ArticlePDF Available

Rapid desensitization of the thyrotropin-releasing hormone receptor expressed in single human embryonal kidney 293 cells

Authors:

Abstract and Figures

This study uses fluorescence microscopy combined with dynamic video imaging to examine the events associated with the rapid desensitization of the thyrotropin-releasing hormone receptor (TRH-R). In single non-pituitary human embryonic kidney 293 (HEK-293) cells, expressing either the rat or human TRH-Rs, TRH produced a rapid dose-dependent monophasic rise in [Ca2+]i. This Ca2+ transient was completely abolished by pretreatment of cells with the intracellular Ca2+ antagonists thapsigargin or cyclopiazonic acid, but not EGTA, the voltage-operated Ca2+ channel (VOCC) antagonist nifedipine or the second-messenger-operated Ca2+ channel antagonist SK&F 96365. These results suggest that TRH causes the mobilization of Ca2+ from thapsigargin/cyclopiazonic acid-sensitive intracellular Ca2+ stores but not the influx of extracellular Ca2+. HEK-293 cells also failed to respond to KCl or the slow Ca(2+)-channel activator BAY K 8644, suggesting that they lack L-type VOCCs. Rat and human TRH-Rs are highly conserved except at the C-terminus where the sequence differs. The C-terminus is believed to be important in receptor desensitization. Despite differences in this region, rat and human TRH-Rs expressed in HEK-293 cells underwent rapid (within 1 min) desensitization. This desensitization was dose-dependent and did not involve receptor loss. Similarly the bradykinin receptor endogenous to HEK-293 cells also displays a rapid desensitization. We conclude that in TRH-R-expressing non-pituitary HEK-293 cells, TRH mobilizes intracellular Ca2+ resulting in a monophasic Ca2+ transient. The rat and human TRH-Rs as well as the endogenous bradykinin receptor also displayed rapid receptor desensitization.
Content may be subject to copyright.
Biochem.
J.
(1995)
311,
385-392
(Printed
in
Great
Britain)
Rapid
desensitization
of
the
thyrotropin-releasing
hormone
receptor
expressed
in
single
human
embryonal
kidney
293
cells
Lorraine
ANDERSON,*
Claire
L.
ALEXANDER,
Elena
FACCENDA
and
Karin
A.
EIDNE
MRC
Reproductive
Biology
Unit,
Centre
for
Reproductive
Biology,
37
Chalmers
Street,
Edinburgh
EH3
9EW,
Scotland,
U.K.
This
study
uses
fluorescence
microscopy
combined
with
dynamic
video
imaging
to
examine
the
events
associated
with
the
rapid
desensitization
of
the
thyrotropin-releasing
hormone
receptor
(TRH-R).
In
single
non-pituitary
human
embryonic
kidney
293
(HEK-293)
cells,
expressing
either
the
rat
or
human
TRH-Rs,
TRH
produced
a
rapid
dose-dependent
monophasic
rise
in
[Ca2+],.
This
Ca2+
transient
was
completely
abolished
by
pre-
treatment
of
cells
with
the
intracellular
Ca2+
antagonists
thapsigargin
or
cyclopiazonic
acid,
but not
EGTA,
the
voltage-
operated
Ca2+
channel
(VOCC)
antagonist
nifedipine
or
the
second-messenger-operated
Ca2+
channel
antagonist
SK&F
96365.
These
results
suggest
that
TRH
causes
the
mobilization
of
Ca2+
from
thapsigargin/cyclopiazonic
acid-sensitive
intra-
cellular
Ca2+
stores
but
not
the
influx
of
extracellular
Ca2
.
INTRODUCTION
Thyrotropin-releasing
hormone
(TRH)
is
a
hypothalamic
peptide
which
acts
on
the
anterior
pituitary
gland
to
induce
release
of
thyrotropin
and
prolactin.
The
TRH
receptor
(TRH-R)
is
a
member
of
the
G-protein-coupled
seven-transmembrane-
spanning
family
of
receptors
(GPCRs)
[1-3].
TRH-R
activation
causes
rapid
stimulation
of
the
phospholipase
C
(PLC)
pathway
via
its
Gq/Gll
G-protein,
producing
Ins(1,4,5)P3
[4,5]
and
1,2-
diacylglycerol.
Ins(l,4,5)P3
causes
a
transient
mobilization
of
Ca2+
from
internal
[6]
and
possibly
indirectly
from
external
[7]
sources,
whereas
diacylglycerol
increases
protein
phosphoryl-
ation
by
activating
protein
kinase
C
[8].
Given
that
changes
in
intracellular
Ca2+
concentration
([Ca2+],)
are
critical
in
the
regulation
of
a
variety
of
cellular
processes,
including
hormone
secretion,
it
is
important
to
understand
the
mechanisms
con-
trolling
these
events
in
different
cell
types.
TRH-induced
changes
in
[Ca2+],
have
been
measured
in
pituitary
and
non-pituitary
cell
types
expressing
TRH-Rs
[9-16].
These
spectrofluorimetric
studies
have
contributed
to
the
analysis
of
TRH-induced
changes
in
[Ca2+]1.
Interpretation
of
these
data
can,
however,
be
misleading,
as
these
studies
were
often
con-
ducted
in
heterogeneous
primary
pituitary
cultures
and/or
sus-
pension
cultures
of
established
pituitary
cell lines.
It
is
generally
believed
that
TRH
causes
a
biphasic
Ca2+
response
but
the
exact
origins
of
the
Ca2`
pools
involved
are
unclear.
Indeed
it
has
been
suggested
that
the
Ca2+
pools
mobilized
after
TRH-R
activation
are
cell-type
specific
[13].
In
pituitary
cell
types,
the
initial
Ca2+
spike
predominantly
involves
the
mobilization
of
Ins(1,4,5)P3-
HEK-293
cells
also
failed
to
respond
to
KCI
or
the
slow
Ca2+-
channel
activator
BAY
K
8644,
suggesting
that
they
lack
L-type
VOCCs.
Rat
and
human
TRH-Rs
are
highly
conserved
except
at
the
C-terminus
where
the
sequence
differs.
The
C-terminus
is
believed
to
be
important
in
receptor
desensitization.
Despite
differences
in
this
region,
rat
and
human
TRH-Rs
expressed
in
HEK-293
cells
underwent
rapid
(within
1
min)
desensitization.
This
desensitization
was
dose-dependent
and
did
not
involve
receptor
loss.
Similarly
the
bradykinin
receptor
endogenous
to
HEK-293
cells
also
displays
a
rapid
desensitization.
We
conclude
that
in
TRH-R-expressing
non-pituitary
HEK-293
cells,
TRH
mobilizes
intracellular
Ca2+
resulting
in
a
monophasic
Ca2+
transient.
The
rat
and
human
TRH-Rs
as
well
as
the
endogenous
bradykinin
receptor
also
displayed
rapid
receptor
desensitization.
sensitive
Ca2+
stores
and
the
longer-lasting
secondary
plateau
phase
relies
on
the
influx
of
extracellular
Ca2
.
In
contrast,
when
expressed
in
cells
of
non-pituitary
origin,
TRH
induces
a
monophasic
Ca2+
response
[13,16].
Receptor
activation
can
lead
not
only
to
stimulation
of
intracellular
events,
for
example
the
mobilization
of
Ca2+,
but
also
to
desensitization
of
these
responses.
Receptor
desensiti-
zation
is
potentially
a
physiologically
important
process,
as
it
provides
a
means
of
regulating
continuous
receptor
stimulation
[17].
Chronic
exposure
of
cells
to
TRH
results
in
the
internaliza-
tion
and
subsequent
loss
of
cell-surface
TRH-Rs
[18-20],
a
decrease
in
the
levels
of
TRH-R
mRNA
[21,22]
and
a
down-
regulation
of
the
Gq
and
Gil
G-proteins
[23].
In
contrast,
rapid
receptor
desensitization
does
not
involve
receptor
loss
but
rather
receptor
phosphorylation.
This
phosphorylation
results
in
an
uncoupling
of
the
receptor
from
its
cognate
G-protein
and
a
loss
of
subsequent
downstream
events.
Although
rapid
desensiti-
zation
of
a
number
of
GPCRs
has
been
demonstrated
[6,17,24],
it
remains
unclear
whether
or
not
the
G-protein-coupled
events
of
the
TRH-R
can
be
acutely
desensitized,
i.e.
within
the
first
few
minutes
of
receptor
activation.
The
present
study
aims
to
eliminate
the
problems
of
cell
heterogeneity
by
directly
monitoring
and
characterizing
TRH-
induced
changes
in
[Ca2+],
in
single
cells
of
TRH-R-expressing
clonal
non-pituitary
cell
lines.
We
demonstrate
that
TRH
causes
a
prompt
dose-dependent
monophasic
rise
in
[Ca2+],
in
HEK-293
cells
expressing
TRH-Rs.
By
using
a
variety
of
intra-
and
extra-
cellular
Ca2+
antagonists,
we
have
also
shown
that
this
response
does
not
involve
the
influx
of
extracellular
Ca2+
through
either
Abbreviations
used:
TRH,
thyrotropin-releasing
hormone;
TRH-R,
TRH
receptor;
BK,
bradykinin;
BK-R,
bradykinin
receptor;
[Ca2+]i,
intracellular
Ca2+
concentration;
VOCCs,
voltage-operated
Ca2+
channels;
SMOCCs,
second-messenger-operated
Ca2+
channels;
HEK-293,
human
embryonal
kidney
293;
TG,
thapsigargin;
CPZ
cyclopiazonic
acid;
GPCRs,
G-protein-coupled
receptors;
PLC,
phospholipase
C;
GTP[S],
guanosine
5'-[y-
thio]triphosphate;
Thi,
thienyl(alanine).
*
To
whom
correspondence
should
be
addressed.
385
386
L.
Anderson
and
others
voltage-
or
second-messenger-operated
Ca2+
channels
(VOCCs
or
SMOCCs),
but
rather
the
mobilization
of
Ca2+
from
thapsigargin/cyclopiazonic
acid
(TG/CPZ)-sensitive
intracellu-
lar
Ca2+
stores.
The
structure
of
the
TRH-R
is
highly
homologous
between
rat
[3],
mouse
[1]
and
human
[25]
except
for
the
stretch
of
amino
acids
at
the
extreme
end
of
the
C-terminus.
The
C-
termini
of
GPCRs
are
thought
to
be
important
in
the
events
associated
with
receptor
desensitization.
We
therefore
compared
acute
desensitization
events
in
cells
expressing
either
rat
or
human
TRH-Rs.
Desensitization
of
the
endogenously
expressed
bradykinin
(BK)
receptor
(BK-R)
was
also
examined.
TRH-Rs
from
both
species
as
well
as
the
endogenous
BK-R
exhibited
a
rapid
desensitization
of
the
agonist-induced
rises
in
[Ca2+]1.
MATERIALS
AND
METHODS
Materials
Tissue
culture
reagents,
media,
lipofectin
and
genetecin
were
supplied
by
Gibco,
Paisley,
Scotland,
U.K.
Glass
coverslips
(22
mm
x
0.175
mm)
were
supplied
by
Arnold
R.
Horwell
Ltd.,
West
Hampstead,
London,
U.K.
[3-Me-His2]TRH
and
guanosine
5'-[y-[35S]thio]triphosphate
([35S]GTP[S]
1245
Ci/mmol)
were
obtained
from
Dupont-NEN
Products,
Hertfordshire,
U.K.
All
other
drugs,
including
fura
2
penta-acetoxymethyl
ester
(fura
2/AM),
were
obtained
from
Calbiochem.
The
Bio-Rad
protein
assay
kit
was
obtained
from
Bio-Rad,
Richmond,
CA,
U.S.A.
Tissue
culture
Cell
lines
were
maintained
routinely
in
Dulbecco's
modified
Eagle's
medium
containing
10%
(v/v)
heat-inactivated
foetal
calf
serum,
glutamine
(0.3
mg/ml),
penicillin
(100
units/ml),
streptomycin
(100
units/ml)
and
geneticin
(800
mg/ml)
and
incubated
at
37
°C
in
a
humidified
atmosphere
of
5
%
(v/v)
CO2
in
air.
Cell
lines
expressing
TRH-Rs
Using
a
stable
transfection
protocol
[23],
the
full-length
rat
[3]
or
human
[25]
TRH-R
was
subcloned
into
the
eukaryotic
expression
vectors
pcDNA1
and
pcDNA3
respectively
and
expressed
in
HEK-293
cells.
Receptor-containing
clones
were
identified
using
a
functional
total
inositol
phosphate
assay
[26]
and
verified
with
a
TRH-R-binding
assay.
The
cell
lines
expanded
from
these
clones
containing
either
the
rat
(293-E2
cell
line)
or
the
human
(293-hlO
cell
line)
TRH-Rs
were
then
used
for
further
study.
Measurement
of
[Ca2+],
Trypsin-treated
single
cells
were
plated
on
to
sterilized
glass
coverslips.
After
2
days,
attached
cells
were
washed
(x
2)
with
buffer
A
[127
mM
NaCl,
5
mM
KCI,
2
mM
MgCl2,
0.5
mM
NaH2PO4,
5
mM
NaHCO3,
1.8
mM
CaC12,
10
mM
Hepes
and
0.1
%
(w/v)
BSA,
pH
7.2].
Cells
were
loaded
in
this
buffer
with
fura
2/AM
(4,uM
final
concentration)
for
30
min
at
37
°C
in
a
5
%
(v/v)
CO2
humidified
incubator.
Unincorporated
dye
was
removed
by
washing
(
x
3)
with
buffer
A.
Coverslips
were
then
transferred
to
a
heated
stage
(37
°C)
of
an
inverted
Nikon
Diaphot
epifluorescence
microscope
with
a
x
40
oil
immersion
objective.
Cells
were
incubated
in
a
fixed
volume
of
buffer
A
(1
ml).
Drug
solutions
(5
ml)
were
added
directly
into
the
coverslip
chamber
and
1
ml
volumes
automatically
obtained
through
suction.
Dynamic
video
imaging
of
changes
in
[Ca2+]2
was
carried
out
in
single
cells
using
the
MagiCal
hardware
and
Tardis
software
Dukesway,
Team
Valley,
Gateshead,
Tyne
and
Weir,
U.K.
as
previously
described
[27].
Fluorescent
images
were
obtained
by
exposing
cells
to
filtered
340
and
380
nm
light
alternated
under
computer
control.
The
image
viewed
at
a
wavelength
of
510
nm
was
focused
on
to
the
face
of
an
intensified
charge-coupled
device
camera
(Photonic
Sciences).
Typically,
eight
images
were
averaged
at
each
wavelength
and
a
similar
number
were
collected
for
background
images
which
were
subsequently
subtracted
on
a
pixel-by-pixel
basis
from
the
imaged
samples.
These
images
were
held
in
memory
for
subsequent
processing
and
analysis.
Fluorescence
excitation
shifts
occur
when
fura
2
binds
Ca2+,
i.e.
the
excitation
efficiency
increases
at
340
nm
and
decreases
at
380
nm.
Ratios
of
values
obtained
at
340/380
nm
therefore
represent
changes
in
[Ca2+]1.
The
340/380
nm
ratio
was
calculated
on
averaged
video
frames
on
a
pixel-by-pixel
basis,
and
this
was
proportional
to
[Ca2+],.
During
non-imaging
periods
in
desensitization
experiments,
photobleaching
of
the
fura
2
dye
was
minimized
by
inserting
a
low-percentage
neutral-density
filter
between
the
light
source
and
the
filter
wheel.
Data
analysis
and
presentation
Software-based
image
analysis
allowed
quantification
of
[Ca2+]1
in
single
cells
versus
time.
ASCII
files
of
these
quantitative
data
were
used
to
derive
plots
of
the
mean
[Ca2+],
versus
time
for
single
cells
from
different
experiments.
A
minimum
of
ten
cells
was
analysed
from
each
experiment,
and
individual
treatment
regimes
were
carried
out
at
least
three
times.
Results
are
given
as
means
+
S.E.M.
Statistical
analysis
was
performed
using
Student's
t
test.
TRH-R-binding
assay
Monolayer
cultures
were
washed
with
PBS
(
x
2)
and
harvested
by
scraping.
The
cells
were
resuspended
in
20
mM
Tris/
HCl/2
mM
MgCl2,
pH
7.4,
and
membranes
prepared
after
homogenization
and
centrifugation
at
20000
g
for
30
min
at
4
°C.
The
membrane
pellet
was
resuspended
in
buffer
B
con-
taining
40
mM
Tris/HCl,
pH
7.4,
and
2
mM
MgCl2.
Ligand-
binding
assays
were
carried
out
with
[3H]
[3-Me-His2]TRH
in
buffer
B
(0.5
ml)
and
various
concentrations
of
unlabelled
peptide.
After
incubation
on
ice
for
I
h,
the
membranes
were
filtered
through
Whatman
GF-B
filters
and
washed
with
buffer
B
(
x
3).
All
assays
were
performed
in
triplicate.
Binding
parameters
were
determined
using
Scatchard
analysis.
Protein
concentrations
were
measured
using
a
Bio-Rad
protein
assay
kit
with
a
BSA
standard.
Binding
of
[NS]GTP[S]
to
membranes
[35S]GTP[S]
binding
to
membranes
prepared
from
293-E2
cells
was
measured
as
previously
described
[28].
Briefly,
reconstituted
293-E2
membranes
were
incubated
in
a
final
assay
volume
of
250,1u
in
buffer
C
[50
mM
Tris/HCl,
2
mM
EDTA,
10
mM
MgCl2,
2
mM
dithiothreitol,
200
mM
NaCl,
0.4
units/ml
adenosine
deaminase,
1
%
(v/v)
BSA]
containing
[35S]GTP[S]
(0.2
nM)
and
GDP
(10
,uM)
at
25
°C
for
45
min.
Samples
were
then
filtered
under
vacuum
through
Whatman
GF-B
filters
presoaked
with
buffer
D
(50
mM
Tris/HCl,
pH
7.4/5
mM
MgCl2).
Filters
were
washed
(
x
2)
with
4
ml
of
buffer
D,
digested
with formic
acid
and
counted
in
Emulsifier
SAFE
scintillation
fluid.
Non-specific
binding
was
determined
in
the
presence
of
provided
by
Applied
Imaging
(formerly
Joyce
Loebl.
Ltd.),
10#M
unlabelled
GTP[S].
Desensitization
of
thyrotropin-releasing
hormone
receptors
in
HEK-293
cells
0.1
nM
I;
1
nM
0.1
PM
400
1
0
120
240 360
480
600
0
120
lime
(s)
240
360
0
120
240
360
Figure
1
Effect
of
various
TRH
concentratons
(10
pM-1
pM)
on
[Ca2+],
in
293-E2
cells
Graphs
represent
an
average
plot
of
[Ca2+]j
measurements
versus
time
(s)
in
a
minimum
of
ten
cells
from
representative
experiments.
TRH
added
at
t=
50
s
produced
a
monophasic
rise
in
[Ca2+]i.
P
<
0.001,
compared
with
control
response
(n
=
10).
RESULTS
Effect
of
TRH
on
[Ca2+],
In
HEK-293
cells
expressing
the
rat
TRH-R
(293-E2
cells)
Figure
1
shows
examples
of
averaged
traces
from
single
experi-
ments
consisting
of
at
least
ten
individual
293-E2
cells
illustrating
TRH-induced
changes
in
[Ca2+],.
The
application
of
various
concentrations
of
TRH
(10
pM-l
,uM)
to
293-E2
cells
produced
dose-dependent
changes
in
[Ca2+]1.
TRH
produced
a
rapid
transient
monophasic
increase
in
[Ca2+1],
and
maximal
responses
were
obtained
with
TRH
concentrations
of
10
nM
and
higher.
Within
10
s
of
the
application
of
1
,uM
TRH,
[Ca2+],
had
risen
from
49.4
+
2.7
nM
at
t
=
50
s
to
348.7
+
25.6
nM
at
t
=
57
s.
The
maximal
concentration
of
1
,#M
TRH
was
used
in
subsequent
experiments.
[Ca2+],
oscillations
were
also
often
observed
after
the
exposure
of
cells
to
concentrations
of
TRH
ranging
from
0.1
nM
to
1
,uM
(Figure
2).
The
origin
of
the
Ca2+
mobilized
by
TRH
was
investigated
using
a
variety
of
intra-
and
extra-cellular
Ca2+
antagonists.
Involvement
of
extracellular
Ca2+
In
the
TRH-induced
[Ca2+J,
response
In
293-E2
cells
Pretreatment
of
293-E2
cells
with
Ca2+-free
buffer
A
containing
the
Ca2+
chelator
EGTA
(2
mM;
6
min)
had
no
effect
on
the
[Ca2+],
response
to
TRH
(Figure
3a).
Similar
treatment
of
cells
400
-
300
+
~200
-I
X
/
j
0
0
120
240
360
480
600 720
840
960
Time
(s)
Figure
2
TRH-induced
(Ca2+1,
oscillations
In
293-E2
cells
This
trace
illustrates
a
typical
example
of
a
single
TRH-treated
(1
,tM;
t=
108
s)
oscillating
cell.
with
the
L-type
VOCC
blocker,
nifedipine
(1
1tM),
did
not
alter
the
response
to
TRH
(Figure
3b).
After
the
application
of
the
SMOCC
blocker,
SK&F
96365
(1
,tM),
293-E2
cells
responded
normally
to
TRH
(Figure
3c).
Collectively
these
data
suggest
10
pM
400
-
300-
200
100
*
i
-)
-W
0
10
nm
300-
200
100
0
387
388
L.
Anderson
and
others
SK&F
96365
TRH
(d)
TRH
(c)
720
0
120
240
360
480
600
720
TRH
(e)
Time
(s)
Figure
3
Effect
of
(a)
EGTA,
(b)
niedipine,
(c)
SK&F
96365,
(d)
TG
and
(e)
CPZ
pretreatment
on
the
TRH-induced
[Ca2+],
response
in
293-E2
cells
(a)
Cells
were
preincubated
for
6 min
in
Ca2+-free
buffer
A
with
2
mM
EGTA, and
TRH
(1
1sM;
t=
420
s)
was
added.
(b)
293-E2
cells
were
pretreated
with
nifedipine
(NIF;
1
,uM;
t=
108
s),
and
TRH
(1
ttM)
was
subsequently
added
at
t
=
420
s.
(c)
293-E2
cells
were
pretreated
with
SK&F
96365
(1
,uM;
t
=
10
s),
and
TRH
(1
1sM;
t
=
420
s)
was
then
added.
(d)
293-E2
cells
were
treated
with
TG
(1
,uM;
t=
108
s),
and
TRH
(1
,uM;
t=
688
s)
was
then
added.
(e)
293-E2
cells
were
pretreated
with
CPZ
(1
,uM;
t=
108
s),
and
TRH
(1
,uM;
t=
533
s)
was
then
added.
that
mobilization
of
extracellular
Ca21
is
not
involved
in
the
TRH-induced
Ca2+
response.
Neither
did
[Ca2+]1
increase
after
treatment
of
cells
with
the
L-type
VOCC
activator
BAY
K8644
or
depolarization
with
KCl
(2-50
mM;
results
not
shown).
Although
[Ca2+]
appeared
not
to
return
to
prestimulated
values
after
the
application
of
TRH
to
either
nifedipine-
or
SK&F
96365-pretreated
cells,
these
values
were
in
fact
not
significantly
higher
than
basal
prestimulated
[Ca2+].
Involvement
of
intracellular
Ca2+
in
the
TRH-induced
Ca2+
response
in
293-E2
cells
The
importance
of
intracellular
Ca2+
mobilization
in
the
TRH-
induced
Ca2+
response
was
investigated
using
the
structurally
unrelated
ATPase
inhibitors
TG
and
CPZ.
Both
compounds
are
thought
to
deplete
Ins(1,4,5)PJ-sensitive
Ca2+
stores
in
the
endoplasmic
reticulum.
Both
compounds
(1
#tM)
produced
a
slow
but
substantial
increase
in
[Ca2+]1
(Figures
3d
and
3e).
Although
the
subsequent
addition
of
TRH
to
these
cells
produced
a
temporary
fall
in
[Ca2+]1,
it
did
not,
however,
produce
the
expected
rise
in
[Ca2+]1
(Figures
3d
and
3e).
These
results
imply
that,
in
293-E2
cells,
the
mobilization
of
Ca2+
from
intracellular
pools
is
responsible
for
the
TRH-induced
Ca2+
response.
The
slight
decrease
in
[Ca2+]1
after
the
application
of
TRH
to
TG-
or
CPZ-pretreated
cells
is
consistent
with
the
idea
that
receptor
stimulation
can
lead
to
activation
of
Ca2+
pumping
from
cells,
an
effect
masked
under
normal
circumstances
[29].
Alternatively,
dye
compartmentalization,
which
can
occur
at
37
°C,
may
also
explain
this
observation.
Involvement
of
G-protein
coupling
in
the
TRH-induced
intracellular
Ca2+
response
Suramin
sodium
has
previously
been
shown
to
uncouple
G-
proteins
from
their
receptors.
The
effect
of
this
compound
on
Gq/G,1-linked
TRH
Ca2+
mobilization
was
examined
(Figure
4a).
Pretreatment
of
cells
with
suramin
sodium
(10
,uM)
signifi-
cantly
reduced
the
Ca2+
response
to
TRH.
[Ca2+],
rose
from
29.4+
8.6
nM
to
155.2+14.9
nM
at
t
=
422
s
compared
with
a
control
response
of
329.7
+
25.6
nM.
Despite
a
dose-dependent
reduction
in
total
binding,
neither
TRH-R
affinity
or
number
were
altered
by
treatment
with
suramin
sodium
(1-100
lM;
results
not
shown).
Suramin
did,
however,
abolish
both
TRH-
and
[3-Me-His2]TRH-stimulated
[35S]GTP[S]
binding
in
mem-
branes
prepared
from
293-E2
cells
(Figure
4b).
400-
300
200
100
i
-S
co
L0
0
400
TG
300-
200
100-
n
0
120
240 360
480
600
720
840 960
Desensitization
of
thyrotropin-releasing
hormone
receptors
in
HEK-293
cells
400
1
(a)
SS
TRH
2.0
0
E
n
1.8
n
0
.0
1.6-
U)
_
1.4J
, .
0
120
240
360
480
600
720
Time
(s)
(b)
TRH
[3-Me-His2I-TRH
cc
Ir
IX
)
2>
CC
ccC
L4
L
+
E e
+
U~~~~~
~~~~
a
Figure
4
Effect
of
suramin
sodium
pretreatment
on
(a)
TRH-induced
Ca2+
response
In
293-E2
cells
and
(b)
TRH-
and
[3Me-HisITRH-simulated
[S]GTP[SJ
binding
In
cell
membranes
prepared
from
293-E2
cells
(a)
Suramin
sodium
(SS)
pretreatment
(10
1sM;
t
=
108
s)
significantly
reduced
the
TRH-induced
[Ca2+]i
response
(1
FM;
t
=
422
s);
P
=
0.02,
compared
with
control
response
(n
=
5).
(b)
SS
(32
FM)
treatment
abolished
both
TRH
(1
/FM)-
and
[3-Me-His2]TRH
(1
FM)-stimulted
[35S]GTP[S]
binding
in
cell
membranes
prepared
from
293-E2
cells.
Results
are
representative
of
those
obtained
in
three
separate
experiments.
DesensitIzation
of
the
TRH-
and
BK-Induced
Intracellular
Ca2+
response
Exposure
of
293-E2
cells
to
a
1
min
pulse
of
TRH
caused
the
expected
transient
increase
in
[Ca2l],
(Figure
5a).
Subsequent
exposure
of
cells
to
further
1
min
pulses
of
TRH
at
various
intervals
(t
=
1000,
2400
and
3600
s;
Figure
5a)
failed
to
elicit
any
further
response.
Similar
results
were
obtained
when
cells
were
continuously
exposed
to
TRH
(results
not
shown).
This
lack
of
response
appeared
not
to
be
related
to
cell
viability.
Using
the
same
experimental
design,
cells
that
initially
showed
no
response
to
buffer
subsequently
responded
normally
to
TRH
(Figure
5b).
This
acute
desensitization
could
not
be
explained
by
changes
in
receptor
number
(Bmax
13.5
pM/mg
of
protein)
or
affinity
(Kd
4
nM)
after
TRH
exposure,
as
these
parameters
were
unaltered
in
control
buffer-treated
cells
compared
with
cells
treated
with
TRH.
The
desensitization
of
the
TRH-induced
Ca2+
response
also
appears
to
be
a
dose-related
phenomenon.
Low-
dose
TRH
pretreatment
(0.1
nM;
Table
1)
partially
desensitized
the
Ca2+
response
to
submaximal
concentrations
of
TRH
(0.1
nM-0.1
uM),
whereas
the
Ca2+
response
to
1
,uM
TRH
was
unaltered.
In
contrast,
pretreatment
with
higher
concentrations
of
TRH
(1
nM,
5
nM,
0.1
M)
completely
desensitized
cells
(results
not
shown).
Treatment
of
293-E2
cells
with
BK
(1
,uM;
Figure
6a)
caused
[Ca2+],
to
rise
from
8.5
+
1.3
at
t
=
47
s
to
145
+
17
nM
at
t
=
59
s.
This
response
was
monophasic
in
nature
and
resembled
the
low-
dose
TRH
responses
observed
in
this
cell
line.
Pretreatment
of
cells
with
a
potent
BK-R
antagonist
(sodium
adamantanecetyl-
D-Arg-[Hyp3,Thi5'8,D-Phe7]bradykinin)
abolished
the
BK-
but
not
the
TRH-induced
mobilization
of
intracellular
Ca2+
(results
not
shown).
As
the
characteristics
of
desensitization,
i.e.
the
rapidity,
extent
and
duration
of
desensitization,
appear
to
be
receptor-specific,
we
also
examined
desensitization
of
the
BK-
induced
[Ca2+],
response
in
293-E2
cells.
After
initial
treatment
with
BK
(1
,uM;
t
=
47
s;
Figure
6a),
cells
subsequently
failed
to
respond
to
BK
(1
,uM;
t
=
2195
s)
but
continued
to
respond
to
TRH
(1
,uM;
t
=
2500
s;
Figure
6a).
So,
although
the
BK-
induced
Ca2+
response
was
smaller
than
that
induced
by
TRH,
BK-R
desensitization
nevertheless
occurred.
These
results
further
imply
that
Ca2+
depletion
is
not
primarily
responsible
for
the
desensitization
phenomenon
observed
in
the
present
experiments.
(a)
TRH
400
-
300
-
200
100*
i
TRH
TRH
TRH
v
4-
c
(b)
Buffer
400
300
-
200
-
100
TRH
0
1000
2000
Time
(s)
L
3000
Figure
5
Effect
of
(a)
several
pulses
of
TRH
and
(b)
buffer
A
and
TRH
on
[Ca2+J,
In
293-E2
cells
(a)
293-E2
cells
initially
challenged
with
TRH
(1
FM;
t
=
98
s)
responded
with
an
increase
in
[Ca2+],
but
failed
to
respond
to
subsequent
regular
TRH
challenges
(1
FM;
t
=
1000,
2500
and
3600
s).
(b)
Application
of
buffer
A
(5
ml;
t
=
100
s)
had
no
effect
whereas
TRH
(1
FuM;
t=
2250
s)
produced
the
expected
increase
in
[Ca2+]i.
300
200
a
2
100*
A
-
4
I
I
I
6...-
m
%#P-
n
v
I
Li
389
m.'
.
390
L.
Anderson
and
others
When
cells
were
pretreated
with
TRH
(1
,#M;
t
=
47
s;
Figure
6b),
no
intracellular
Ca2+
mobilization
was
observed
after
the
application
of
either
BK
(1
,uM;
t
=
2230
s)
or
TRH
(1
sM;
t
=
2430
s).
Similar
results
were
observed
in
293-h20
cells
ex-
pressing
the
human
TRH-R
(Figures
6c
and
6d),
suggesting
that
this
pattern
of
desensitization
is
not
species-specific.
Again
this
desensitization
could
not
be
accounted
for
by
a
reduction
in
receptor
number
(Bmax
0.52
pM/mg
of
protein;
Kd
4
nM).
In
293-E2
cells
this
effect
was
dose-dependent,
as
pretreatment
of
cells
with
lower
TRH
concentrations
(0.1
nM;
t
=
110
s;
Figure
Table
1
DesensfflzatIon
of
the
TRH-induced
Ca2+
response
In
293-E2
cells
after
pretreatment
wIth
low-dose
TRH
293-E2
cells
initially
treated
with
TRH
(0.1
nM)
were
subsequently
challenged
with
various
concentrations
of
TRH
(0.1
nM-1
,uM).
Compared
with
control
responses,
low-dose
TRH
pretreatment
desensitized
subsequent
[Ca2+]
responses
to
low-dose
but
not
to
high-dose
TRH
treatment;
*P
<
0.05,
**P
<
0.001
(n
=
3).
[Ca2+]1
response
in
293-E2
cells
Control
[Ca2+];
after
pretreatment
response
in
293-E2
with
0.1
nM
TRH
[TRH]
cells
(nM) (nM)
0.1
nM
10
nM
0.1
,M
1
,uM
178
+
6
265
+121
242
±293
352
+144
103
+3*
141
+50*
139
+
42**
292
+89
6e)
only
partially
desensitized
the
[Ca2l],
response
to
BK
(1
#etM;
t
=
2125
s).
These
cells
were
then
treated
with
Ca2+-free
buffer
A
containing
EGTA
(2
mM;
5
min).
The
subsequent
addition
of
1
1tM
ionomycin
(Ca2+-free
salt)
produced
a
rapid
rise
in
[Ca2+]i,
suggesting
that
intracellular
Ca2+
stores
had
not
been
depleted.
DISCUSSION
Time-related
changes
in
[Ca2+]1
in
single
HEK-293
cells
expressing
either
the
rat
or
human
TRH-Rs
were
characterized.
Although
HEK-293
cells
do
not
normally
express
TRH-R,
this
cell
line
has
been
commonly
used
for
the
expression
of
a
variety
of
recently
cloned
GPCRs
[30,31].
The
application
of
TRH
produced
a
prompt
dose-dependent
monophasic
increase
in
[Ca2+]1
in
all
imaged
TRH-R-expressing
cells.
As
the
influx
of
extracellular
Ca2+
through
either
SMOCCs
or
L-type
VOCCs
were
not
involved
in
this
TRH-induced
Ca2+
response,
the
involvement
of
intracellular
Ca2+
was
examined.
Treatment
of
293-E2
cells
with
either
of
the
Ca2+-ATPase
inhibitors
TG
and
CPZ
resulted
in
a
transient
but
pronounced
rise
in
[Ca2+]1.
The
subsequent
addition
of
TRH
failed
to
provoke
a
Ca2+
response,
suggesting
that
Ins(1,4,5)P3-sensitive
Ca2+
stores
normally
released
during
TRH-
R
activation
are
absent
and
that
TRH-induced
Ca2+
mobilization
is
dependent
on
these
intracellular
stores.
These
data
are
in
accord
with
the
findings
of
a
recent
study
that
showed
that,
in
cell
lines
lacking
L-type
VOCC
activity
(for
example
HeLa
or
C6
cells),
TRH
similarly
generated
a
monophasic
Ca2+
response
which
was
also
blocked
by
TG
[13].
Physiological
concentrations
of
Ca2+-mobilizing
hormones
(b
BK
TRH
TRI
1 3
1000
2000
3000
0
(c)
BK
TRH
BK
1000
2000 3000
0
BK
TRH
1000
2000
3000
(e)
TRH
TRH
BK
TRH
BK
ION
10,
10
'0.
0
II
0
1000
2000
3000
0
1000
2000
3000
Time
(s)
Figure
6
Effect
of
BK
on
[Ca2+],
response
(a)
Effect
of
BK
pretreatment
on
[Ca2+]i
response
to
subsequent
BK
and
TRH
challenge
in
293-E2
cells.
Although the
application
of
BK
(1
,uM;
t
=
47
s)
produced
a
transient
rise
in
[Ca2+],
it
also
completely
desensitized
the
cells
to
the
effect
of
BK
(1
#M;
t
=
2195
s)
but
only
partially
desensitized
cells
to
the
effects
of
TRH
(1
1zM;
t
=
2500
s).
(b)
TRH
pretreatment
(1
FuM;
t
=
47
s)
abolished
[Ca2+]i
responses
to
both
BK
(1
FM;
t
=
2230
s)
and
TRH
(1
,M;
t
=
2430
s)
in
293-E2
cells.
(c)
Effect
of
BK
pretreatment
on
[Ca2+]1
response
to
subsequent
BK
and
TRH
challenge
in
293-h20
cells
expressing
the
human
TRH-R.
Although
the
application
of
TRH
(1
FM;
t
=
47
s)
produced
a
transient
rise
in
[Ca2+]i,
it
markedly
desensitized
cells
to
the
effect
of
BK
(1
FM;
t=
2080
s)
but
not
TRH
(1
FM;
t
=
2263
s).
(d)
Effect
of
TRH
pretreatment
on
[Ca2+]i
responses
to
TRH
and
BK
in
293-h20
cells.
TRH
pretreatment
(1
F%M;
t
=
47
s)
abolished
subsequent
effects
of
TRH
(1
uM;
t
=
2286
s)
and
BK
(1
FM;
t
=
2523
s)
on[Ca2+]i.
(e)
Low-dose
TRH
pretreatment
(0.1
nM;
t
=
110
s)
partially
desensitized
the
[Ca2+]i
response
to
BK
(1
FM;
t=
2125
s).
Ca2+-free
buffer
A
containing
EGTA
(2
mM;
5
min)
was
then
added,
followed
by
1
FM
ionomycin
(ION).
(a)
BK
I
.i
30u
300-
(d)
200
100
0
2
C4
C)
401
30
20
10
_
anti
I
Desensitization
of
thyrotropin-releasing
hormone
receptors
in
HEK-293
cells
have been
shown
to
produce
repetitive
and
periodic
Ca2l
spikes/oscillations.
These
oscillations
are
believed
to
play
an
important
role
in
cellular
signal-transduction
processes
[15,161.
In
the
present
study,
TRH-induced
Ca2l
oscillations
were
occasionally
observed.
As
these
oscillations
were
an
apparently
random
occurrence
in
only
a
small
proportion
of
the
cells
imaged,
we
were
unable
to
investigate
the
source
of
Ca2+
involved
in
these
responses.
Similar
observations
have,
however,
been
demonstrated
in
HeLaR
cells
transfected
with
the
mouse
TRH-
R
[13].
These
responses
were
independent
of
extracellular
Ca2+
and
more
likely
arose
from
Ca2+
release
and
reuptake
at
intracellular
Ca2+
storage
pools.
Indeed
the
Ca21
oscillations/
spikes
observed
in
both
this
and
the
present
study
are
charac-
teristic
of
those
produced
by
the
cytoplasmic
oscillator
[32]
which
are
typically
insensitive
to
extracellular
Ca2
.
TRH-R
belongs
to
a
family
of
hormone
and
neurotransmitter
receptors
whose
actions
are
mediated
via
the
activation
of
G-proteins.
Suramin
sodium
has
previously
been
shown
to
antagonize
the
interaction
between
receptors
and
the
G-proteins
that
regulate
adenylate
cyclase
[33,34]
or
PLC
activity
[35,36].
It
is
believed
to
intercalate
directly
with
the
receptor-G-protein
complex
thereby
forming
an
a-helix
which
in
some
way
uncouples
the
receptor
from
its
G-protein.
In
the
present
study
the
partial
inhibition
by
suramin
of
the
TRH-induced
rise
in
[Ca2+]1
could
reflect
an
interaction
of
the
compound
with
either
the
TRH-R
or
its
associated
G-protein(s).
Owing
to
the
lack
of
a
specific
TRH-
R
antagonist,
the
effects
of
suramin
on
TRH
binding
and
TRH-
stimulated
[35S]GTP[S]
binding
were
investigated
instead.
The
dose-dependent
inhibition
of
both
TRH
and
[3-Me-His2]TRH-
induced
[35S]GTP[S]
binding
in
293-E2
cells
by
suramin
sodium
without
altering
either
TRH-R
affinity
or
number
suggests
that
suramin
acts
directly
at
the
site
of
receptor-G-protein
coupling
to
prevent
any
interaction
between
these
two
proteins.
Similar
data
have
been
reported
for
the
polyanionic
compound
L-
451,167
in
CHO
cells
expressing
the
cc2-adrenoceptor
[34].
Some
controversy
surrounds
the
desensitization
of
the
G-
protein-coupled
events
of
the
TRH-R.
Whereas
some
groups
have
shown
that
TRH
can
maintain
second-messenger
pro-
duction
at
a
constant
level
for
periods
of
up
to
1
h
[20,37],
others
have
shown
a
marked
reduction
in
the
rate
of
second-messenger
production
within
minutes
of
exposure
to
the
peptide
[4,38,39].
The
methodology
employed
to
measure
the
events
of
receptor
desensitization,
experimental
design,
data
interpretation
and
varying
cell
types
may
all
contribute
to
these
conflicting
results.
To
assess
desensitization
more
directly
at
the
level
of
the
single
cell,
we
examined
the
effects
of
TRH
pretreatment
on
the
TRH-
induced
Ca2+
response
in
293-E2
cells.
After
either
a
1
min
pulse
or
continuous
exposure
of
cells
to
a
high
concentration
of
TRH,
293-E2
cells
at
first
responded
with
the
expected
monophasic
rise
in
[Ca2+]i.
Subsequent
exposure
of
these
cells
to
TRH
at
various
intervals
for
a
peri-od
of
up
to
1
h
failed
to
promote
Ca2+
mobilization.
Desensitization
of
the
TRH-induced
Ca2+
response
was
dose-dependent,
as
low-dose
TRH
pretreatment
reduced,
whereas
high-dose
pretreatment
completely
abolished,
Ca2+
responses
to
subsequent
TRH
treatment.
Although
these
data
suggest
that
TRH-Rs
expressed
in
HEK-293
cells
undergo
rapid
desensitization,
this
desensitization
could
be
attributed
to
the
depletion
of
internal
Ca2+
stores.
We
therefore
challenged
TRH-
treated
cells
with
ionomycin
in
buffer
containing
EGTA.
Under
these
conditions
ionomycin,
but
not
TRH,
was
able
to
mobilize
intracellular
Ca2+.
These
results
indicate
that
intracellular
Ca2+
stores
have
not
been
totally
depleted.
It
should
also
be
noted
that
fura-2-labelled
cells,
which
showed
no
initial
response
to
buffer,
were
capable
of
responding
normally
to
TRH
for
up
to
1
h.
This
ability
could
account
for
the
TRH-induced
desensitization
observed
in
the
present
study.
Although
in
general
the
rat,
human
and
mouse
TRH-Rs
show
high
sequence
homology,
considerable
sequence
variation
exists
at
the
C-terminus
of
these
receptors
[25].
Compared
with
the
human
TRH-R,
the
long
form
of
the
rat
TRH-R
has
an
extended
terminus
containing
an
extra
13
amino
acids
including
an
additional
potential
phosphorylation
site.
The
terminal
region
of
a
variety
of
receptors
including
TRH-R
have
been
associated
with
receptor-G-protein-coupled
events
and
receptor
desen-
sitization
[40].
We
therefore
examined
whether
the
truncated
C-terminus
of
human
TRH-R
altered
the
ability
of
the
receptor
to
desensitize.
Despite
sequence
differences,
the
TRH-induced
Ca2+
response
in
293-h20
cells
expressing
human
TRH-R
was
similarly
desensitized.
Interestingly,
in
a
recent
study
by
Pedersen
et
al.
[39],
desensitization
of
mouse
TRH-R
was
shown
to
be
cell-type-
specific.
Although
the
internalization
and
eventual
loss
of
TRH-Rs
are
associated
with
long-term
exposure
to
TRH
[18-20],
the
rapid
desensitization
observed
in
the
present
study
cannot
be
accounted
for
by
receptor
loss,
as
receptor
number
and
affinity
were
unaltered
after
the
acute
exposure
of
cells
to
TRH.
A
possible
explanation
for
this
rapid
desensitization
is
receptor
phosphoryl-
ation.
The
occupation
of
the
receptor
with
its
agonist
is
believed
to
induce
receptor
phosphorylation
at
multiple
sites
resulting
in
an
uncoupling
of
the
receptor
from
its
cognate
G-protein.
This
is
certainly
the
case
for
the
widely
studied
,-adrenoceptor
[41].
Agonist-dependent
phosphorylation
of
the
PLC-linked
mus-
carinic
M3
[42],
substance
P
[43]
and
cholecystokinin
[44]
receptors
has
also
been
demonstrated.
The
time
course
measured
for
this
receptor
phosphorylation
is
indeed
rapid
enough
to
account
for
the
acute
desensitization
of
the
TRH-induced
Ca2+
response.
Although
rapid
receptor
desensitization
appears
to
be
a
feature
common
to
PLC-linked
receptors,
the
level
of
desensitization
has
been
reported
to
vary
from
receptor
to
receptor
[45,46].
BK-Rs
are
expressed
in
normal
rat
kidney
fibroblasts
[47]
and
in
the
kidney-derived
HEK-293
cells.
In
293-E2
and
293-hlO
cells,
BK
produced
a
transient
monophasic
rise
in
[Ca2+],.
This
appears
to
be
a
receptor-specific
effect,
as
a
potent
BK-R
antagonist
abolished
the
BK-
but
not
the
TRH-induced
mobilization
of
intracellular
Ca2
.
We
subsequently
carried
out
a
series
of
experiments
comparing
the
level
and
patterns
of
desensitization
of
the
endogenously
expressed
BK-R
and
exogenously
expressed
TRH-R
in
293-E2
cells.
High-dose
TRH
treatment
of
cells
resulted
in
desensitization
of
both
the
TRH-
and
BK-induced
Ca2+
response.
In
contrast,
high
concentrations
of
BK
desensitized
the
BK-
but
not
the
TRH-induced
Ca2+
response.
Therefore
despite
similar
intracellular
signalling
pathway,
it
initially
appears
that
these
two
receptors
have
different
desensitization
mechanisms.
However,
it
is
also
possible
that
this
might
simply
reflect
differences
in
the
efficiency
of
receptor-G-
protein
coupling.
It
has
been
suggested
that
there
is
a
regulatory
negative
feedback
loop
between
cytosolic
Ca2+
and
PLC
activity
[48].
The
mobilization
of
intracellular
Ca2+
could
therefore
influence
rapid
desensitization
by
virtue
of
its
ability
to
feedback
negatively
on
PLC
activity.
Perhaps
the
extensive
Ca2+
mobilization
induced
by
TRH
results
in
a
marked
negative
regulatory
effect
on
PLC
activity
thereby
desensitizing
the
responses
to
both
TRH
and
BK.
Compared
with
TRH,
BK
is
less
effective
in
mobilizing
intracellular
Ca2+
and
may
therefore
have
a
smaller
negative
regulatory
effect
on
PLC
activity
such
t-hat
the
subsequent
exposure
to
a
more
vigorous
stimulus,
i.e.
TRH,
is
still
capable
of
mobilizing
Ca2+.
It
has
been
proposed
that
within
any
given
cell
there
is
a
limited
pool
of
G-proteins
391
demonstrates
that
neither
dye-quenching
nor
reduced
cell
vi-
392
L.
Anderson
and
others
shared
by
a
variety
of
endogenous
receptors.
It
is
possible
that
TRH-R
stimulation
and
subsequent
receptor
uncoupling
limits
the
G-protein
pool
available
to
couple
to
other
endogenously
expressed
receptors,
in
this
case
the
BK-R.
The
problems
often
associated
with
the
measurement
of
agonist-induced
changes
in
[Ca2+]1
in
heterogeneous
populations
of
cells
were
eliminated
in
the
present
study
by
combining
dynamic
video
imaging
with
fluorescence
microscopy
to
measure
acute
time-related
changes
in
[Ca2+],
in
single
HEK-293
cells
expressing
either
the
rat
or
human
TRH-Rs.
These
cells
are
non-
pituitary
in
origin
and
apparently
lack
L-type
VOCC
activity.
TRH
produces
a
prompt
dose-dependent
rise
in
[Ca2+]1
in
these
cells,
an
effect
solely
dependent
on
the
mobilization
of
TG/CPZ-
sensitive
intracellular
Ca2+
stores.
We
have
also
clearly
demon-
strated
a
rapid
desensitization
of
the
rat
and
human
TRH-Rs
as
well
as
the
endogeneously
expressed
BK-R.
We
thank
Professor
D.
W.
Lincoln
for
support,
J.
Zabavnik
for
generating
the
293-
E2
cell
line
and
Dr.
P.
L.
Taylor,
J.
V.
Cooke
and
A.
McGregor
for
assistance
with
TRH-R-binding
assays.
REFERENCES
1
Straub,
R.
E.,
Frech,
G.
C.,
Joho,
R.
H.
and
Gershengorn,
M.
C.
(1990)
Proc.
Natl.
Acad.
Sci.
U.S.A.
87,
9514-9518
2
Probst,
W.
C.,
Snyder,
L.
A.,
Schuster,
D.
I.,
Brosius,
J.
and
Sealfon,
S.
C.
(1992)
Mol.
Cell.
Biol.
11,
1-20
3
Sellar,
R.
E.,
Taylor,
P.
L.,
Lamb,
R.
F.,
Zabavnik,
J.,
Anderson,
L.
and
Eidne,
K.
A.
(1993)
J.
Mol.
Endocrinol.
10,199-206
4
Torjesen,
P.
A.,
Bjoro,
T.,
Ostberg,
B.
C.
and
Haug,
E.
(1988)
Mol.
Cell.
Endocrinol.
56,107-114
5
Hsieh,
K.
P.
and
Martin,
T.
F.
J.
(1992)
Mol.
Endocrinol.
6,
1673-1681
6
Richardson,
R.
M.,
Kim,
C.,
Benovic,
J.
L.
and
Hosey,
M. M.
(1993)
J.
Biol.
Chem.
268,
13650-13656
7
Bird,
G.,
Rossier,
M.
F.,
Hughes,
A.
R.,
Shears,
S.
B.,
Armstrong,
D.
L.
and
Putney,
J. J.
W.
(1991)
Nature
(London)
352,
162-165
8
Nishizuka,
Y.
(1992)
Science
258,
607-614
9
Winiger,
B. P.
and
Schlegel,
W.
(1988)
Biochem.
J.
255,
161-167
10
Law,
G.
J.,
Pachter,
J.
A.
and
Dannies,
P.
S.
(1989)
Mol.
Endocrinol.
3,
539-546
11
Shorte,
S.
L.
and
Schofield,
J.
G.
(1991)
Mol.
Cell.
Endocrinol.
79,
167-176
12
Chiavaroli,
C.,
Vacher,
P.
and
Schlegel,
W.
(1992)
Eur.
J.
Pharmacol.
227,173-180
13
Peizhi,
L.,
Thaw,
C.
N.,
Sempowski,
G.
D.,
Gershengorn,
M.
C.
and
Hinkle,
P.
M.
(1992)
Mol.
Endocrinol
6,
1393-1402
14
Bauer,
C.
K.,
Davison,
I.,
Kubasov,
I.,
Schwarz,
J.
R.
and
Mason,
W.
T.
(1994)
Pflugers
Arch.
428,17-25
15
Wagner,
K.
A.,
Yacono,
P.
W.,
Golan,
D.
E.
and
Tashjian,
A.
H.
J.
(1993)
Biochem.
J.
292,175-182
16
Nelson,
E.
J.
and
Hinkle,
P.
M.
(1994)
Endocrinology
135,
1084-1092
17
Kobilka,
B.
(1992)
Annu.
Rev.
Neurosci.
15,
87-114
18
Imai,
A.
and
Gershengorn,
M.
C.
(1985)
J.
Biol.
Chem.
260,
10536-10540
19
Winicov,
I.
and
Gershengorn,
M.
C.
(1989)
J.
Biol.
Chem.
264,
9438-9443
20
Cubbitt,
A.
B.,
Geras-Raaka,
E.
and
Gershengorn,
M.
C.
(1990)
Biochem.
J.
271,
331-336
21
Fujimoto,
J.,
Narayanan,
C.
S.,
Benjamin,
J.
E.
and
Gershengorn,
M.
C.
(1992)
Endocrinology
131,
1716-1720
22
Yamada,
M.,
Monden,
T.,
Satoh,
T.
et
al.
(1992)
Biochem.
Biophys.
Res.
Commun.
184,
367-372
23
Kim,
G.
D.,
Carr,
C.,
Anderson,
L.,
Zabavnik,
J.,
Eidne,
K.
A.
and
Milligan,
G.
(1994)
J.
Biol.
Chem.
269,19933-19940
24
Hawes,
B.
E.
and
Conn,
P.
M.
(1992)
Endocrinology
131,
2681-2690
25
Duthie,
S.
M.,
Taylor,
P.
L.,
Anderson,
L.,
Cook,
J.
V.
and
Eidne,
K.
A.
(1993)
Mol.
Cell.
Endocrinol.
95,
R11-R15
26
Anderson,
L.,
Milligan,
G.
and
Eidne,
K.
A.
(1992)
J.
Endocrinol.
136,
51-58
27
Anderson,
L.,
Hoyland,
J.,
Mason,
W.
T.
and
Eidne,
K.
A.
(1992)
Mol.
Cell.
Endocrinol.
86,
167-175
28
Lorenzen,
A.,
Fuss,
M.,
Vogt,
H.
and
Schwabe,
U.
(1993)
Mol.
Pharmacol.
U,
115-123
29
Zhang,
B.-X.,
Zhao,
H.,
Loessberg,
P.
A.
and
Muallem,
S.
(1992)
J.
Biol.
Chem.
267,
15419-15424
30
Lakkakorpi,
J.
T.
and
Rajaniemi,
H.
J.
(1994)
Mol.
Cell.
Endocrinol.
99,
39-47
31
Nagayama,
Y.,
Chazenbalk,
G.
D.,
Takeshita,
A.
et
al.
(1994)
Endocrinology
135,
1060-1065
32
Berridge,
M.
J.
(1990)
J.
Biol.
Chem.
265,
9583-9586
33
Butler,
S.
J.,
Kelly,
E.,
McKenzie,
F.,
Guild,
S.,
Wakelam,
M.
and
Milligan,
G.
(1988)
Biochem.
J.
251,
201-205
34
Huang,
R.
R.
C.,
Dehaven,
R.
N.,
Cheung,
A.
H.,
Diehl,
R.
E.,
Dixon,
R.
A.
F.
and
Strader,
C.
D.
(1989)
Mol.
Pharmacol.
37,
304-310
35
Kopp,
R.
and
Pfeiffer,
A.
(1990)
Cancer
Res.
50,
6490-6496
36
Dasso,
L.
L.
T.
and
Taylor,
C.
W.
(1991)
Biochem.
J.
280,
791-795
37
Drummond,
A.
H.,
Bushfield,
M.
and
Macphee,
C.
H.
(1984)
Mol.
Pharmacol.
25,
201-208
38
Perlman,
J.
H.
and
Gershengorn,
M.
C.
(1991)
Endocrinology
129,
2679-2686
39
Pedersen,
E.,
Heinflink,
M.,
Alvira,
M.,
Nussenzveig,
D.
R.
and
Gershengorn,
M.
C.
(1994)
Mol.
Pharmacol.
45,
684-689
40
Nussenzveig,
D.
R.,
Heinflink,
M.
and
Gershengorn,
M.
C.
(1993)
J.
Biol.
Chem.
268,
2389-2392
41
Hausdorff,
W.
P.,
Bouvier,
M.,
O'Dowd,
B.
F.,
Irons,
G.
P.,
Caron,
M.
G.
and
Lefkowitz,
R.
J.
(1989)
J.
Biol.
Chem.
21,
12657-12665
42
Tobin,
A.
B.
and
Nahorski,
S.
R.
(1993)
J.
Biol.
Chem.
268,
9817-9823
43
Kwatra,
M.
M.,
Schwinn,
D.
A.,
Schreurs,
J.
et
al.
(1993)
J.
Biol.
Chem.
268,
9161-9164
44
Klueppelberg,
U.
G.,
Gates,
L.
K.,
Gorelick,
F.
S.
and
Miller,
L.
J.
(1991)
J.
Biol.
Chem.
266,
2403-2408
45
Stauderman,
K.
A.
and
Pruss,
R.
M.
(1990)
J.
Neurochem.
54,
946-953
46
Menniti,
F.
S.,
Takemura,
H.,
Oliver,
K.
G.
and
Putney,
J.
W.
(1991)
Mol.
Pharmacol.
40,
727-733
47
Van
Zoelen,
E.
J.
J.,
Peters,
P.
H.
J.,
Afink,
G.
B. et
al.
(1994)
Biochem.
J.
298,
335-340
48
Stauderman,
K.
A.
and
Pruss,
R.
M.
(1990)
J.
Neurochem.
54,
946-953
Received
11
November
1994/30
May
1995;
accepted
8 June
1995
... As such, the cell line of choice was a compromise. HEK293 were shown to express the neurofilament subunits NF-L, NF-M, NF-H, and α-internexin [50], endogenous β2-adrenergic receptors and βadrenergic receptor kinase 1 [51,52], sphingosine-1-phosphate receptors, somatostatin receptor subtype SSTR2, P2Y1, and P2Y2 receptors [53,54], and thyrotropin releasing hormone receptor [55], all of which are expressed in neuronal cells. Although derived from kidney tissue, HEK293 cells exhibited characteristics of immature neurons. ...
Article
Full-text available
Seizure threshold 2 (SZT2) is a component of the KICSTOR complex which, under catabolic conditions, functions as a negative regulator in the amino acid-sensing branch of mTORC1. Mutations in this gene cause a severe neurodevelopmental and epileptic encephalopathy whose main symptoms include epilepsy, intellectual disability, and macrocephaly. As SZT2 remains one of the least characterized regulators of mTORC1, in this work we performed a systematic interactome analysis under catabolic and anabolic conditions. Besides numerous mTORC1 and AMPK signaling components, we identified clusters of proteins related to autophagy, ciliogenesis regulation, neurogenesis, and neurodegenerative processes. Moreover, analysis of SZT2 ablated cells revealed increased mTORC1 signaling activation that could be reversed by Rapamycin or Torin treatments. Strikingly, SZT2 KO cells also exhibited higher levels of autophagic components, independent of the physiological conditions tested. These results are consistent with our interactome data, in which we detected an enriched pool of selective autophagy receptors/regulators. Moreover, preliminary analyses indicated that SZT2 alters ciliogenesis. Overall, the data presented form the basis to comprehensively investigate the physiological functions of SZT2 that could explain major molecular events in the pathophysiology of developmental and epileptic encephalopathy in patients with SZT2 mutations.
... Chromatin immunoprecipitation (ChIP) assays were used to assess FOXO3a recruitment to Egr1 and LHβ promoters. FOXO3a targets Egr1 expression to, at least in part, indirectly regulate LHβ promoter activity.These findings suggest that GnRH regulates LHβ-subunit expression through one or more FOXO3a-mediated mechanisms.HEK293 embryonic kidney cells (originally obtained from ATCC) stably expressing the rat Type I GnRH-R (designated SCL60 (Stable Cell Lines)) were generated within our laboratory(Anderson et al., 1995). Cells were maintained in complete medium (as described in the table below) supplemented with G418 (Sigma) (unless otherwise specified, all materials utilised in this study, were supplied from Sigma) and plasmocin (all cell lines were grown at 37 o C in a humidified 5% CO2 atmosphere unless otherwise indicated). ...
Thesis
G protein-coupled receptors (GPCRs) are a large family of trans-membrane receptors that transmit signals from extracellular stimuli to target intracellular signal transduction pathways. The gonadotropin-releasing hormone receptor (GnRH-R) is a GPCR which binds the decapeptide GnRH. In the pituitary gonadotrope, GnRH stimulates gonadotropin (LH and FSH) biosynthesis and secretion to regulate reproduction. GnRH and the GnRH-Rs are also present in many extra-pituitary tissues, although their role at these sites remains largely undetermined. GnRH-Rs are known to recruit a diverse array of signalling pathway mediators in different cell-types. These include; Gq/11-PLCβ-IP3/DAG-Ca2+/PKC signalling, monomeric G-proteins and integrins to mediate cell adhesion and migration, the activation of the major members of the mitogen-activated protein kinase (MAPK) super-family (extracellular signal-regulated kinase (ERK), c-Jun N-terminal Kinase (JNK) and p38MAPK), and β-catenin and other mediators of the canonical Wnt signalling pathway. This thesis describes the regulation of Forkhead Box O (FOXO) transcription factors by GnRH. The mammalian FOXO transcription factors, FOXO1, FOXO3a and FOXO4, are emerging as an important family of proteins that modulate the expression of genes involved in cell-cycle regulation, induction of apoptosis, DNA damage repair and response to oxidative stress. In this thesis, emphasis is placed on delineating the novel role of FOXO transcription factors in mediating two important and widely-researched areas of GnRH biology. Firstly, the role of FOXO transcription factors in mediating cell-growth inhibition in response to GnRH treatment is assessed in a heterologous HEK293/GnRH-R expressing cell line. Secondly, the role of transcription factors in regulating luteinising hormone-β (LHβ)-subunit expression is investigated in the LβT2 gonadotrope cell line. Activation of the GnRH-R can inhibit cell proliferation and induce apoptosis in certain tumour-derived cell lines. Several studies have reported that these events can occur as a result of changes in the expression profiles of specific cell-cycle regulatory and apoptotic genes, many of which are FOXO-target genes, including GADD45, FasL, p21Cip1 and p27Kip1. In this thesis, a role for FOXOs in targeting the expression of several of these genes in response to GnRH is assessed, highlighting a specific role for FOXO3a in mediating GADD45 and FasL expression. The signalling mechanisms through which FOXO3a regulates GADD45 expression in response to GnRH is also described. Finally, a stable FOXO3a-knock-down cell line was generated in order to further examine FOXO3a involvement in GnRH-induced cell-growth inhibition. GnRH is an essential regulator of the reproductive process by stimulating the synthesis of LH and FSH in pituitary gonadotropes, thereby regulating gametogenesis and steroidogenesis. Diverse signalling pathways have been reported to regulate LHβ-subunit expression in response to GnRH, including the ERK/JNK/p38MAPK cascades and factors such as Egr1, SF1 and β-catenin. In the second part of this thesis, the role of FOXOs in regulating LHβ-subunit expression in response to GnRH is described. The data presented suggests that GnRH can regulate LHβ-subunit expression through both indirect and direct FOXO3a-mediated mechanisms. Firstly, FOXO3a was found to regulate Egr1 expression to indirectly target LHβ-promoter activity. Secondly, a role for β-catenin as a FOXO3a co-factor to directly regulate LHβ-subunit expression, together with Egr1 and SF1, is also proposed. FOXO3a expression and sub-cellular localisation was assessed and demonstrated in LβT2 cells and in adult human male pituitary sections. The research presented in this thesis adds to the diversity of signalling pathways and mediators that GnRH can target in different cellular backgrounds in order to mediate a variety of cellular processes. The antiproliferative and apoptotic effects of GnRH on tumour-derived cell lines are well-documented, and this research highlights a novel role for FOXO3a in mediating these events. The regulation of gonadotropin synthesis remains an important topic of research, and the novel implication of FOXO3a in mediating LHβ-subunit expression adds further complexity to gonadotrope physiology.
... For these experiments we utilized the HEK293 cell line, which was derived from transformed primary cultures of human embryonic kidney cells with sheared adenovirus (Ad)5 DNA, for its high transfection efficiency and ability to robustly express GFP from our constructs. Although kidney derived cell lines are not generally considered to be optimal model systems for the study of neurological disorders, the HEK293 cell line exhibits many common genetic markers and phenotypic resemblance with early differentiating neurons or neuronal stem cells [108][109][110][111][112][113][114][115][116]. Beyond this, neuronal cells are more prone to transformation by adenovirus and it is thought that a neuronal cell may have been present in the original culture from whence the HEK293 cell line was derived [108]. ...
Article
Full-text available
Thyrotropin-releasing hormone (TRH) is an important endocrine agent that regulates the function of cells in the anterior pituitary and the central and peripheral nervous systems. By controlling the synthesis and release of thyroid hormones, TRH affects many physiological functions, including energy homeostasis. This hormone exerts its effects through G protein-coupled TRH receptors, which signal primarily through Gq/11 but may also utilize other G protein classes under certain conditions. Because of the potential therapeutic benefit, considerable attention has been devoted to the synthesis of new TRH analogs that may have some advantageous properties compared with TRH. In this context, it may be interesting to consider the phenomenon of biased agonism and signaling at the TRH receptor. This possibility is supported by some recent findings. Although knowledge about the mechanisms of TRH receptor-mediated signaling has increased steadily over the past decades, there are still many unanswered questions, particularly about the molecular details of post-receptor signaling. In this review, we summarize what has been learned to date about TRH receptor-mediated signaling, including some previously undiscussed information, and point to future directions in TRH research that may offer new insights into the molecular mechanisms of TRH receptor-triggered actions and possible ways to modulate TRH receptor-mediated signaling.
Article
The HEK-293 cell line was created in 1977 by transformation of primary human embryonic kidney cells with sheared adenovirus type 5 DNA. A previous study determined that the HEK-293 cells have neuronal markers rather than kidney markers. In this study, we tested the hypothesis whether Zika virus (ZIKV), a neurotropic virus, is able to infect and replicate in the HEK-293 cells. We show that the HEK-293 cells infected with ZIKV support viral replication as shown by indirect immunofluorescence (IFA) and quantitative reverse transcriptase-PCR (qRT-PCR). We performed RNA-seq analysis on the ZIKV-infected and the control uninfected HEK-293 cells and find 659 genes that are differentially transcribed in ZIKV-infected HEK-293 cells as compared to uninfected cells. The results show that the top 10 differentially transcribed and upregulated genes are involved in antiviral and inflammatory responses. Seven upregulated genes, IFNL1, DDX58, CXCL10, ISG15, KCNJ15, IFNIH1, and IFIT2, were validated by qRT-PCR. Altogether, our findings show that ZIKV infection alters host gene expression by affecting their antiviral and inflammatory responses.
Article
Full-text available
This study quantified the agonist-induced endocytotic and recycling events of the mammalian gonadotropin releasing hormone receptor (GnRH-R) and investigated the role of the intracellular carboxyl (C)-terminal tail in regulating agonist-induced receptor internalization kinetics. The rate of internalization for the rat GnRH-R was found to be exceptionally low when compared with G-protein coupled receptors (GPCRs) which possess a cytoplasmic C-terminal tail (thyrotropin-releasing hormone receptor (TRH-R), catfish GnRH-R (cfGnRH-R) and GnRH/TRH-R chimeric receptor). These data provide evidence that the presence of a functional intracellular cytoplasmic C-terminal tail is essential for rapid internalization of the studied GPCRs.
Article
Full-text available
This study examined the mechanism underlying the rat GnRH receptor (GnRH-R) internalization pathway by investigating the role of added/extended C-terminal tails and the effect of b-arrestins and dynamin. The internalization of the wild-type (WT) rat GnRH-R, stop codon mutants, GnRH-R/TRH receptor (TRH-R) chimera, rat TRH-R, and catfish GnRH-R was examined using radioligand binding assay. Overexpression of b-arrestin in COS-7 cells expressing each of the receptor constructs substantially increased endocytosis rate con- stants (ke) of the TRH-R, catfish GnRH-R, and GnRH-R/TRH-R chi- mera, but not of the WT rat GnRH-R and stop codon mutants. The b-arrestin-promoted increase in the ke value was diminished by co- transfecting cells with the dominant negative b-arrestin-(319 - 418) mutant, whereas WT GnRH-R and stop codon mutant internalization were unaffected. Additionally, confocal microscopy showed that ac- tivated GnRH-Rs failed to induce time-dependent redistribution of either b-arrestin-1- or b-arrestin-2-green fluorescent protein conju- gate to the plasma membrane. However, the dominant negative dy- namin (DynK44A) mutant impaired internalization of all of the re- ceptors regardless of their b-arrestin dependency, indicating that they internalize via a clathrin-mediated pathway. We conclude that the mammalian GnRH-R uses a b-arrestin-independent, dynamin- dependent internalization mechanism distinct from that employed by the other receptors studied. (Endocrinology 141: 299 -306, 2000)
Article
This study examined the mechanism underlying the rat GnRH receptor (GnRH-R) internalization pathway by investigating the role of added/extended C-terminal tails and the effect of beta-arrestins and dynamin. The internalization of the wild-type (WT) rat GnRH-R, stop codon mutants, GnRH-R/TRH receptor (TRH-R) chimera, rat TRH-R, and catfish GnRH-R was examined using radioligand binding assay. Overexpression of beta-arrestin in COS-7 cells expressing each of the receptor constructs substantially increased endocytosis rate constants (k(e)) of the TRH-R, catfish GnRH-R, and GnRH-R/TRH-R chimera, but not of the WT rat GnRH-R and stop codon mutants. The beta-arrestin-promoted increase in the k(e) value was diminished by cotransfecting cells with the dominant negative beta-arrestin-(319-418) mutant, whereas WT GnRH-R and stop codon mutant internalization were unaffected. Additionally, confocal microscopy showed that activated GnRH-Rs failed to induce time-dependent redistribution of either beta-arrestin-1- or beta-arrestin-2-green fluorescent protein conjugate to the plasma membrane. However, the dominant negative dynamin (DynK44A) mutant impaired internalization of all of the receptors regardless of their beta-arrestin dependency, indicating that they internalize via a clathrin-mediated pathway. We conclude that the mammalian GnRH-R uses a beta-arrestin-independent, dynamin-dependent internalization mechanism distinct from that employed by the other receptors studied.
Article
Thesis (Ph.D.)--University of Rochester. School of Medicine and Dentistry. Dept. of Pharmacology and Physiology, 2008. G protein-coupled receptors (GPCRs) regulate numerous signaling pathways. Their activity is tightly regulated through phosphorylation at multiple sites. These studies were performed to elucidate the mechanisms and consequences of thyrotropin-releasing hormone (TRH) receptor phosphorylation. Phosphorylation was measured by mobility shift on SDS-PAGE, incorporation of 32P-orthophosphate, and binding of phosphosite-specific polyclonal antibodies. Receptor activation caused translocation of endogenous G protein-coupled receptor kinase-2 (GRK2) to the plasma membrane. Over-expression of dominant negative GRK2 or small interfering RNA against GRK2 blocked receptor phosphorylation. The major phosphorylation site of receptor endogenously expressed in pituitary GH3 cells was Thr-365 in the receptor tail. Phosphorylation peaked within seconds of adding TRH, was TRH concentration-dependent, and was required for arrestin binding. Regulated receptor dimerization increased receptor phosphorylation. Neither G protein signaling nor activation of protein kinase C was required for receptor phosphorylation or dephosphorylation. Receptors were dephosphorylated after TRH was removed. Internalized receptors dephosphorylated slower than receptors on the plasma membrane. Receptors on the plasma membrane could be rephosphorylated in response to TRH after agonist removal and dephosphorylation. Phosphorylated receptor staining was visible in prolactin- and thyrotropin-producing cells in rat pituitary tissue from animals injected with TRH. The results show that receptor dephosphorylation can take place on the plasma membrane, that receptor can rapidly cycle between a phosphorylated and nonphosphorylated state in response to changing agonist concentrations, and that phosphorylation can be used as an indicator of receptor activity in vivo. Arrestins are cytosolic proteins that bind activated, phosphorylated GPCRs and inhibit G protein signaling and increase receptor internalization. The importance of arrestin was studied by co-expressing TRH receptor with or without arrestin in fibroblasts from mice lacking arrestin 2 and/or arrestin 3. Affinity of receptor for TRH was increased several-fold in cells expressing arrestin. Arrestin was required for TRH-induced receptor endocytosis and rapid and sustained desensitization. Arrestin did not promote internalization or desensitization of a receptor that had phosphosites mutated to Ala (4Ala receptor), but did increase 4Ala receptor affinity for TRH and resistance to acid/salt washout of TRH. Internalization was not required for arrestin-dependent high affinity TRH binding, acid/salt resistance, and desensitization. A sterically restricted arrestin mutant did not cause receptor internalization or desensitization, but did promote acid/salt resistance and high agonist affinity. The results demonstrate that arrestin binds two different sites in the receptor tail and arrestin binding at either site causes increased agonist affinity and acid/salt resistance, but only the proximal phosphosites evoke the necessary conformational changes in arrestin for receptor desensitization and internalization. Rab GTPases regulate intracellular membrane trafficking to distinct compartments. Trafficking of phosphorylated TRH receptors with fluorescent protein-tagged Rabs and arrestin was measured in HEK293 cells. Phosphorylated receptor recruited arrestin to the plasma membrane and internalized via Rab5. Phospho-receptor moved primarily into endosomes containing both Rab4 and Rab5. Very little internalized receptor co-localized with Rab11 and none with Rab7. Wildtype or dominant negative Rab over-expression did not affect phosphorylation or dephosphorylation. Dephosphorylated receptor co-localized exclusively with Rab4. Dephosphorylated receptors did not recycle and the plasma membrane. The results show that the phospho-state of the receptor determines its subcellular trafficking and that dephosphorylated receptors do not immediately recycle, suggesting that the plasma membrane is repopulated by a naïve intracellular pool.
Article
TRH acts on specific G-protein coupled receptors sited in cells of the anterior pituitary gland. Pituitary tumours expressing either TSH, PRL or GH may respond to TRH by enhanced, blunted or paradoxical hormone release. Non-functioning pituitary tumours may also show abnormal responses to TRH. Little is understood of the mechanisms regulating inappropriate hormone release in these tumours. Activating or inactivating mutations found in G-protein coupled receptors have been implicated in human pathological conditions. Mutations in the G-protein coupled TRH receptor might be involved in the aetiology of pituitary adenomas resulting in aberrant hormone secretion. We therefore screened samples of pituitary adenomas for the presence of somatic mutations in the TRH receptor gene. Pituitary adenoma tissue samples were obtained at surgery from 50 patients with pituitary adenoma (17 acromegaly, 15 prolactinoma, 11 TSH-secreting and 7 non-functioning adenoma) along with blood samples to provide lymphocyte DNA as control sequence. Genomic DNA was extracted from adenoma and lymphocyte samples and the entire coding region of the TRH receptor was amplified using 5 overlapping pairs of PCR primers. The PCR products were analysed for mutations by non-denaturing polyacrylamide gel electrophoresis which reveals single-strand conformational polymorphisms (SSCP) as a mobility shift in product migration. Wild-type and mutant TRH receptor cDNA were similarly analysed to confirm the sensitivity of the method. Additionally, PCR products were ligated into a PCR cloning vector and DNA sequencing carried out to confirm the findings of SSCP analysis. The human TRH receptor retained normal wild-type sequence in the large group of TSH secreting, PRL secreting, GH secreting and non-functioning pituitary adenomas investigated in this study. Our observations suggest that the TRH receptor structure is normal in TSH secreting, PRL secreting, GH secreting and non-functioning pituitary adenomas. It is therefore unlikely that the TRH receptor is involved in the pathology associated with the types of pituitary adenomas investigated in this study. It is possible that some other component of the pathway controlling TRH-signalling events may be implicated in pituitary tumorigenesis.
Article
Full-text available
The multitude of G-protein coupled receptor (GPR) superfamily cDNAs recently isolated has exceeded the number of receptor subtypes anticipated by pharmacological studies. Analysis of the sequence similarities and unique features of the members of this family is valuable for designing strategies to isolate related cDNAs, for developing hypotheses concerning substrate-ligand and receptor-effector interactions, and for understanding the evolution of these genes. We have compiled and aligned the 74 unique amino acid sequences published to date and review the present understanding of the structural motifs contributing to ligand binding and G-protein coupling.
Article
Full-text available
The role of internal stores and plasma membrane Ca2+ pumps in controlling [Ca2+]i during agonist stimulation and their regulation by agonists are not well understood. We report here measurements of intracellular ([Ca2+]i) and extracellular ([Ca2+]o) Ca2+ concentrations in agonist-stimulated pancreatic acini in an effort to directly address these questions. Stimulation of acini suspended in Ca(2+)-free or Ca(2+)-containing medium with Ca2+ mobilizing agonists resulted in a typical transient increase in [Ca2+]i. Thapsigargin, a specific inhibitor of internal Ca2+ pumps, inhibited the rate of [Ca2+]i reduction after agonist stimulation by approximately 40%. Under the same conditions, thapsigargin had no effect on the rate of the unidirectional Ca2+ efflux across the plasma membrane as revealed by measurements of [Ca2+]o. These findings suggest that internal Ca2+ pumps actively remove Ca2+ from the cytosol during continued agonist stimulation. The correlation between the reduction in [Ca2+]i and the increase in [Ca2+]o showed that Ca2+ efflux from cells stimulated with agonist and thapsigargin represent Ca2+ efflux across the plasma membrane. Inhibition of cells exposed to agonist and thapsigargin with a specific antagonist sharply reduced the rates of the [Ca2+]i decrease and the accompanied [Ca2+]o increase. Hence, at comparable [Ca2+]i, Ca2+ efflux from stimulated cells was about 3-fold faster than that from resting cells, indicating that agonists directly activate the plasma membrane Ca2+ pump. To study the role of [Ca2+]i increase in plasma membrane Ca2+ pump activation the acini were loaded with 1,2-bis-(2-aminophenoxyethane-N,N,N',N')-tetraacetic acid (BAPTA), and [Ca2+]o was measured during agonist stimulation. Surprisingly, although BAPTA completely prevented the increase in [Ca2+]i, Ca2+ efflux rate was reduced by only 34%. These findings provide the first evidence for Ca(2+)-independent activation of the plasma membrane Ca2+ pump by Ca2+ mobilizing agonists.
Article
After initial GnRH pretreatment (10 nM, 5 h), subsequent GnRH-stimulated LH release from the gonadotrope was diminished (1 microM GnRH stimulated release of 36.4 +/- 1.4% total cellular LH over 3 h in cells initially pretreated with medium alone compared to 27.4 +/- 1.2% in GnRH-pretreated cells); however, inositol phosphate (IP) production in response to the releasing hormone remained unaffected (1 microM GnRH provoked IP accumulation of 161 +/- 9% above basal levels after 45 min in control cells and 162 +/- 11% in GnRH-pretreated cells). Pretreatment of pituitary cell cultures with NaF (a guanyl nucleotide binding protein activator, 10 mM, 3 h) also decreased subsequent GnRH-stimulated LH release, and in addition, provoked a decrease in GnRH receptor number, an increase in GnRH receptor affinity, reduction of GnRH-stimulated IP production to basal levels, and an increase in the amount of LH released in response to stimulation with the calcium ionophore A23187. In order to determine if the changes in LH ...
Article
It is well known that the TSH receptor (TSHR) undergoes homologous desensitization. That is, prolonged stimulation of thyroid cells with TSH attenuates the cAMP response to subsequent TSH stimulation. However, the existence of homologous desensitization of the recombinant TSHR expressed in nonthyroidal eukaryotic cells is controversial. In the present studies, therefore, we first investigated whether or not the TSHR was desensitized by TSH in 293 human embryonal kidney cells, a cell line in which the LH/CG receptor (LH/CGR) is reported to undergo homologous desensitization. The wild type (wt) TSHR and the wt-LH/CGR stably expressed in 293 cells bound to their respective hormones with high affinity and produced a dose-dependent intracellular cAMP response to hormone stimulation. Pretreatment of cells expressing the TSHR or the LH/CGR with their respective hormones attenuated the cAMP response to subsequent hormone stimulation without down-regulation of the receptors, demonstrating that the TSHR, as well as the LH/CGR, undergoes homologous desensitization in 293 cells. With this cell type expressing mutant TSHRs, we then studied some aspects of the molecular mechanism of TSHR desensitization and compared our data to those obtained with the β-adrenergic receptor (β-AR), which is widely regarded as the prototype for receptor desensitization. We cotransfected the wt-TSHR and a chimeric receptor consisting of the LH/CGR extracellular ligand binding domain with the TSHR transmembrane/cytoplasmic signal transducing region. These two receptors have distinct hormone specificities but share common signal regulatory mechanisms. We observed that, like the β-AR, only hormone-occupied receptor is likely to be involved in homologous desensitization. On the other hand, studies with a truncated TSHR indicated that, in contrast to the β-AR, the serine/threonine-rich region in the carboxyl two thirds of the cytoplasmic tail of the TSHR is not involved in homologous desensitization.
Article
TRH stimulates a biphasic increase in intracellular free calcium ion, [Ca2+]i. Cells stably transfected with TRH receptor cDNA were used to compare the response in lines with and without L type voltage-gated calcium channels. Rat pituitary GH-Y cells that do not normally express TRH receptors, rat glial C6 cells, and human epithelial Hela cells were transfected with mouse TRH receptor cDNA. All lines bound similar amounts of [3H][N3-Me-His2]TRH with identical affinities (dissociation constant = 1.5 nM). Both pituitary lines expressed L type voltage-gated calcium channels; depolarization with high K+ increased 45Ca2+ uptake 20- to 25-fold and [Ca2+]i 12- to 14-fold. C6 and Hela cells, in contrast, appeared to have no L channel activity. GH4C1 cells responded to TRH with a calcium spike (6-fold) followed by a sustained second phase. When TRH was added after 100 nM nimodipine, an L channel blocker, the initial calcium burst was unaffected but the second phase was abolished. GH-Y cells transfected with TRH receptor cDNA responded to TRH with a 6-fold [Ca2+]i spike followed by a plateau phase (>8 min) in which [Ca2+]i remained elevated or increased. Nimodipine did not alter the peak TRH response or resting [Ca2+]i but reduced the sustained phase, which was eliminated by chelation of extracellular Ca2+. In the transfected glial C6 and Hela cells without calcium channels, TRH evoked transient, monophasic 7- to 9-fold increases in [Ca2+]i, and [Ca2+]i returned to resting levels within 3 min. Thapsigargin stimulated a gradual, large increase in [Ca2+]i in transfected C6 cells, and subsequent addition of TRH caused no further rise. Removal of extracellular Ca2+ from transfected C6 cells shortened the [Ca2+]i responses to TRH, to endothelin 1, and to thapsigargin. The TRH responses were pertussis toxin-insensitive. In summary, TRH can generate a calcium spike in pituitary, C6, and Hela cells transfected with TRH receptor cDNA, but the plateau phase of the [Ca2+]i response is not observed when the receptor is expressed in a cell line without L channel activity.
Article
We found previously that the level of endogenous TRH receptor (TRH-R) mRNA in pituitary (GH3) cells and the level of mouse TRH-R mRNA in GH3 cells stably transfected with mouse pituitary TRH-R cDNA are down-regulated by TRH. This down-regulation is caused by TRH stimulation of TRH-R mRNA degradation via a mechanism that appears to involve protein kinase-C. In this report we study regulation of TRH-R mRNA in monkey kidney (COS-1) cells transiently transfected with mouse pituitary TRH-R cDNA. In transfected COS-1 cells, TRH and phorbol 12-myristate 13-acetate (PMA) caused increases in the level of TRH-R mRNA. In contrast, TRH caused only a small transient increase in the level of the mRNA for the neomycin resistance gene, which was cotransfected with TRH-R, and did not affect the level of the mRNA for glyceraldehyde phosphate dehydrogenase, an endogenous gene. The increases in TRH-R mRNA caused by TRH and PMA were inhibited to similar extents by H-7 (1-[5-isoquinolinesulfonyl]2-methyl piperazine dihydrochloride), an inhibitor of protein kinases. The effect of TRH was observed in cells transfected with expression vectors in which TRH-R cDNA was controlled by cytomegalovirus or Rous sarcoma virus promoters. There was no effect of TRH or PMA on the rate of transcription of the transfected TRH-R cDNA. In contrast, TRH caused the rate of degradation of TRH-R mRNA to decrease from 8.0% to 5.1%/h. Hence, TRH, most likely via a protein kinase-C-mediated mechanism, up-regulates TRH-R mRNA levels in transfected COS-1 cells by decreasing the rate of TRH-R mRNA degradation. Since TRH and PMA down-regulate TRH-R mRNA in GH3 cells, posttranscriptional regulation of TRH-R mRNA is a cell-type specific process.
Article
In pituitary cells, electrical activity generates characteristic oscillations of the cytosolic free Ca2+ concentration, [Ca2+]i. These oscillations are controlled by activators as well as by inhibitors of secretion. We studied, in single fura-2-loaded cells, the role of protein phosphorylation in modulating [Ca2+]i oscillations, using either okadaic acid, an inhibitor of protein phosphatases, or activators of protein kinases A and C. Okadaic acid always increased rapidly both the frequency and amplitude of [Ca2+]i oscillations. In contrast, activation of protein kinases A or C generated more complex kinetic [Ca2+]i patterns: phosphorylation due to both kinases resulted in a sustained activation of [Ca2+]i oscillations in about one-third of the cells, whereas two-thirds of the cells responded by an arrest of [Ca2+]i oscillations. This transient phase of arrest was followed, after a few minutes, by a recovery of [Ca2+]i oscillations, often with enhanced frequency. During the arrest, depolarizing the cells with an external microelectrode could not trigger an increase in [Ca2+]i. We conclude that: (i) the fine regulation between phosphorylation/dephosphorylation events is crucial for the modulation of [Ca2+]i oscillations, and (ii) protein kinases A and C can control Ca2+ influx bidirectionally.
Article
The regulation of pituitary hormone secretion by TRH and GnRH proceeds through similar mechanisms which employ phosphoinositide hydrolysis to generate intracellular signals. Proximal events involve receptor activation of heterotrimeric (alpha beta gamma) GTP-binding (G) proteins which regulate phospholipase (PLC) activity. Since TRH and GnRH actions are not affected by cholera or pertussis toxin, a novel G protein (Gp) was suggested to mediate receptor regulation. The required Gp protein has not been identified and this was the focus of the present study. Recent molecular cloning and biochemical studies have characterized two novel, pertussis toxin-insensitive alpha-subunit proteins of the Gq subfamily (alpha q and alpha 11) which regulate the activity of the beta 1 isoenzyme of PLC. Gq and G11 represent the best candidates for the PLC-activating G proteins which mediate the actions of TRH and GnRH. To test this directly, an antibody to the common Gq/11 alpha-subunit carboxyterminal sequence was generated and shown to react with unique 42-kilodalton Gq alpha and 43-kilodalton G11 alpha proteins in membranes from TRH-responsive GH3 cells and GnRH-responsive alpha T3-1 pituitary cells. The Gq/11 alpha peptide antibody was shown to immunodeplete the Gp activity of GH3 cell membrane extracts measured by reconstitution of the guanine nucleotide regulation of PLC-beta 1. In addition, the immunoglobulin G fraction of Gq/11 alpha peptide immune serum specifically inhibited TRH- and GnRH-stimulated PLC activity measured in the membranes of GH3 and alpha T3-1 cells, respectively. The results indicate that TRH and GnRH activation of PLC requires receptor coupling to a Gp protein(s) which corresponds to Gq, G11 or both.
Article
After initial GnRH pretreatment (10 nM, 5 h), subsequent GnRH-stimulated LH release from the gonadotrope was diminished (1 microM GnRH stimulated release of 36.4 +/- 1.4% total cellular LH over 3 h in cells initially pretreated with medium alone compared to 27.4 +/- 1.2% in GnRH-pretreated cells); however, inositol phosphate (IP) production in response to the releasing hormone remained unaffected (1 microM GnRH provoked IP accumulation of 161 +/- 9% above basal levels after 45 min in control cells and 162 +/- 11% in GnRH-pretreated cells). Pretreatment of pituitary cell cultures with NaF (a guanyl nucleotide binding protein activator, 10 mM, 3 h) also decreased subsequent GnRH-stimulated LH release, and in addition, provoked a decrease in GnRH receptor number, an increase in GnRH receptor affinity, reduction of GnRH-stimulated IP production to basal levels, and an increase in the amount of LH released in response to stimulation with the calcium ionophore A23187. In order to determine if the changes in LH release were a result of decreased IP production and/or decreased GnRH receptor binding, the time course of recovery to control levels of these processes was assessed. GnRH receptor binding continued to decrease after NaF pretreatment, reaching a nadir (62% of control) at 6 h after the pretreatment period and recovering at 48 h (90% of control). In contrast, GnRH-provoked IP accumulation did not return to control levels even after 48 h of recovery after NaF pretreatment (1 microM GnRH-stimulated IP accumulation in NaF-pretreated cells was 57% compared to control cells after 48 h of recovery). GnRH-stimulated LH release was inhibited immediately after NaF pretreatment (1 microM GnRH-stimulated LH release in NaF-pretreated cells was 65% of control levels). Cells began to recover within 3 h (80% of control) and were almost completely recovered by 6 h (90% of control). A23187-provoked LH release was enhanced immediately after NaF pretreatment (30 microM A23187-stimulated LH release in NaF-pretreated cells was 170% of control levels). Responsiveness to ionophore was 133% of control by 0.5 h, and complete recovery was measured within 1 h (100% of control). Furthermore, both NaF and GnRH pretreatment still provoked a decrease in gonadotrope responsiveness when IP production was inhibited by the phospholipase C inhibitor U-73122. The results suggest that the development of gonadotrope desensitization (by either NaF or GnRH pretreatment) can be uncoupled from changes in IP production.(ABSTRACT TRUNCATED AT 400 WORDS)