation for neuropro-
tective actions of adenosine through a unique signaling mecha-
nism and raise the possibility that small molecules may be used to
elicit neurotrophic effects for the treatment of neurodegenerative
diseases.
Neurotrophins play a prominent role in the development of
the vertebrate nervous system by inf luencing cell survival,
differentiation, and cell death events (1, 2). Neurotrophins also
exhibit acute regulatory effects on neurotransmitter release,
synaptic strength, and connectivity (3, 4). In addition to pro-
moting axonal and dendritic branching, neurotrophins serve as
chemoattractants for extending growth cones in vitro (5). These
actions are mediated by neurotrophin binding to two separate
receptor classes, the Trk family of tyrosine kinase receptors and
the p75 neurotrophin receptor, a member of the tumor necrosis
factor receptor superfamily (6).
Mutations in Trk neurotrophin receptor function lead to
deficits in survival, axonal and dendritic branching, long-term
potentiation, and behavior (7–9). Nerve growth factor (NGF),
brain-derived neurotrophic factor (BDNF), neurotrophin-3, and
neurotrophin-4 also bind to the p75 neurotrophin receptor, a
potential cell death receptor whose actions are negated by Trk
tyrosine kinase signaling (10, 11). Therefore, the ability to
regulate Trk tyrosine kinase activity is critical for neuronal
survival and differentiation.
Ligands for G protein-coupled receptors are capable of acti-
vating the mitogen-activated protein (MAP) kinase signaling
pathway, in addition to classic G protein-dependent signaling
pathways involving adenylyl cyclase and phospholipase C (12,
13). Induction of mitogenic receptor tyrosine kinase phosphor-
ylation also occurs through signaling from several G protein-
coupled receptors (14). In particular, receptors for epidermal
growth factor, platelet-derived growth factor, and insulin-like
growth factor 1 can be transactivated by G protein-coupled
receptors(12, 15, 16). Whether transactivation of neurotrophic
receptor tyrosine kinases occurs by means of G protein-coupled
receptors has not been demonstrated to date.
We have tested the possibility that ligands of G protein-
coupled receptors might activate neurotrophin receptors of the
Trk tyrosine kinase subfamily. Here, we report that adenosine
and adenosine agonists can activate Trk receptor phosphoryla-
tion, through a mechanism that requires the adenosine 2A (A
2A
)
receptor. The activation does not require neurotrophin binding
and is observed in PC12 cells, as well as primary cultures of
hippocampal neurons. Unlike the results obtained with other
tyrosine kinase receptors, increased Trk receptor activity pro-
vides increased cell survival over a prolonged time course that
requires Akt, and not MAP kinase, signaling. These findings
suggest alternative approaches of stimulating trophic functions
in the nervous system by linking different receptor signaling
pathways.
Materials and Methods
CGS 21680, CPA, A23187, and insulin-like growth factor-1 were
purchased from Sigma-RBI. ZM 241385 was from Tocris Neu-
rochemicals (Ballwin, MO), PP1 from Alexis Biochemicals (San
Diego, CA), LY294002 from Biomol, K252a from Calbiochem,
and PD98059 from New England Biolabs. NGF was obtained
from Harlan Bioproducts (Indianapolis, IN) and BDNF from
PeproTech (Rocky Hill, NJ). All other compounds were from
Sigma. An anti-pan-Trk rabbit antiserum raised against the
C-terminal region of the Trk receptor was from Barbara Hemp-
stead (Cornell University); anti-NGF antibody was obtained
from Chemicon. Anti-phosphotyrosine and anti-Akt antibodies
were from Santa Cruz Biotechnologies. Anti-phospho-Akt, anti-
MAP kinase, and anti-phospho-MAP kinase antibodies were
from New England Biolabs.
Immunoprecipitation and Immunoblotting. PC12 cells or PC12
(615) cells (17), were maintained in DMEM containing 10%
FBS supplemented with 100 unitsml penicillin, 100
gml
streptomycin, and 2 mM glutamine plus 200
gml G418. Cells
were placed in low-serum medium (1% FBS, 0.5% horse serum)
overnight before experiments. Cell lysates from PC12, 615 cells,
or hippocampal cells were incubated in lysis buffer (1% Nonidet
P-40) for4htoovernight at 4°C with anti-pan-Trk polyclonal
antibody followed by incubation with protein A-Sepharose
beads. Equivalent amounts of protein were analyzed for each
condition. The beads were washed five times with lysis buffer,
and the immune complexes were boiled in SDS-sample buffer
Abbreviations: NGF, nerve growth factor; BDNF, brain-derived neurotrophic factor; A2A,
adenosine 2A; PI3-kinase, phosphatidylinositol 3-kinase; MAP, mitogen-activated protein;
LDH, lactate dehydrogenase; En, embryonic day n.
To whom reprint requests should be addressed. E-mail: chao@saturn.med.nyu.edu.
The publication costs of this article were defrayed in part by page charge payment. This
article must therefore be hereby marked advertisement in accordance with 18 U.S.C.
§1734 solely to indicate this fact.
www.pnas.orgcgidoi10.1073pnas.061020198 PNAS
March 13, 2001
vol. 98
no. 6
3555–3560
NEUROBIOLOGY
and loaded on SDS-PAGE gels for immunoblot analysis. The
immunoreactive protein bands were detected by enhanced
chemiluminescence (Amersham Pharmacia).
125
I-NGF Binding Analysis. For equilibrium binding studies,
125
I-
NGF was prepared as described previously (18). PC12 cells
stably overexpressing TrkA (2 10
5
cells) and HEK 293 cells
expressing TrkA (2 10
5
) were incubated with
125
I-NGF in the
absence and presence of adenosine compounds for 30 min at
25°C. The cells were then washed twice with PBS, and
125
I-NGF
was stripped with an acid solution (0.2 M acetic acid, 0.5 M
NaCl). Nonspecific binding was assessed by adding unlabeled
NGF at a final concentration of 1000 ngml and represented
20% of total binding. Specific binding was defined as total
binding minus nonspecific binding. All conditions were carried
out in triplicate and SEM calculated.
Hippocampal Cell Cultures. Dissociated primary cultures of hip-
pocampal neurons from embryonic day 17 (E17) rats were
prepared from timed-pregnant Sprague–Dawley rats as de-
scribed previously (19). Fetuses were removed under sterile
conditions and kept in PBS on ice for microscopic dissection of
the hippocampus. The meninges were removed and the tissue
was placed in Neurobasal media (GIBCOBRL). The tissue was
briefly minced with fine forceps and then triturated with a
fire-polished pasteur pipet. Cells were counted and plated on
culture wells coated with 0.01 mgml poly-D-lysine overnight.
Hippocampal cells were maintained in Neurobasal media, con-
taining B27 supplement and L-glutamine (0.5 mM). Experiments
were conducted 7–10 days after plating.
Cell Death Assay. PC12 cells were differentiated in DMEM
supplemented with 0.33% FBS, 0.67% heat-inactivated horse
serum, 2 mM L-glutamine, and NGF (50 ngml) for 7 days.
Serum and NGF were then removed, and adenosine agonists or
growth factors were added to the media. After 48 h, cell death
was quantified by measuring lactate dehydrogenase (LDH)
released from injured cells into the media by using the Cytox 96
Cytotoxicity Assay Kit (Promega). LDH values were normalized
by subtracting the LDH released by cells maintained in NGF (50
ngml) and scaling to a full kill (100%) reference induced by
5 min of treatment with 1% Triton, an exposure that consistently
killed all PC12 cells.
Hippocampal neurons were maintained in Neurobasal media
containing B27 supplement and 0.5 mM L-glutamine for 10 days.
B27 was then removed, and adenosine agonists or BDNF (100
ngml) were added to the media. MK-801 (1
M) was added to
all conditions to decrease the contribution of N-methyl-D-
aspartate-mediated cell death. After 48 h, cell death was assessed
by measurement of LDH released into media. LDH values were
normalized by subtracting the LDH released by cells maintained
in BDNF (100 ngml) and scaling to full kill (100%) reference
induced by 24 h of treatment with A23187 (30
M), a condition
that resulted in complete cell death of all neurons (20).
Results
Transactivation of mitogenic tyrosine kinase receptors through
G protein-coupled receptors has been described previously (12,
15, 16). To explore whether any G protein-coupled receptors
exert an effect on neurotrophin receptor signaling, several
ligands were tested for their ability to influence TrkA tyrosine
kinase activity in PC12 cells. Receptors for each ligand are found
on PC12 cells (21–24). TrkA receptors were immunoprecipitated
from PC12 cell lysates and then probed with an anti-
phosphotyrosine antibody. Activated TrkA receptors were ob-
served with adenosine treatment (10
M), but not with nucle-
otides such as ATP or GTP (Fig. 1). The TrkA doublet
represents an underglycosylated form of 110 kDa and the fully
glycosylated form of 140 kDa (17). Activation of TrkA receptors
was not observed with other G protein-coupled ligands, includ-
ing bradykinin and dopamine agonists, apomorphine and quin-
pirole (Fig. 1). The specificity of adenosine’s effects was also
confirmed by the use of CGS 21680, 2-[(4-(2-carboxyethyl)phe-
nylethyl)] aminoadenosine-5-N-ethylcarboxamide, a selective
adenosine A
2A
agonist (25).
The effect of adenosine on TrkA receptor activity occurred in
a low concentration range (Fig. 2A). This response was verified
by the use of 1 nM CGS 21680. A time course of adenosine action
showed that the increase in TrkA activation was slow and
required at least 90 min (Fig. 2B), which is delayed compared
with NGF treatment. This increase was inhibited by K252a, a
known inhibitor of Trk tyrosine kinases (see below). It is
formally possible that adenosine treatment leads to the produc-
tion of NGF by PC12 cells that could act in an autocrine fashion
to stimulate TrkA receptors. This possibility was discounted by
the absence of neurite outgrowth activity of supernatants taken
from PC12 cells treated with adenosine and by a lack of effect
of anti-NGF antibody on adenosine’s action (data not shown).
Adenosine interacts with four different G protein-coupled
receptors, designated A
1
,A
2A
,A
2B
, and A
3
receptors (26). The
A
2
class of adenosine receptors are expressed in PC12 cells and
have been detected by radioligand binding (23). Adenosine does
not bind to the TrkA receptor. There was no displacement of
125
I-NGF binding with an excess of adenosine (1 mM) or CGS
21680 (1
M) in PC12 cells overexpressing TrkA (Table 1). As
PC12 cells express the p75 neurotrophin receptor, which also
binds
125
I-NGF, similar experiments were carried out in HEK
293 cells after transfection with TrkA. Again, excess concentra-
tions of adenosine or CGS 21680 did not displace
125
I-NGF
binding to HEK 293 cells that expressed TrkA (Table 1). The
concentrations of adenosine and CGS 21680 were approximately
100-fold greater than those normally used in A
2A
receptor
binding and signaling studies (27).
To verify adenosine interacted specifically with the A
2A
re-
ceptor, several adenosine analogs were used. A low concentra-
tion (10 nM) of CGS 21680 gave a similar increase in phosphor-
ylated TrkA receptors with the same time course as adenosine
(Figs. 2Aand 3A). In contrast, a selective A
1
receptor agonist,
CPA, N(6)-cyclopentyladenosine, had no effect (Fig. 3A). Incu-
Fig. 1. Activation of TrkA receptors with G protein-coupled receptor ligands.
Stably transfected PC12 cells expressing high levels of TrkA (615) were treated
with the indicated compounds for 2 h. The cells were subsequently harvested
in lysis buffer as described in Materials and Methods. Lysates were immuno-
precipitated with anti-pan Trk rabbit antiserum. Immunocomplexes were
analyzed by immunoblotting with anti-phosphotyrosine antibody (PY99).
Immunoprecipitation of TrkA receptors was then confirmed by immunoblot-
ting of the immunocomplex with anti-pan Trk antiserum.
3556
www.pnas.orgcgidoi10.1073pnas.061020198 Lee and Chao
bation of PC12 cells with the A
2A
antagonist, ZM 241385, 4-
(2-[7-amino-2-(2-furyl)-1,2,4-triazolo[1,5-a] (1, 3, 5)triazin-5-
ylamino]ethyl] phenol, that binds the A
2A
receptor with high
affinity (28) antagonized the effects of adenosine on the phos-
phorylation of TrkA receptors (Fig. 3A). These results are
consistent with the involvement of adenosine A
2A
receptors in
mediating the increase in Trk receptor activity.
The mechanism by which G protein-coupled receptors are
linked to the activation of receptor tyrosine kinases is not well
understood. Src family kinases have been implicated as media-
tors of mitogenic receptor tyrosine kinase transactivation by
several G protein-coupled receptor agonists, such as lysophos-
phatidic acid, angiotensin II, thrombin, and bradykinin (14, 29).
To test whether a Src family member is involved in the activation
of Trk receptors by adenosine, the inhibitor PP1 (30) was used.
Treatment of PC12 cells with 1
M PP1 resulted in a marked
decrease in the level of tyrosine phosphor ylated TrkA receptors
elicited by adenosine (Fig. 3B). Increasing concentrations of PP1
produced a progressively stronger inhibition. These results sug-
gest that the regulation of TrkA activity by adenosine may be
mediated by a Src family member. An involvement of Src was
previously implicated in NGF signaling downstream of its re-
ceptor (31). However, it is conceivable that members of the Src
tyrosine kinase activity may be activated by G proteins. This has
been demonstrated for Lck, which acts in thymocytes down-
stream of the
-adrenergic receptor and whose activity can be
increased in vitro by Gs (32).
Hippocampal Neurons. To extend the generality of adenosine
effects on Trk receptors, we established primary hippocampal
neuronal cultures from E17 rat embryos. Hippocampal neurons
predominately express the TrkB receptor and respond to BDNF
(Fig. 4). These neurons also express and A
1
and A
2A
receptors
(33), but not TrkA receptors. Treatment with 10
M adenosine
or 10 nM CGS 21680 for 2 h gave rise to phosphorylated TrkB
receptors in hippocampal neurons (Fig. 4 Aand B), similar to the
activation of TrkA receptors by adenosine. An A
1
-specific
agonist, CPA, however, did not activate TrkB receptors (Fig.
4B), confirming the specificity of this effect to the A
2A
receptor.
These results not only extend the effects of adenosine to
hippocampal neurons, but also demonstrate that TrkB can also
be activated by signaling through the A
2A
receptors.
Downstream Signal Transduction. To characterize the signaling
pathways activated by adenosine, further experiments were
carried out. Pretreatment with 100 nM K252a abolished ade-
nosine’s activation of TrkA tyrosine kinase activity (Fig. 5). This
concentration of K252a has been used to block NGF activation
of TrkA receptors (34) and the subsequent biological effects of
neurotrophins.
Many G protein-coupled receptors activate the MAP kinase
pathway. Indeed, within 10 min of adenosine treatment, a
marked increased in phosphorylated MAP kinase was detected
Fig. 2. Time course and dose of adenosine activation of TrkA receptors. (A)
Different concentrations of adenosine were administered to PC12 615 cells for
2 h or NGF (1 ngml) for 10 min. Cells were also treated with CGS 21680 at the
indicated doses for 2 h. (B) PC12 cells (615) were treated with adenosine (10
M) for various times or with 5 ngml NGF for 10 min. Phosphorylated TrkA
receptors were detected by immunoblot analysis using PY99 anti-phosphoty-
rosine antibodies. The amount of Trk receptors in each condition was verified
by immunoblotting.
Table 1. No effect of adenosine on
125
I-NGF binding
Condition Specific binding
PC12
Control 14316 350
Adenosine (1 mM) 14403 888
CGS 21680 (1
M) 14237 1055
293TrkA
Control 39078 2885
Adenosine (1 mM) 36618 4185
CGS 21680 (1
M) 39568 2032
Fig. 3. Adenosine activation of TrkA by A2A receptors. (A) PC12 cells (615)
were treated with the A2A agonist CGS 21680 (10 nM) and an A1agonist CPA
(10 nM) for 2 h. ZM 241385 (10 nM), an A2A antagonist, was incubated with the
cells for 15 min before treatment with adenosine (10
M)for2h.(B) PC12 cells
(615) were incubated with the indicated concentrations of PP1, a Src family
kinase inhibitor (30), for 30 min, and then treated with adenosine (10
M) for
2 h. Activation of TrkA was assessed by immunoprecipitation and immunoblot
analysis using PY99 anti-phosphotyrosine antibody.
Lee and Chao PNAS
March 13, 2001
vol. 98
no. 6
3557
NEUROBIOLOGY
in PC12 cells (Fig. 5), consistent with previous observations
(35–37). After 10 min, the levels of activated MAP kinase
declined to a baseline level. Activation of MAP kinases can be
achieved either by A
2A
receptors, or through Trk receptor
signaling. To distinguish between these alternatives, PC12 cells
were treated with adenosine in the presence and absence of
K252a. Using a concentration of K252a that specifically blocks
TrkA signaling (34), we found that MAP kinase activity was not
altered (Fig. 5). Thus, MAP kinase induction occurred quickly,
whereas Trk activation by adenosine followed a slower time
course and did not influence MAP kinase activity. This result is
in contrast to other examples of G protein-coupled receptor
transactivation, in which MAP kinase activities are directly
stimulated downstream of the tyrosine kinase receptor (14).
Another pathway activated by receptor tyrosine kinases is
phosphatidylinositol 3-kinase (PI3-kinase)Akt. Interestingly,
adenosine (Fig. 5) or CGS 21680 treatment (data not shown) of
PC12 cells was also able to activate Akt as detected by a
phospho-specific antibody. This response has not been previ-
ously associated with adenosine action. The time course of Akt
activation was very similar to Trk autophosphorylation induced
by adenosine. This effect was eliminated either by pretreatment
with K252a (100 nM) or LY2494002 (10
M), a PI3-kinase
inhibitor. These results indicate that Akt activation by adenosine
is Trk- and PI3-kinase-dependent.
Trophic Effects. To test the functional consequences of adenosine-
activated Trk receptor activity, we assessed the ability of aden-
osine to maintain survival of differentiated PC12 cells after
withdrawal of NGF. After culture for 48 h in the absence of
NGF, cell survival was assessed by measuring LDH release.
Whereas cells grown without NGF underwent rapid cell death,
a one-time treatment with CGS 21680 effectively rescued nearly
50% of the cells (Fig. 6A). The action of CGS 21680 required
TrkA receptors, because K252a (100 nM) eliminated the positive
effects of CGS 21680 under similar conditions that blocked the
activation of TrkA receptors (Fig. 5). Likewise, a similar dose of
K252a reversed the survival effects of NGF, but not of insulin-
like growth factor 1, in this deprivation assay.
Similar survival results with CGS 21680 were obtained in
hippocampal neurons grown in the absence of BDNF (Fig. 6B).
The action of CGS 21680 was again dose-dependent and K-252a-
sensitive. Treatment with CGS 21680 effectively rescued 60%
of the cells (Fig. 6B). Hence, a potent adenosine agonist at
nanomolar concentrations was able to reverse cell death in both
PC12 cells and hippocampal neurons initiated by withdrawal of
trophic support by neurotrophins.
The ability of K252a to block adenosine’s trophic effect as well
as induction of Trk receptor activity suggested that Trk receptor
downstream signaling was involved in this process. This was
confirmed by the ability of LY294002 to eliminate the trophic
effect of CGS 21680 after NGF withdrawal (Fig. 6), indicating
that the PI3-kinaseAkt pathway was involved in the survival
effects of adenosine. Consistent with the lack of a MAP kinase
response, we did not find that the mitogen-activated kinase
kinase inhibitor PD98059 had any effect on survival imparted by
CGS 21680.
Discussion
Adenosine receptor activation leads to many modulatory effects
on neuropeptide and neurotransmitter systems (38). Here, we
report a property of adenosine in neuronal cells that affects
neurotrophin signaling. Through crosstalk with Trk receptor
tyrosine kinases, adenosine is capable of activating the PI3-
kinaseAkt cascade, resulting in a survival response in PC12 and
hippocampal cells. This response is similar to the effect of NGF
and BDNF on their Trk receptors, but differs in the longer time
course.
Neurotrophin receptors and A
2A
receptors have considerable
overlap in their central and peripheral nervous system distribu-
tion. In the central nervous system, A
2A
receptors are expressed
in striatum, amygdala, and olfactory tubercles, and in cerebral
cortex, hippocampus, and cerebellum (39). All of these regions
express TrkB receptors. In the peripheral nervous system, A
2A
receptor expression has been localized primarily to dorsal root
Fig. 4. Adenosine activation of TrkB receptors in hippocampal neurons.
Primary cultures of E17 hippocampal neurons were prepared as described in
Materials and Methods and treated with (A) CGS 21680 (10 nM) or BDNF (1
ngml) for various times and (B) adenosine (10
M), CGS 21680 (10 nM), CPA
(10 nM), or BDNF (10 ngml) for 2 h. Activation of TrkB receptors was assessed
by immunoprecipitation and Western blotting with anti-phosphotyrosine
antibody.
Fig. 5. Effects of adenosine on MAP kinase and Akt activation. PC12 cells
(615) were treated with adenosine (10
M) for various times in the presence
or absence of K252a (100 nM) or LY294002 (10
M). The cells were subse-
quently harvested in lysis buffer; lysates and immunoprecipitated samples
were subsequently immunoblotted with anti-phospho-MAP kinase, anti-
phospho-Akt, and anti-phosphotyrosine. Reprobing with anti-MAP kinase
and anti-pan-Trk antibodies was carried out to ensure equal protein loading.
3558
www.pnas.orgcgidoi10.1073pnas.061020198 Lee and Chao
ganglion and superior cervical ganglion (40), two regions that
express TrkA receptors. Interestingly, mice deficient in the
adenosine A
2A
receptor display decreased sensitivity to thermal
stimulation (41). It is noteworthy that mice with mutations in
NGF or TrkA also display hypoalgesia to thermal and mechan-
ical stimuli.
What are the in vivo consequences of these events observed in
culture? During hypoxia or ischemic conditions, adenosine is
released in large amounts and can mediate cellular protection.
A
2A
receptor agonists, such as CGS 21680, have been shown to
be neuroprotective against ischemia (42, 43) and kainate-
induced neuronal damage (44) in animals. However, A
2A
an-
tagonists have been also reported to reduce hypoxic-ischemic
neuronal injury (43, 45). The differential effects of A
2A
receptor
ligands may reflect short term versus long term effects by
adenosine receptors (46). Acute effects of adenosine analogs
may lead to opposite effects on neuroprotection than chronic
treatment. Engagement of receptor tyrosine kinases such as the
Trk subfamily may account for differences in the functional
consequences of adenosine action. A distinctive feature of
adenosine’s transactivation of Trk is the longer time course of
Trk mediated signaling, which is similar to neurotrophin-induced
signaling.
Adenosine has been proposed as a potential treatment for a
wide number of neurological disorders, including cerebral isch-
emia, sleep disorders, hyperalgesia, Parkinson’s disease, and
other neurodegenerative conditions (47). Previous effects on
growth arrest and enzyme induction by adenosine in PC12 cells
(48, 49) might well be explained by transactivation of TrkA
receptors. Our findings on adenosine delineate a pathway for
activating the neurotrophin signaling system in the absence of
neurotrophins. In contrast to other transactivation events in-
volving receptor tyrosine kinases that lead to transient increases
in MAP kinase activity, G protein-coupled receptor signaling to
neurotrophin receptors leads to selective activation of the PI3-
kinaseAkt pathway over a prolonged time course.
These findings provide a mechanism for the neuroprotective
actions of adenosine involving engagement of the A
2A
receptor,
transactivation of Trk tyrosine kinase receptors, and selective
induction of the PI3-kinaseAkt pathway. A number of ap-
proaches have been taken to use neurotrophins to treat Alzhei-
mer’s dementia, amyotrophic lateral sclerosis, and peripheral
sensory neuropathy (50, 51). However, there are considerable
hurdles in the use of neurotrophic molecules that are related to
difficulties in their delivery and pharmcokinetics and unantici-
pated side effects (51). The selective and sustained effects of
adenosine on survival signaling pathways suggest that small
molecules may be used to target populations of neurons that
express both adenosine and Trk receptors. The identification of
small ligands in the G protein-coupled receptor family that
regulate tyrosine kinase activity in neural cells offers a strategy
for promoting trophic effects during normal and neurodegen-
erative conditions.
We thank B. Hempstead and A. Kim for advice and R. Rajagopal and
L. Aibel for assistance. F.S.L. was supported by the DeWitt–Wallace
Fund in the New York Community Trust and an American Psychiatric
Association Program for Minority Research Training in Psychiatry
Fellowship, and M.V.C. was supported by grants from the National
Institutes of Health.
1. Lev i-Montalcini, R. (1987) Science 237, 1154–1164.
2. Lew in, G. R. & Barde, Y.-A. (1996) Annu. Rev. Neurosci. 19, 289–317.
3. Thoenen, H. (1995) Science 270, 593–598.
4. Bonhoeffer, T. (1996) Curr. Opin. Neurobiol. 6, 119–126.
5. Gallo, G., Lefcort, F. & Letourneau, P. (1997) J. Neurosci. 17, 5445–5454.
6. Chao, M. V. & Hempstead, B. L. (1995) Trends Neurosci. 19, 321–326.
7. McAllister, A., Katz, L. & Lo, D. (1999) Annu. Rev. Neurosci. 22, 295–318.
8. Minichiello, L., Korte, M., Wolfer, D., Kuhn, R., Unsicker, K., Cestari, V., Rossi-
Arnaud, C., Lipp, H. P., Bonhoeffer, T. & Klein, R. (1999) Neuron 24, 401– 414.
9. Lyons, W. E., Mamounas, L. A., Ricaurte, G. A., Coppola, V., Reid, S. W.,
Bora, S. H., Wihler, C., Koliatsos, V. E. & Tessarollo, L. (1999) Proc. Natl.
Acad. Sci. USA 96, 15239 –15244.
10. Dobrowsky, R. T., Jenkins, G. M. & Hannun, Y. A. (1995) J. Biol. Chem. 270,
22135–22142.
11. Yoon, S. O., Carter, B. D., Casaccia-Bonnefil, P. & Chao, M. V. (1998)
J. Neurosci. 18, 3273–3281.
12. Daub, H., Weiss, F. U., Wallasch, C. & Ullrich, A. (1996) Nature (London) 379,
557–560.
13. Castellino,A. M. & Chao, M. V. (1997) Cytokine Grow th Factor Rev. 7, 297–302.
14. Luttrell, L., Daaka, Y. & Lefkowitz, R. (1999) Curr. Opin. Cell Biol . 11,
177–183.
15. Linseman, D. A., Benjamin, C. W. & Jones, D. A. (1995) J. Biol. Chem. 270,
12563–12568.
16. Rao, G., Delafontaine, P. & Runge, M. (1995) J. Biol. Chem. 270, 27871–27875.
Fig. 6. Trophic effects of adenosine agonist in PC12 and hippocampal cells
deprived of neurotrophins. (A) NGF-differentiated PC12 cells were pre-
pared, and then NGF and serum were withdrawn for 48 h as described in
Materials and Methods. On NGF withdrawal, various concentrations of CGS
21680 (CGS) were added to the media. CON, no addition. NGF (50 ngml),
insulin-like growth factor 1 (IGF-1) at 100 ngml, and CGS 21680 (10 nM)
were added together with K252a (100 nM), LY294002 (10
M), or PD98059
(25
M) upon NGF withdrawal. (B) Hippocampal neurons were prepared
and B27 was withdrawn for 48 h as described in Materials and Methods.
Upon B27 withdrawal, various concentrations of CGS 21680 (CGS) were
added to the media. CGS 21680 (10 nM) and BDNF (100 ngml) were added
together with K252a (100 nM) on B27 withdrawal. All LDH levels
were quantitated and % cell death calculated as described in Materials
and Methods. All bars depict mean SEM from three independent
experiments.
Lee and Chao PNAS
March 13, 2001
vol. 98
no. 6
3559
NEUROBIOLOGY
17. Hempstead, B. L., Rabin, S. J., Kaplan, L., Reid, S., Parada, L. F. & Kaplan,
D. R. (1992) Neuron 9, 883–896.
18. Hempstead, B. L., Schleifer, L. & Chao, M. V. (1989) Science 243,
373–375.
19. Aibel, L., Martin-Zanca, D., Perez, P. & Chao, M. (1998) J. Neurosci. Res. 54,
424– 431.
20. Kim, A., Sheline, C., Tian, M., Higashi, T., McMahon, R., Cousins, R. & Choi,
D. (2000) Brain Res. 886, 99–107.
21. Etschied, B., Ko, P. & Villereal, M. (1991) Br. J. Phamacol. 103, 1347–1350.
22. Inoue, K., Nakazawa, K., Watano, T., Ohara-Imaizumi, M., Fujimori, K. &
Takanaka, A. (1992) Eur. J. Pharmacol. 215, 321–324.
23. Williams, M., Abreu, M., Jarvis, M. & Noronha-Blob, L. (1987) J. Neurochem.
48, 498–502.
24. Kim, W. & Rabin, R. (1994) J. Biol. Chem. 269, 6471–6477.
25. Jarvis, M., Schulz, R., Hutchinson, A., Do, U., Sills, M. & Williams, M. (1989)
J. Pharmacol. Exp. Ther. 251, 888 –893.
26. Near y, J., Rathbone, M., Cattabeni, F., Abbracchio, M. & Burnstock, G. (1996)
Trends Neurosci. 19, 1996.
27. Ralev ic, V. & Burnstock, G. (1998) Pharmacol. Rev. 50, 413–492.
28. Poucher, S. M., Keddie, J. R., Singh, P., Stoggall, S. M., Caulkett, P. W., Jones,
G. & Coll, M. G. (1995) Br. J. Pharmacol. 115, 1096 –1102.
29. Daub, H., Wallash, C., Lankenau, A., Herrlich, A. & Ullrich, A. (1997) EMBO
J. 16, 7032–7044.
30. Hanke, J. H., Gardner, J. P., Dow, R. L., Changelian, P. S., Brissette, W. H., Weringer,
E. J., Pollok, B. A. & Connelly, P. A. (1996) J. Biol. Chem. 271, 695–701.
31. D’Arcangelo, G. & Halegoua, S. (1993) Mol. Cell. Biol. 13, 3146–3155.
32. Gu, C., Ma, Y.-C., Benjamin, J., Littman, D., Chao, M. & Huang, X.-Y. (2000)
J. Biol. Chem. 275, 20726–20733.
33. Dixon, A., Gabitz, A., Sirinathsinghji, D., Richardson, P. & Freeman, T. (1996)
Br. J. Pharmacol. 118, 1461–1468.
34. Berg, M. M., Sternberg, D., Parada, L. F. & Chao, M. V. (1992) J. Biol. Chem.
267, 13–16.
35. Sexl, V., Mancusi, G., Holler, C., Gloria-Maercker, E., Schutz, W. & Freiss-
muth, M. (1997) J. Biol. Chem. 272, 5792–5799.
36. Gao, Z., Chen, T., Weber, M. & Linden, J. (1999) J. Biol. Chem. 274, 5972–5980.
37. Seidel, M., Klinger, M., Freissmuth, M. & Holler, C. (1999) J. Biol. Chem. 274,
25833–25841.
38. Sebastiao, A. & Ribeiro, J. (2000) Trends Pharmacol. Sci. 21, 341–346.
39. Rosin, D., Robeva, A., Woodard, R., Guyenet, P. & Linden, J. (1998) J. Comp.
Neurol. 401, 163–186.
40. Kaelin-Lang, A., L auterburg, T. & Burgunder, J.-M. (1998) Neurosci. Lett. 246,
21–24.
41. Ledent, C., Vaugeois, J.-M., Schiffmann, S. N., Pedrazzini, T., El Yacoubi, M.,
Vanderhaeghen, J.-J., Costentin, J., Heath, J. K., Vassart, G. & Parmentier, M.
(1997) Nature (London) 388, 674– 678.
42. Scheardown, M. & Knutsen, L. (1996) Drug Dev. Res. 39, 108–114.
43. von Lubitz, D., Lin, R. & Jacobson, K. (1995) Eur. J. Pharmacol. 287,
295–302.
44. Jones, P., Smith, R. & Stone, T. (1998) Neuroscience 85, 229 –237.
45. Phillis, J. (1995) Brain Res. 705, 79 –84.
46. Jacobson, K., von Lubitz, D., Daly, J. & Fredholm, B. (1996) Trends Phar macol.
Sci. 17, 108–113.
47. Moreau, J.-L. & Huber, G. (1999) Brain Res. Rev. 31, 65– 82.
48. Guroff, G., Dickens, G., End, D. & Londos, C. (1981) J. Neurochem. 37,
1431–1439.
49. Huffaker, T., Corc oran, T. & Wagner, J. A. (1984) J. Cell. Physiol. 120, 188–196.
50. Hefti, F. (1994) J. Neurobiol. 25, 1418–1435.
51. Thoenen, H. (2001) in Axonal Regeneration in the Central Nervous System, eds.
Ingoglia, N. & Murray, M. (Dekker, New York), pp. 675–697.
3560
www.pnas.orgcgidoi10.1073pnas.061020198 Lee and Chao
... With the exception of FGF, all the factors displayed statistically non-significant effects on cell survival (Figure 8). We have used deprivation of hippocampal neurons from lack of B27 (Lee and Chao, 2001;Jeanneteau et al., 2008) with reproducible results. However, despite much work on the viability of hippocampal neurons, a bona fide hippocampal growth factor has yet to be confirmed in vivo. ...
... As B27 contains many factors that can change gene expression, it is an unnatural way of removing trophic support. We have used the lack of B27 nutrients previously (Lee and Chao, 2001;Jeanneteau et al., 2008) as a means to study the consequences of deprivation of hippocampal neurons. To align and verify the cell culture approach, microarray experiments to compare with the glucocorticoid transcriptomes have been undertaken (Lambert et al., 2013;Mariga et al., 2014). ...
... Previously we used deprivation of hippocampal neurons from lack of B27 with consistent results to identify small molecule ligands of GPCRs that provide a survival outcome (Lee and Chao, 2001;Jeanneteau et al., 2008). The interest in the glucocorticoid receptor (GR) is through its role as a ligand-activated transcription factor that mediates many physiological processes, including stress and inflammation. ...
Article
Full-text available
Classic experiments with peripheral sympathetic neurons established an absolute dependence upon NGF for survival. A forgotten problem is how these neurons become resistant to deprivation of trophic factors. The question is whether and how neurons can survive in the absence of trophic support. However, the mechanism is not understood how neurons switch their phenotype to lose their dependence on trophic factors, such as NGF and BDNF. Here, we approach the problem by considering the requirements for trophic support of peripheral sympathetic neurons and hippocampal neurons from the central nervous system. We developed cellular assays to assess trophic factor dependency for sympathetic and hippocampal neurons and identified factors that rescue neurons in the absence of trophic support. They include enhanced expression of a subunit of the NGF receptor (Neurotrophin Receptor Homolog, NRH) in sympathetic neurons and an increase of the expression of the glucocorticoid receptor in hippocampal neurons. The results are significant since levels and activity of trophic factors are responsible for many neuropsychiatric conditions. Resistance of neurons to trophic factor deprivation may be relevant to the underlying basis of longevity, as well as an important element in preventing neurodegeneration.
... Finally, considering the role of Src family kinases in the neurotrophin-independent Trk receptor transactivation, 67 as well as mediating temperature-associated changes in the metabolic rate, 68 we assessed the active phosphorylation of Src in brain homogenates that showed increased TrkB phosphorylation. Indeed, Src Y416 phosphorylation was rapidly induced by amitriptyline, isoflurane, metabolic inhibitors 2-DG + MA, and during recovery from SD and flurothyl seizure ( Figures 4A and S7). ...
... 12,24 While the precise mechanism of the activation remains to be fully elucidated, one possibility is the involvement of Src family kinases, which have been previously associated with transactivation of the TrkB. 67 Previous studies have shown that Src, as well as several other targets associated with TrkB signaling (most notably GSK3β), are indeed regulated by the metabolic milieu through mechanisms such as oxidative stress and altered neuronal lipid environment. 90,91 Beyond the implications to therapeutic mechanisms of various drugs, the findings of this report raise crucial questions regarding the reproducibility and translational value of the neuropharmacological research. ...
Article
Many mechanisms have been proposed to explain acute antidepressant drug-induced activation of TrkB neurotrophin receptors, but several questions remain. In a series of pharmacological experiments, we observed that TrkB activation induced by antidepressants and several other drugs correlated with sedation, and most importantly, coinciding hypothermia. Untargeted metabolomics of pharmacologically dissimilar TrkB activating treatments revealed effects on shared bioenergetic targets involved in adenosine triphosphate (ATP) breakdown and synthesis, demonstrating a common perturbation in metabolic activity. Both activation of TrkB signaling and hypothermia were recapitulated by administration of inhibitors of glucose and lipid metabolism, supporting a close relationship between metabolic inhibition and neurotrophic signaling. Drug-induced TrkB phosphorylation was independent of electroencephalography slow-wave activity and remained unaltered in knock-in mice with the brain-derived neurotrophic factor (BDNF) Val66Met allele, which have impaired activity-dependent BDNF release, alluding to an activation mechanism independent from BDNF and neuronal activity. Instead, we demonstrated that the active maintenance of body temperature prevents activation of TrkB and other targets associated with antidepressants, including p70S6 kinase downstream of the mammalian target of rapamycin (mTOR) and glycogen synthase kinase 3β (GSK3β). Increased TrkB, GSK3β, and p70S6K phosphorylation was also observed during recovery sleep following sleep deprivation, when a physiological temperature drop is known to occur. Our results suggest that the changes in bioenergetics and thermoregulation are causally connected to TrkB activation and may act as physiological regulators of signaling processes involved in neuronal plasticity.
... The restoration of plasticity-associated endpoints via the exogenous application of neurotrophins is a Trk isoform-dependent process and appears to be a unique property of BDNF among neurotrophins [127]. The stimulation of adenosine-2A receptors, believed to directly activate TrkB receptors and potentiate the BDNF-mediated modulation of synaptic transmission, also appeared to restrain excitotoxicity in R6/2 rodents [128,129]. The expression of mHTT is not merely associated with reduced plastic capacity within LTP paradigms, but also affects the activity-dependent release of BDNF into the synaptic cleft, at least at corticostriatal synapses. ...
Article
Full-text available
Brain-derived neurotrophic factor (BDNF) is a major neurotrophin whose loss or interruption is well established to have numerous intersections with the pathogenesis of progressive neurological disorders. There is perhaps no greater example of disease pathogenesis resulting from the dysregulation of BDNF signaling than Huntington’s disease (HD)—an inherited neurodegenerative disorder characterized by motor, psychiatric, and cognitive impairments associated with basal ganglia dysfunction and the ultimate death of striatal projection neurons. Investigation of the collection of mechanisms leading to BDNF loss in HD highlights this neurotrophin’s importance to neuronal viability and calls attention to opportunities for therapeutic interventions. Using electronic database searches of existing and forthcoming research, we constructed a literature review with the overarching goal of exploring the diverse set of molecular events that trigger BDNF dysregulation within HD. We highlighted research that investigated these major mechanisms in preclinical models of HD and connected these studies to those evaluating similar endpoints in human HD subjects. We also included a special focus on the growing body of literature detailing key transcriptomic and epigenetic alterations that affect BDNF abundance in HD. Finally, we offer critical evaluation of proposed neurotrophin-directed therapies and assessed clinical trials seeking to correct BDNF expression in HD individuals.
... The second endogenous TrkB ligand NT4 does not account for the ligandreceptor imbalance as NT4 is barely detectable in the adult CNS. 1 Observations indicating that TrkB can be activated by non-neurotrophin ligands, including antidepressants (see below), may contribute to explain the quantitative discrepancies between the levels of TrkB and those of BDNF. In addition, zinc, various G-protein coupled receptor agonists as well as EGF receptor ligands have all been shown to activate TrkB (Lee and Chao, 2001;Huang et al., 2008;Puehringer et al., 2013;Zagrebelsky et al., 2018;Casarotto et al., 2021). ...
Article
Full-text available
This review focuses on neurotrophins and their tyrosine kinase receptors, with an emphasis on their relevance to the function and dysfunction in the human nervous system. It also deals with measurements of BDNF levels and highlights recent findings from our laboratory on TrkB and TrkC signalling in human neurons. These include ligand selectivity and Trk activation by neurotrophins and non-neurotrophin ligands. The ligand-induced down-regulation and re-activation of Trk receptors is also discussed.
Article
Full-text available
Metabotropic glutamate receptor 2 (mGlu 2) attracts particular attention as a possible target for a new class of antipsychotics. However, the signaling pathways transducing the effects of mGlu 2 in the brain remain poorly characterized. Here, we addressed this issue by identifying native mGlu 2 interactome in mouse prefrontal cortex. Nanobody-based affinity purification and mass spectrometry identified 149 candidate mGlu 2 partners, including the neurotrophin receptor TrkB. The later interaction was confirmed both in cultured cells and prefrontal cortex. mGlu 2 activation triggers phosphorylation of TrkB on Tyr 816 in primary cortical neurons and prefrontal cortex. Reciprocally, TrkB stimulation enhances mGlu 2-operated G i/o protein activation. Furthermore, TrkB inhibition prevents the rescue of behavioral deficits by glutamatergic antipsychotics in phencyclidine-treated mice. Collectively, these results reveal a cross-talk between TrkB and mGlu 2 , which is key to the behavioral response to glutamatergic antipsychotics.
Article
Full-text available
Cellular signaling involves a large repertoire of membrane receptors operating in overlapping spatiotemporal regimes and targeting many common intracellular effectors. However, both the molecular mechanisms and the physiological roles of crosstalk between receptors, especially those from different superfamilies, are poorly understood. We find that the receptor tyrosine kinase (RTK) TrkB and the G-protein-coupled receptor (GPCR) metabotropic glutamate receptor 5 (mGluR5) together mediate hippocampal synaptic plasticity in response to brain-derived neurotrophic factor (BDNF). Activated TrkB enhances constitutive mGluR5 activity to initiate a mode switch that drives BDNF-dependent sustained, oscillatory Ca²⁺ signaling and enhanced MAP kinase activation. This crosstalk is mediated, in part, by synergy between Gβγ, released by TrkB, and Gαq-GTP, released by mGluR5, to enable physiologically relevant RTK/GPCR crosstalk.
Article
Neurotrophic factors are a family of growth factors that modulate cellular growth, survival, and differentiation. For many decades, it has been generally believed that a lack of neurotrophic support led to the decreased neuronal synaptic plasticity, death, and loss of non-neuronal supportive cells seen in neuropsychiatric disorders. Traditional psychiatric medications that lead to immediate increases in neurotransmitter levels at the synapse have been shown also to elevate synaptic neurotrophic levels over weeks, correlating with the time course of the therapeutic effects of these drugs. Recent advances in psychiatric treatments, such as ketamine and psychedelics, have shown a much faster onset of therapeutic effects (within minutes to hours). They have also been shown to lead to a rapid release of neurotrophins into the synapse. This has spurred a significant shift in understanding the role of neurotrophins and how the receptor tyrosine kinases that bind neurotrophins may work in concert with other signaling systems. In this review, this renewed understanding of synaptic receptor signaling interactions and the clinical implications of this mechanistic insight will be discussed within the larger context of the well-established roles of neurotrophic factors in psychiatric disorders and treatments.
Article
The insulin superfamily of peptides is essential for homeostasis as well as neuronal plasticity, learning, and memory. Here, we show that insulin-like growth factors 1 and 2 (IGF1 and IGF2) are differentially expressed in hippocampal neurons and released in an activity-dependent manner. Using a new fluorescence resonance energy transfer sensor for IGF1 receptor (IGF1R) with two-photon fluorescence lifetime imaging, we find that the release of IGF1 triggers rapid local autocrine IGF1R activation on the same spine and more than several micrometers along the stimulated dendrite, regulating the plasticity of the activated spine in CA1 pyramidal neurons. In CA3 neurons, IGF2, instead of IGF1, is responsible for IGF1R autocrine activation and synaptic plasticity. Thus, our study demonstrates the cell type-specific roles of IGF1 and IGF2 in hippocampal plasticity and a plasticity mechanism mediated by the synthesis and autocrine signaling of IGF peptides in pyramidal neurons.
Chapter
Purinergic nucleosides are for a long time known as neuromodulators, controlling the release and action of neurotransmitters. More recently, evidence emerged for adenosine and guanosine to have trophic actions. In this chapter, we address evidence for neurotrophic actions mediated by adenosine and guanosine in cell proliferation, cell migration, cell differentiation, neurite outgrowth, and synaptogenesis. Alterations in any of these processes may compromise brain function, leading to cognitive impairment and neurodegenerative diseases. Therefore, understanding the molecular and cellular mechanisms that modulate these trophic actions is the center of attention of regenerative research. Purine nucleosides could certainly play a role there.KeywordsAdenosineGuanosineCellular proliferationNeurite outgrowthNeuronal differentiation
Chapter
Brain tumors are primary or metastatic tumors that proliferate in the brain tissue itself or surrounding tissues, such as pituitary and pineal glands, brain-covering membranes, and cranial nerves. Gliomas alone account for 80% of all malignant primary brain tumors; however, metastatic brain tumors are estimated to have a 10–20 times higher than all other primary brain tumors together. Although there are similarities between primary and metastatic brain tumors, to colonize the brain parenchyma, tumor cells must be able to cross the blood-brain barrier (BBB), evade the actions of CNS immune cells, and finally grow in the highly unique brain microenvironment due to distinct brain features. The purinergic system is implicated in the acquisition of a more neural phenotype of tumor cells and in the preparation of the metastatic niche conferring a growth advantage to a population of metastatic cells, which allows colonization of the brain and subsequent neurological injury. Metabolic adaptations are also required for brain colonization. Brain tumor cells, even in an aerobic microenvironment, use the energy of glucose by oxidizing pyruvate to lactate, just as they would under hypoxia, a metabolic pathway that produces less ATP. Tumor cells expressing purinoreceptors overexpress glycolytic enzymes and have altered fatty acid metabolism to guarantee sufficient ATP levels for cancer development and survival. Furthermore, purinergic signaling is involved in the autoregulatory program of the brain tumor microenvironment, which involves phenotype regulation of tumor-associated macrophage-microglia (TAMs) for an antitumoral or protumoral response. The purinergic system has been related to different protumoral responses of macrophages, such as angiogenesis, tumor invasion, chemoresistance, and immunosuppression. On the other hand, purinergic signaling also mediates several antitumoral responses, such as TAM phagocytic activity and microglial proliferation and chemotaxis to the affected site to destroy foreign cells or particles and alert the immune system to preeminent danger. Last, there is evidence of the involvement of purinergic signaling in the symptoms of brain tumor development, including neurological deficits, confusion, memory loss, and changing personality. However, studies are still needed to elucidate these pathways for the mitigation of these symptoms. In summary, the purinergic system is an important exploration field for further approaches to avoid and treat brain tumors and their symptoms. In this chapter, details of purinergic signaling for brain tumor development and colonization are described.KeywordsTumor microenvironmentTumor-associated macrophage-microgliaExtracellular ATP
Article
Full-text available
It has recently been reported that protein-tyrosine kinase activity is required for thrombin-induced growth in vascular smooth muscle cells (VSMC). In the present study, we have identified several phosphoproteins that are tyrosine-phosphorylated in response to thrombin in quiescent VSMC. These proteins are insulin-like growth factor-1 receptor β-subunit (IGF-IRβ), insulin receptor substrate-1 (IRS-1), and phospholipase C-γ1 (PLC-γ1). Thrombin-stimulated phosphorylation of these proteins was rapid; it was maximal at 1 min and reduced thereafter. Thrombin also activated mitogen-activated protein kinases (MAPK) in quiescent VSMC in a biphasic manner with a rapid and larger peak at 10 min (6-fold) followed by a sustained smaller second peak at 2 h (2-fold). Inhibition of protein-tyrosine kinase activity by the use of two structurally different protein-tyrosine kinase inhibitors, genistein and herbimycin A, significantly blocked the thrombin-induced tyrosine phosphorylation of IGF-1Rβ, IRS-1, and PLC-γ1 and decreased thrombin-stimulated DNA synthesis. In contrast, however, inhibition of protein-tyrosine kinase activity had no effect on thrombin activation of MAPK. Collectively, these findings suggest a role for tyrosine phosphorylation of IGF-IRβ, IRS-1, and PLC-γ1 in thrombin-induced mitogenic signaling events in VSMC. Furthermore, while protein tyrosine phosphorylation is essential for thrombin-induced DNA synthesis, it is not required for thrombin-stimulated MAPK activation. Since thrombin rapidly activated Src in VSMC, Src may be involved in the cross-talk between the G-protein-coupled receptor agonist and a tyrosine kinase receptor such as IGF-1R.
Article
Full-text available
The A2A-adenosine receptor, a prototypical Gs-coupled receptor, activates mitogen-activated protein (MAP) kinase in a manner independent of cAMP in primary human endothelial cells. In order to delineate signaling pathways that link the receptor to the regulation of MAP kinase, the human A2A receptor was heterologously expressed in Chinese hamster ovary (CHO) and HEK293 cells. In both cell lines, A2A agonist-mediated cAMP accumulation was accompanied by activation of the small G protein rap1. However, rap1 mediates A2A receptor-dependent activation of MAP kinase only in CHO cells, the signaling cascade being composed of Gs, adenylyl cyclase, rap1, and the p68 isoform of B-raf. This isoform was absent in HEK293 cells. Contrary to CHO cells, in HEK293 cells activation of MAP kinase by A2A agonists was not mimicked by 8-bromo-cAMP, was independent of Gαs, and was associated with activation of p21ras. Accordingly, overexpression of the inactive S17N mutant of p21ras and of a dominant negative version of mSos (the exchange factor of p21ras) blocked MAP kinase stimulation by the A2A receptor in HEK 293 but not in CHO cells. In spite of the close homology between p21ras and rap1, the S17N mutant of rap1 was not dominant negative because (i) overexpression of rap1(S17N) failed to inhibit A2A receptor-dependent MAP kinase activation, (ii) rap1(S17N) was recovered in the active form with a GST fusion protein comprising the rap1-binding domain of ralGDS after A2A receptor activation, and (iii) A2A agonists promoted the association of rap1(S17N) with the 68-kDa isoform of B-raf in CHO cells. We conclude that the A2A receptor has the capacity two activate MAP kinase via at least two signaling pathways, which depend on two distinct small G proteins, namely p21ras and rap1. Our observations also show that the S17N version of rap1 cannot be assumed a priori to act as a dominant negative interfering mutant.
Article
Full-text available
Brain-derived neurotrophic factor (BDNF) has trophic effects on serotonergic (5-HT) neurons in the central nervous system. However, the role of endogenous BDNF in the development and function of these neurons has not been established in vivo because of the early postnatal lethality of BDNF null mice. In the present study, we use heterozygous BDNF+/− mice that have a normal life span and show that these animals develop enhanced intermale aggressiveness and hyperphagia accompanied by significant weight gain in early adulthood; these behavioral abnormalities are known to correlate with 5-HT dysfunction. Forebrain 5-HT levels and fiber density in BDNF+/− mice are normal at an early age but undergo premature age-associated decrements. However, young adult BDNF+/− mice show a blunted c-fos induction by the specific serotonin releaser-uptake inhibitor dexfenfluramine and alterations in the expression of several 5-HT receptors in the cortex, hippocampus, and hypothalamus. The heightened aggressiveness can be ameliorated by the selective serotonin reuptake inhibitor fluoxetine. Our results indicate that endogenous BDNF is critical for the normal development and function of central 5-HT neurons and for the elaboration of behaviors that depend on these nerve cells. Therefore, BDNF+/− mice may provide a useful model to study human psychiatric disorders attributed to dysfunction of serotonergic neurons.
Article
A myriad of gene induction events underlie nerve growth factor (NGF)-induced differentiation of PC12 cells. To dissect the signal transduction pathways which lead to NGF actions, we have assessed the relative roles of NGF receptor, Src, Ras, and Raf activities in mediating specific gene inductions. We have used the PC12 cell line as well as sublines which inducibly express activated forms of either Src, Ras, or Raf or a dominant inhibitory form of Ras (p21N17 Ras) to study the expression of multiple NGF-inducible mRNAs. The NGF induction of NGFI-A, transin, and VGF mRNAs was mimicked by activated forms of Src, Ras, or Raf and was blocked by p21N17 Ras. The NGF induction of SCG10 mRNA was mimicked only by activated Src and Ras and was blocked by p21N17 Ras, while the induction of Thy-1 mRNA was mimicked only by activated Src and was not blocked by p21N17 Ras. The NGF induction of mRNAs for two sodium channel types was neither mimicked by any activated oncoprotein nor blocked by p21N17 Ras. From these and previous results, we suggest a model in which a linear order of NGF receptor, Src, Ras, and Raf activities is used by NGF to elicit gene inductions. These signaling components define branchpoints in the pathway to specific gene induction events, providing a mechanism for generating a host of diverse NGF actions.
Article
In general, neurons in the adult central nervous system of higher vertebrates are incapable of regeneration after injury. This is in contrast to the situation in the mammalian peripheral nervous system and in lower vertebrates where severed axons can regenerate, form correct synaptic connections and restore function. This enigma has been the subject of extensive research and in this review, we describe the recent developments in this area concentrating on growth inhibitory proteins associated with myelis. New studies have shown that these proteins may mask a remarkable capacity for sprouting and plasticity of uninjured neural connections.
Article
The neuroprotective role of adenosine receptor agonists in various models of ischaemia and neuronal excitotoxicity has been attributed to adenosine A1 receptor activation. In this study we examine the role of the A2A receptor in the kainate model of excitotoxicity. Kainate (10mg/kg) was administered systemically 10min after the intraperitoneal injection of adenosine analogues. The A2A agonist 2-p-(2-carboxyethyl)phenethylamino-5′-N-ethylcarboxamidoadenosine hydrochloride (CGS21680) protected the hippocampus at concentrations of 0.1 and 0.01mg/kg, but not at 2μg/kg. The addition of the centrally acting adenosine A1 receptor antagonist 8-cyclopentyl-1,3-dipropylxanthine partially reduced protection only in the CA3a region, suggesting that only a small proportion of the protection was attributable to the A1 receptor. A less potent A2A agonist, N6-[2-(3,5-dimethyoxyphenyl)-2-(2-methylphenyl)-ethyl]adenosine (1mg/kg), provided only partial protection against kainate. 4-(2-[7-Amino-2-{2-furyl}{1,2,4}triazolo{2,3-a}{1,3,5}triazin-5-yl-amino]ethyl)phenol, a selective A2A antagonist, also showed protection against kainate-induced neuronal death, when administered alone or in combination with CGS21680.These results show that adenosine A2A receptor activation is protective against excitotoxicity. The protection is largely independent of A1 receptor activation or blockade.
Article
Despite considerable evidence that neuronal activity influences the organization and function of circuits in the developing and adult brain, the molecular signals that translate activity into structural and functional changes in connections remain largely obscure. This review discusses the evidence implicating neurotrophins as molecular mediators of synaptic and morphological plasticity. Neurotrophins are attractive candidates for these roles because they and their receptors are expressed in areas of the brain that undergo plasticity, activity can regulate their levels and secretion, and they regulate both synaptic transmission and neuronal growth. Although numerous experiments show demonstrable effects of neurotrophins on synaptic plasticity, the rules and mechanisms by which they exert their effects remain intriguingly elusive.
Article
The selective adenosine A2A receptor agonists 2-[p-(2-carboxethyl)phenylethylaminol-5′-N-ethylcarboxyamidoadenosine (CGS 21680), N-[2-(3,5-dimethoxyphenyl)ethyl]adenosine (DPMA) and metrifudil were investigated for their ability to prevent the loss of pyramidal CA1 neurons in the hippocampus following 5 min of severe temporary forebrain ischemia in mongolian gerbils. CGS 21680, when administered at 18.7 μmol/kg 30 and 120 min following reperfusion, exhibited highly significant protection against neuronal loss, but was inactive at 5.6 μmol/kg. DPMA, a more potent agonist at A1 and A2A receptors, was inactive up to a dose of 19 μmol/kg. Metrifudil (equipotent with CGS 21680 at A2A >25 times more potent at A1) gave a modest degree of protection at 26 μmol/kg and was inactive at 7.8 μmol/kg. CGS 21680 showed an equal degree of hypothermia at 5.6 and 18.7 μmol/kg, suggesting that this was not the prime mode of action. While the effect of metrifudil may be the result of its higher A1 receptor affinity, the mode of action of CGS 21680 has not been established; the data, however, suggest that a non-A1 non-A2A receptor mechanism may possibly be involved. Drug Dev. Res. 39:108–114 © 1997 Wiley-Liss, Inc.
Article
The ability of neurotrophic factors to regulate developmental neuronal survival and adult nervous system planticity suggests the use of these molecuales to treat neurodegeneration associated with human diseases. Solid rationales exist for the use of NGF and neurotrophin-3 in the treatment of neuropathies of the peripheral sensory system, insulin-like growth factor and ciliary neurotrophic factor in motor neuron atrophy, and NGF in Alzheimer's disease. Growth factors have been identified for neurons affected in Parkinson's disease, Huntington's disease, and acute brain and spinal cord injury. Various strategies are actively pursued to deliver neurotrophic factors to the brain, and develop therapeutically useful molecules that mimic neurotrophic factor actions or stimulate their production or receptor mechanisms. 1994 John Wiley & Sons, Inc.