scispace - formally typeset
Open AccessJournal ArticleDOI

Leptin activates anorexigenic POMC neurons through a neural network in the arcuate nucleus

TLDR
An integrated model of leptin action and neuronal architecture in the arcuate nucleus of the hypothalamus is proposed and it is shown that melanocortin peptides have an autoinhibitory effect on this circuit.
Abstract
The administration of leptin to leptin-deficient humans, and the analogous Lepob/Lepob mice, effectively reduces hyperphagia and obesity. But common obesity is associated with elevated leptin, which suggests that obese humans are resistant to this adipocyte hormone. In addition to regulating long-term energy balance, leptin also rapidly affects neuronal activity. Proopiomelanocortin (POMC) and neuropeptide-Y types of neurons in the arcuate nucleus of the hypothalamus are both principal sites of leptin receptor expression and the source of potent neuropeptide modulators, melanocortins and neuropeptide Y, which exert opposing effects on feeding and metabolism. These neurons are therefore ideal for characterizing leptin action and the mechanism of leptin resistance; however, their diffuse distribution makes them difficult to study. Here we report electrophysiological recordings on POMC neurons, which we identified by targeted expression of green fluorescent protein in transgenic mice. Leptin increases the frequency of action potentials in the anorexigenic POMC neurons by two mechanisms: depolarization through a nonspecific cation channel; and reduced inhibition by local orexigenic neuropeptide-Y/GABA (gamma-aminobutyric acid) neurons. Furthermore, we show that melanocortin peptides have an autoinhibitory effect on this circuit. On the basis of our results, we propose an integrated model of leptin action and neuronal architecture in the arcuate nucleus of the hypothalamus.

read more

Content maybe subject to copyright    Report

letters to nature
480 NATURE
|
VOL 411
|
24 MAY 2001
|
www.nature.com
adjusted error rate (
a9) for the six subsequent planned pairwise comparisons was 0.009.
n = 8 PIs for each group.
PIs are distributed normally. Hence, these data were analysed by analysis of variance
with subsequent pairwise comparisons adjusted to maintain an experiment-wise error rate
of
a = 0.05. In Fig. 3, the error rate for the one subsequent planned comparison was 0.05.
In Fig. 4, to maintain an experiment-wise error rate of a = 0.05, the adjusted error rates
(a9) were P = 0.004 and 0.01, respectively, for the 13 subsequent planned pairwise
comparisons in Fig. 4a and for the 5 subsequent planned comparisons in Fig. 4b. The
number of PIs included in each group mean is listed above the corresponding bar in all
®gures.
Received 25 January; accepted 28 March 2001.
1. Scoville, W. B. & Milner, B. Loss of recent memory after bilateral hippocampal lesions.
J. Neuropsychiat. Clin. Neurosci. 12, 103±113 (2000).
2. Milner, B., Squire, L. R. & Kandel, E. R. Cognitive neuroscience and the study of memory. Neuron 20,
445±468 (1998).
3. Mayford, M. et al. Control of memory formation through regulated expression of a CaMKII
transgene. Science 274, 1678±1683 (1996).
4. Shimizu, E., Tang, Y. P., Rampon, C. & Tsien, J. Z. NMDA receptor-dependent synaptic reinforcement
as a crucial process for memory consolidation. Science 290, 1170±1174 (2000).
5. de Belle, J. S. & Heisenberg, M. Associative odor learning in Drosophila abolished by chemical ablation
of mushroom bodies. Science 263, 692±695 (1994).
6. Hammer, M. & Menzel, R. Multiple sites of associative odor learning as revealed by local brain
microinjections of octopamine in honeybees. Learn. Mem. 5, 146±156 (1998).
7. Kitamoto, T. Conditional modi®cation of behavior in Drosophila by targeted expression of a
temperature-sensitive shibire allele in de®ned neurons. J. Neurobiol. 47, 81±92 (2001).
8. Rybak, J. & Menzel, R. Anatomy of the mushroom bodies in the honey bee brain: the neuronal
connections of the a-lobe. J. Comp. Neurol. 334, 444±465 (1993).
9. Strausfeld, N. J. & Li, Y. Organization of olfactory and multimodal afferent neurons supplying the
calyx and pedunculus of the cockroach mushroom bodies. J. Comp. Neurol. 409, 603±625 (1999).
10. Ito, K. et al. The organization of extrinsic neurons and their implications in the functional roles of the
mushroom bodies in Drosophila melanogaster Meigen. Learn. Mem. 5, 52±77 (1998).
11. Connolly, J. B. et al. Associative learning disrupted by impaired Gs signaling in Drosophila mushroom
bodies. Science 274, 2104±2106 (1996).
12. Zars, T., Fischer, M., Schulz, R. & Heisenberg, M. Localization of a short-term memory in Drosophila.
Science 288, 672±675 (2000).
13. Poodry, C. A., Hall, L. & Suzuki, D. T. Developmental properties of Shibire: a pleiotropic mutation
affecting larval and adult locomotion and development. Dev. Biol. 32, 373±386 (1973).
14. Poodry, C. A. & Edgar, L. Reversible alteration in the neuromuscular junctions of Drosophila
melanogaster bearing a temperature-sensitive mutation, shibire. J. Cell Biol. 81, 520±527 (1979).
15. Ramaswami, M., Rao, S., van der Bliek, A., Kelly, R. B. & Krishnan, K. S. Genetic studies on dynamin
function in Drosophila. J. Neurogenet. 9, 73±87 (1993).
16. Koenig, J. H. & Ikeda, K. Disappearance and reformation of synaptic vesicle membrane upon
transmitter release observed under reversible blockage of membrane retrieval. J. Neurosci. 9, 3844±
3860 (1989).
17. Chen, M. S. et al. Multiple forms of dynamin are encoded by shibire,aDrosophila gene involved in
endocytosis. Nature 351, 583±586 (1991).
18. van der Bliek, A. M. & Meyerowitz, E. M. Dynamin-like protein encoded by the Drosophila shibire gene
associated with vesicular traf®c. Nature 351, 411±414 (1991).
19. Damke, H., Baba, T., Warnock, D. E. & Schmid, S. L. Induction of mutant dynamin speci®cally blocks
endocytic coated vesicle formation. J. Cell Biol. 127, 915±934 (1994).
20. Grant, D., Unadkat, S., Katzen, A., Krishnan, K. S. & Ramaswami, M. Probable mechanisms
underlying interallelic complementation and temperature-sensitivity of mutations at the shibire locus
of Drosophila melanogaster. Genetics 149, 1019±1030 (1998).
21. Brand, A. H. & Perrimon, N. Targeted gene expression as a means of altering cell fates and generating
dominant phenotypes. Development 118, 401±415 (1993).
22. Yang, M. Y., Armstrong, J. D., Vilinsky, I., Strausfeld, N. J. & Kaiser, K. Subdivision of the Drosophila
mushroom bodies by enhancer-trap expression patterns. Neuron 15, 45±54 (1995).
23. Joiner, M. A. & Grif®th, L. C. Mapping of the anatomical circuit of CaM kinase-dependent courtship
conditioning in Drosophila. Learn. Mem. 6, 177±192 (1999).
24. Benzer, S. Behavioral mutants of Drosophila isolated by countercurrent distribution. Proc. Natl Acad.
Sci. USA 58, 1112±1119 (1967).
25. Tully, T., Preat, T., Boynton, S. C. & Del Vecchio, M. Genetic dissection of consolidated memory in
Drosophila. Cell 79, 35±47 (1994).
26. Wang, Y. T. & Linden, D. J. Expression of cerebellar long-term depression requires postsynaptic
clathrin-mediated endocytosis. Neuron 25, 635±647 (2000).
27. Carroll, R. C. et al. Dynamin-dependent endocytosis of ionotropic glutamate receptors. Proc. Natl
Acad. Sci. USA 96, 14112±14117 (1999).
28. Dudai, Y., Corfas, G. & Hazvi, S. What is the possible contribution of Ca
2+
-stimulated adenylate
cyclase to acquisition, consolidation and retention of an associative olfactory memory in Drosophila?
J. Comp. Physiol. A 162, 101±109 (1988).
29. Waddell, S., Armstrong, J. D., Kitamoto, T., Kaiser, K. & Quinn, W. G. The amnesiac gene product is
expressed in two neurons in the Drosophila brain that are critical for memory. Cell 103, 805±813 (2000).
30. Dubnau, J. & Tully, T. Functional anatomy: from molecule to memory. Curr. Biol.11, R240±R243 (2001).
Supplementary information is available on Nature's World-Wide Web site
(http://www.nature.com) or as paper copy from the London editorial of®ce of Nature.
Acknowledgements
We thank H. Cline, E. Drier, J. Huang, M. Regulski and K. Svoboda for comments on the
manuscript. This work was supported by the NIH (J.T.D. and T.T.) and the NSF (T.K.).
Correspondence and requests for materials should be addressed to J.D.
(e-mail: dubnau@cshl.org).
.................................................................
Leptin activates anorexigenic POMC
neurons through a neural network
in the arcuate nucleus
Michael A. Cowley*
²
, James L. Smart*
²
, Marcelo Rubinstein
³
,
Marcelo G. Cerda
Â
n
³
, Sabrina Diano§, Tamas L. Horvath§k, Roger D. Cone*
& Malcolm J. Low*
* The Vollum Institute, Oregon Health Sciences University, Portland,
Oregon 97201-3098, USA
³
Instituto de Investigaciones en Ingenierõ
Â
a Gene
Â
tica y Biologõ
Â
a Molecular,
CONICET and Department of Biology, School of Sciences,
University of Buenos Aires 1428, Argentina
§ Reproductive Neurosciences Unit, Department of Obstetrics and Gynecology and
k Department of Neurology, Yale Medical School, New Haven, Connecticut 06520,
USA
²
These authors contributed equally to this work
..............................................................................................................................................
The administration of leptin
1
to leptin-de®cient humans, and the
analogous Lep
ob
/Lep
ob
mice, effectively reduces hyperphagia and
obesity
2,3
. But common obesity is associated with elevated leptin,
which suggests that obese humans are resistant to this adipocyte
hormone. In addition to regulating long-term energy balance,
leptin also rapidly affects neuronal activity
4±6
. Proopiomelano-
cortin (POMC) and neuropeptide-Y types of neurons in the
arcuate nucleus of the hypothalamus
7
are both principal sites of
leptin receptor expression and the source of potent neuropeptide
modulators, melanocortins and neuropeptide Y, which exert
opposing effects on feeding and metabolism
8,9
. These neurons
are therefore ideal for characterizing leptin action and the
mechanism of leptin resistance; however, their diffuse distribu-
tion makes them dif®cult to study. Here we report electrophysio-
logical recordings on POMC neurons, which we identi®ed by
targeted expression of green ¯uorescent protein in transgenic
mice. Leptin increases the frequency of action potentials in the
anorexigenic POMC neurons by two mechanisms: depolarization
through a nonspeci®c cation channel; and reduced inhibition by
local orexigenic neuropeptide-Y/GABA (g-aminobutyric acid)
neurons. Furthermore, we show that melanocortin peptides
have an autoinhibitory effect on this circuit. On the basis of our
results, we propose an integrated model of leptin action and
neuronal architecture in the arcuate nucleus of the hypothalamus.
Previous studies suggest that leptin does not have equal effects on
all neuronal subtypes. Acute leptin treatment presumably activates
POMC, but not neuropeptide Y (NPY) neurons in the arcuate
nucleus of the hypothalamus (ARC), because c-Fos protein is
increased only in the former population and Socs3 messenger
RNA is increased in both
10
. Furthermore, a population of ARC
neurons seems to be inhibited directly by leptin, but the peptide
phenotype of these neurons has not been directly established
11±14
.In
addition to leptin receptors, both POMC and NPY neurons express
a receptor, MC3-R, for POMC-derived melanocortin peptides
15
.
The physiological role of this receptor is not well understood,
although MC3-R null mice have increased adiposity compared
with wild-type mice
16
.
To test the hypothesis that leptin selectively activates POMC
neurons, we ®rst generated a strain of transgenic mice expressing
green ¯uorescent protein (EGFP; Clontech) under the transcrip-
tional control of mouse Pomc genomic sequences, including a
region located between -13 kilobases (kb) and -2 kb that is required
for accurate neuronal expression
17
(Fig. 1a). Bright green ¯uores-
cence (509 nm) was seen in the two central nervous system regions
where POMC is produced: the ARC and the nucleus of the solitary
tract (data not shown).
© 2001 Macmillan Magazines Ltd

letters to nature
NATURE
|
VOL 411
|
24 MAY 2001
|
www.nature.com 481
Under ultraviolet (450±480 nm) excitation, POMC neurons were
clearly distinguished from adjacent, non-¯uorescent neurons (Fig.
1b) visualized under infrared optics. Double immuno¯uorescence
showed that there was more than 99% cellular colocalization of
EGFP and POMC peptides in the ARC (Fig. 1c). There was close
apposition of both tyrosine-hydroxylase- and NPY-stained termi-
nals on EGFP-expressing POMC neurons, but no evidence of
colocalization of the tyrosine hydroxylase or NPY immunoreactiv-
ity with EGFP (data not shown). Total ¯uorescent cell counts
of coronal hypothalamic sections determined that there were
3,148 6 62 (mean 6 s.e.m.; n = 3) POMC-EGFP neurons distrib-
uted through the whole ARC
18
(Fig. 1d). POMC neurons are located
both medially and ventrally in mouse ARC, in contrast to a
predominantly lateral position in rat ARC.
POMC-EGFP neurons in hypothalamic slices had a resting
membrane potential of -40 to -45 mV and exhibited frequent
spontaneous action potentials. The non-selective opioid agonist
met-enkephalin (Met-Enk, 30 mM; Sigma) caused a rapid (35±
40 s), reversible hyperpolarization (10±20 mV) of the membrane
potential of POMC cells (n = 10) and prevented spontaneous action
potential generation (Fig. 2a). In normal (2.5 mM K
+
) Krebs
solution, the reversal potential of the inwardly rectifying opioid
current was about -90 mV, whereas in 6.5 mM K
+
Krebs solution the
reversal potential shifted to about -60 mV (n = 3; Fig. 2b). The m-
opioid receptor (MOP-R) antagonist CTAP (1 mM; Phoenix Phar-
maceuticals) completely prevented the current induced by Met-Enk
in POMC cells (n = 3; Fig. 2c). These characteristics indicate that the
opioid current was caused by both activation of MOP-R and
increased ion conductance through G-protein-coupled, inwardly
rectifying potassium channels
19
. The similarity of opioid responses
in EGFP-labelled POMC neurons to those of guinea-pig
19
or
I
M
–13 –2 0 9
EcoRI EcoRINotl
Length (kb)
transcriptional start
translational stop
EGFP
123
Bregma –1.06 mm –1.22 mm
–1.58 mm
–1.34 mm
–1.70 mm
–1.46 mm
–1.82 mm –1.94 mm
Bregma –2.06 mm –2.18 mm –2.46 mm
–2.54 mm –2.80 mm–2.70 mm –2.92 mm
IR-DIC
Ultraviolet
Colocalization
β-endorphin
EGFP
d
b
a
c
2468
f
SOX
PaPo
PaPo
Pe
Pe
Pe
Pe
Pe
PeP
PeP
PeP
Pe
Arc
Arc
Arc
Arc
Arc
Arc
Arc
Arc
Arc
Arc
Arc
mt
mt
mt
mt
Gen
Gen
A12
Arc
Arc
Arc
VMH
VMH
SubI
VMH
VMH
PMD
PMV
VMH
AHP
AHP
AHP
ME
ME
ME
ME
ME
ME
TC
TC
VMH
VMH
VMH
VMH
DM
DM
DM
DM
PH
PH
PMV
PH
PH
PH
DM
DM
DM
DTM
DTM
DTM
DTM
SMT
SMT
SMT
SuMM
PMD
PMD
PMV
SuM
SuM
SuM
SuMM
sumx
sumx
sumx
pm
pm
mp
pm
MM
MM
MM
LH
LM
VTM
VTM
MRe
LH
L
SMT
DM
3V
3V
3V
3V
3V
3V
3V
3V
3V
3V
MRe
3V
3V
3V
f
f
f
r
f
f
f
f
f
f
f
f
f
f
f
f
–2.30 mm
ME
Figure 1 Generation of transgenic mice expressing EGFP in ARC POMC neurons.
a, Structure of the POMC-EGFP transgene. b, Identi®cation of a single POMC neuron (red
arrowhead on recording electrode tip) by infrared-differential interference contrast
(IR-DIC) microscopy (upper) and EGFP ¯uorescence (lower) in a living ARC slice before
electrophysiological recordings. c, Colocalization (yellow) of EGFP (green) and b-
endorphin immunoreactivity (red) in ARC POMC neurons. Scale bars, 50 mm(b, c).
d, Distribution of EGFP-positive neuronal soma throughout the ARC nucleus. Green open
circles = 5 cells, green ®lled circles = 10 cells.
2,450
–10
–20
–30
–40
–50
–60
–70
–80
2,200 2,250 2,300 2,350 2,400
Time (s)
Potential (mV)
Met-Enk (30 µM)
–30
20
–120 –110 –100 –90
–80 –70 –60 –50 –40
–30
–20
–10
10
Met-Enk
Met-Enk CTAP
Voltage (mV)
–130
–130 –110
–90
–70
–50
–75
–50
–25
25
2.5 mM [K
+
]
6.5 mM [K
+
]
Voltage (mV)
Net current (pA)
–150
Net current (pA)
a
b
c
Figure 2 Activation of MOP-Rs hyperpolarizes the EGFP-labelled POMC neurons by
opening G-protein-coupled inwardly rectifying potassium channels. a, Met-enkephalin
(Met-Enk) hyperpolarizes POMC neurons and inhibits all action potentials. Horizontal bar
indicates the time when 30 mM Met-Enk was bath applied to the slice. b, Met-Enk current
and reversal potential is shifted by extracellular K
+
concentration. c, Met-Enk activates
MOP-Rs on POMC neurons. A Met-Enk (30 mM) current was observed, and the MOP-R-
speci®c antagonist CTAP (1 mM) was applied for 1 min. After CTAP, Met-Enk elicited no
current. Figure is representative of three experiments.
© 2001 Macmillan Magazines Ltd

letters to nature
482 NATURE
|
VOL 411
|
24 MAY 2001
|
www.nature.com
mouse
20
POMC cells, identi®ed by post-recording immunohisto-
chemistry, suggests that expression of the EGFP transgene does not
compromise expression of receptors nor their coupling to second
messenger systems in POMC neurons.
We next determined the direct effects of leptin on identi®ed
POMC cells in slice preparations. Leptin (0.1±100 nM) depolarized
72 out of 77 POMC cells by 3±30 mV (Fig. 3a; depolarization at
100 nM leptin, 9.7 6 1.2 mV (mean 6 s.e.m.; n = 45)) in 2±10 min,
in a concentration-responsive manner (Fig. 3b). There were two
components to the depolarization and neither were fully reversible
within 40 min. First, the depolarization was due to a small inward
current that reversed at about -20 mV (Fig. 3c), suggesting the
involvement of a nonspeci®c cation channel
21
. Second, leptin
treatment decreased the GABA-mediated tone onto POMC cells.
GABA-mediated inhibitory postsynaptic currents (IPSCs) were
observed in POMC cells, and leptin (100 nM) decreased their
frequency by 25% (Fig. 3d) in 5 out of 15 cells, indicating that it
may act presynaptically to reduce GABA release (leptin had no effect
on IPSCs in 10 out of 15 POMC neurons).
The effect on IPSC frequency occurred with a similar lag to the
effect on membrane potential. Thus, leptin not only directly
depolarizes POMC neurons but also acts at GABA-secreting nerve
terminals to reduce the release of GABA onto POMC neurons,
allowing them to adopt a more depolarized resting potential. The
consistent depolarization of POMC cells by leptin was speci®c
because leptin had no effect on 5 out of 13 adjacent non-¯uorescent
cells tested (Fig. 3e), whereas it hyperpolarized ®ve (Fig. 3f) and
depolarized three other non-POMC neurons in the ARC (data
not shown). The electrophysiological effects of leptin reported
here are consistent with leptin's biological actionsÐleptin rapidly
causes release of a-melanocyte-stimulating hormone (a-MSH)
from rat hypothalami
4
, presumably by activating POMC neurons.
Previous reports of neuronal hyperpolarization by leptin
11,12
and
the colocalization of GABA and NPY
22
in subpopulations of ARC
neurons led us to speculate that leptin hyperpolarizes NPY/GABA
cells that directly innervate POMC neurons, and thus reduces
GABA-mediated drive onto POMC cells. Both leptin and NPY Y2
receptors are expressed on NPY neurons in the ARC
7,23
. Further-
more, activation of Y2 receptors inhibits NPY release from NPY
neurons
24
, and presumably would also diminish GABA release from
NPY/GABA terminals. This provided us with an alternative phar-
macological approach, independent of leptin, to test the proposed
innervation of POMC neurons by GABA-secreting NPY neurons.
Indeed, NPY (100 nM; Bachem) decreased the frequency of
GABA-mediated IPSCs by 55% in 3 min in all 12 POMC cells
tested (Fig. 4a). Both NPY and leptin still inhibited IPSCs in the
presence of tetrodotoxin (TTX; 6/6 and 3/5 cells, respectively),
indicating that some of the inhibition of IPSCs was occurring
through direct effects at presynaptic nerve terminals. POMC neu-
rons express the NPY Y1 receptor
23
, and NPYalso hyperpolarized all
POMC neurons tested by an average of 9 6 6mV(n = 3; data not
shown).
To con®rm the origin of GABA-mediated innervation on POMC
neurons from NPY/GABA terminals, we also tested the effect of the
highly selective MC3-R agonist
D-Trp
8
-gMSH (ref. 25) on local
GABA release.
D-Trp
8
-gMSH (7 nM) increased the frequency of
GABA-mediated IPSCs (280 6 90%) recorded from three out of
four POMC neurons (Fig. 4b). It had no effect on one cell. The
positive effect of MC3-R activation, together with the negative
effects of NPY and leptin, shows the dynamic range of the NPY/
GABA synapse onto POMC neurons and points to the important
role of this synapse in modulating signal ¯ow in the ARC.
D-Trp
8
-
gMSH (7 nM) also hyperpolarized (-5.5 6 2.4 mV) 9 out of 15
POMC neurons tested, and decreased the frequency of action
potentials (Fig. 4c); the remaining cells showed no signi®cant
response to
D-Trp
8
-gMSH. We could not determine whether these
effects were entirely due to increased GABA release onto the POMC
cells, or whether there was an additional postsynaptic action of
D-
Trp
8
-gMSH on POMC neurons, roughly half of which also express
the MC3-R
15
. Thus, MC3-R acts in a similar autoreceptor manner to
MOP-Rs on POMC neurons, diminishing POMC neuronal activity
in response to elevated POMC peptides.
To determine that the IPSCs in POMC neurons were due to local
innervation by NPY/GABA cells, we carried out multi-label immuno-
25
0
–25
–50
–75
–100
500 550
Time (s)
Leptin (100 nM)
600
–100 –80 –60 –40 -20
0
–60
–50
–40
–30
–20
–10
10
Voltage (mV)
Net current (pA)
log [Leptin (M)]
–11
–10 –9 –8 –7
12.5
10.0
7.5
5.0
2.5
0
–2.5
EC
50
= 5.9 nM
Depolarization (mV)
Leptin (100 nM)
020
1.5
1.0
0.5
0
Time (min)
25
0
–25
–50
–75
–100
500 550
Time (s)
Potential (mV)
Potential (mV)
Leptin (100 nM)
600
IPSC frequency (Hz)
–40
–60
–80
–100
600 700
Time (s)
Potential (mV)
Leptin (100 nM)
800
20
0
–20
a
b
c
de
f
Figure 3 Leptin depolarizes POMC neurons through a nonspeci®c cation channel, and
decreases the GABA-mediated tone onto POMC cells. a, Leptin depolarizes POMC
neurons and increases the frequency of action potentials within 1±10 min of addition.
Figure is representative of recordings made from 77 POMC neurons. b, Leptin causes a
concentration-dependent depolarization of POMC cells. Depolarization by leptin was
determined at 0.1, 1, 10, 50 and 100 nM (effector concentration at half-maximum
response, EC
50
= 5.9 nM) in 8, 7, 9, 3 and 45 cells, respectively. c, Leptin depolarizes
POMC cells by activating a nonspeci®c cation current. Figure is representative of the
response in ten cells. d, Leptin decreases the frequency of IPSCs in POMC cells. Figure is
an example of ®ve cells in which leptin (100 nM) decreased the frequency of IPSCs.
e, Leptin had no effect on ®ve adjacent non-¯uorescent ARC neurons. f, Leptin
hyperpolarized ®ve non-¯uorescent ARC neurons.
© 2001 Macmillan Magazines Ltd

letters to nature
NATURE
|
VOL 411
|
24 MAY 2001
|
www.nature.com 483
histochemistry using light and electron microscopy. Although
independent NPY
27
and GABA
27
innervation of POMC cells has
been reported, colocalization of NPY and GABA in nerve terminals
forming synapses onto POMC cells has not been shown. Similar to
the rat
26
, a dense innervation of POMC cells by NPYaxon terminals
was detected in the mouse (Fig. 4d). Electron microscopy con®rmed
the co-expression of NPYand GABA in axon terminals, and showed
that these boutons established synapses on the perikarya of all 15
ARC POMC neurons analysed (Fig. 4e).
These observations allow us to propose a detailed model of
regulation of this circuit including: dual mechanisms of leptin
action in the ARC, interactions between NPY/GABA and POMC
neurons, and autoregulatory feedback from opioid and melano-
cortin peptides as well as NPY (Fig. 4f). In this model, leptin directly
depolarizes the POMC neurons while simultaneously hyper-
polarizing the somata of NPY/GABA neurons, and diminishes
release from NPY/GABA terminals. This diminished GABA release
disinhibits the POMC neurons. Together, the direct and indirect
effects of leptin result in an activation of POMC neurons and an
increased frequency of action potentials. In addition, both POMC
and NPY neurons express autoreceptors for some of their respective
neuropeptide products (b-endorphin or a-MSH, and NPY, respec-
tively) and activation of these autoreceptors may provide ultrashort
feedback loops that further modulate the effects of leptin on POMC
neurons. The effects of NPY and melanocortin agonists on IPSC
frequency are similar to the effect of NPY on IPSCs in the
paraventricular nucleus of the hypothalamus
28
.
Thus, there seem to be two classes of neurons accounting for
leptin sensitivity in the brain: those activated (depolarized) to
release anorexigenic peptides; and those inhibited (hyperpolarized)
with a consequent reduction in release of orexigenic peptides.
Identifying viable POMC neurons with EGFP provides an invalu-
able tool for determining complex actions on these important
neurons. Furthermore, an understanding of the acute effects of
leptin provides a model system for studying the mechanisms
underlying the development of leptin resistance
29
. M
Methods
Generation of POMC-EGFP mice
The EGFP cassette contains its own Kozak consensus translation initiation site, along with
SV40 polyadenylation signals downstream of the EGFP coding sequences, which directs
proper processing of the 39 end of the EGFP mRNA. We introduced the EGFP cassette by
standard techniques into the 59 untranslated region of exon 2 of a mouse Pomc genomic
clone containing 13 kb of 59 and 2 kb of 39 ¯anking sequences
17
. The transgene was
microinjected into pronuclei of one-cell-stage embryos of C57BL/6J mice (Jackson
Laboratories) as described
17
. One founder was generated and bred to wild-type C57BL/6J
to produce N
1
hemizygous mice. In addition, we also generated N
2
and subsequent
generations of mice homozygous for the transgene. The mice are fertile and have normal
growth and development.
Immuno¯uorescence and GFP colocalization
Anaesthetized mice were perfused transcardially with 4% paraformaldehyde, and free-
¯oating brain sections were prepared with a vibratome. We processed sections for
immuno¯uoresence and colocalization of GFP ¯uoresence using standard techniques.
Primary antisera and their ®nal dilutions were rabbit anti-b-endorphin, 1:2500 (v/v);
rabbit anti-NPY, 1:25,000 (v/v) (Alanex Corp.); rabbit anti-ACTH, 1:2000 (v/v); and
mouse anti-TH, 1:1000 (v/v) (Incstar). After rinsing, sections were incubated with
10 mg ml
-1
biotinylated horse anti-mouse/rabbit immunoglobulin-g (IgG) (Vector
Laboratories) followed by Cy-3 conjugated streptavidin, 1:500 (v/v) (Jackson Immuno-
research Laboratories). Photomicrographs were taken on an Axioscop (Zeiss) using
¯uoroscein isothiocyanate (FITC) and rhodamine isothiocyanate (RITC) ®lter sets
(Chroma Technology Corp.).
Electrophysiology
We cut 200-mm thick coronal slices from the ARC of 4-week-old male POMC-EGFP mice.
Slices were maintained at 35 8C in Krebs solution ((in mM): 126 NaCl, 2.5 KCl, 1.2 MgCl
2
,
2.4 CaCl
2
.2H
2
O, 1.2 NaH
2
PO
4
.H
2
O, 21.4 NaHCO
3
, 11.1 glucose), and were saturated with
95% O
2
5% CO
2
for 1 h before recordings. Recordings were made in Krebs solution at
35 8C. Slices were visualized on an Axioskop FS2 (Zeiss) through standard infrared optics
and using epi¯uoresence through a FITC ®lter set (see Fig. 1c). Whole-cell recordings were
made from ¯uorescent neurons using an Axopatch 1D ampli®er (Axon Instruments) and
Clampex 7 (Axon Instruments).
We determined resting membrane potentials using an event-detection protocol on a
PowerLab system (AD Instruments) to average expanded traces of the membrane
potential. Drugs were applied to the bath over the times indicated. The resting membrane
potential was stable for up to 1 h in cells treated with Krebs solution alone (data not
shown). I±V relationships for the Met-Enk currents were established using a step protocol
(-60 mV holding potential, sequentially pulsed (40 ms) from -120 to -50 mV; cells were
returned to -60 mV for 2 s between voltage steps). The protocol was repeated after
addition of Met-Enk. The net current was the difference between the two I±V relation-
51015
5
4
3
2
1
0
D-Tr p
8
-
γ
MSH (7 nM)
Time (min)
IPSC frequency (Hz)
Time (min)
NPY (100 nM)
0123456
1.5
1.0
0.5
0
IPSC frequency (Hz)
25
0
–25
–50
–75
–100
100
Time (s)
Potential (mV)
200 300 400
D-Tr p
8
-
γ
MSH (7 nM)
a
b
c
f
d
e
Target site
NPY/AGRP
POMC
GABA
POMC
POMC
NPY
NPY
Y2-R
Y2-R
Y1-R
MC3-R
GABA
MC3-R
MOP-R
MC3-R
Lep-R
Lep-R
Figure 4 The GABA-mediated inputs to POMC cells are from NPY neurons that co-express
GABA. a, NPY decreases the frequency of mini-IPSCs in POMC neurons. b,
D-Trp
8
gMSH
(7nM)Ða dose that selectively activates MC3-RÐincreases the frequency of GABA-
mediated IPSCs in POMC neurons. c,
D-Trp
8
-gMSH hyperpolarizes POMC neurons.
Panels a±c are representative results. d, Expression of NPY in nerve terminals adjacent to
POMC neurons in the ARC. NPY nerve terminals (black, green arrowheads); POMC
neuronal soma (brown). Scale bar, 10 mm. e, Expression of GABA and NPY in nerve
terminals synapsing onto POMC neurons in the ARC. GABA immunoreactivity (10-nm gold
particles, red arrowheads) and NPY immunoreactivity (25-nm gold particles, green
arrows) are in separate vesicle populations colocalized in synaptic boutons that make
direct contact with the soma of POMC neurons (DAB contrasted with uranyl acetate and
lead citrate, diffuse black in cytoplasm). Scale bar, 1 mm. f, Model of leptin regulation of
NPY/GABA and POMC neurons in the ARC (see text).
© 2001 Macmillan Magazines Ltd

letters to nature
484 NATURE
|
VOL 411
|
24 MAY 2001
|
www.nature.com
ships. This protocol was repeated in Krebs solution with 6.5 mM K
+
. To identify the
postsynaptic leptin current, I±V relationships were performed similarly with slow voltage
ramps (5 mV s
-1
from -100 to -20 mV) before and 10 min after adding leptin (100 nM).
GABA-mediated IPSCs were recorded using a CsCl internal electrode solution ((in
mM): 140 CsCl, 10 HEPES, 5 MgCl
2
, 1 BAPTA, 5 Mg-ATP, 0.3 Na-GTP). Both mini IPSCs
and large amplitude (presumably multisynaptic) IPSCs were observed in the untreated
slices. TTX (1 mM) abolished large IPSCs. We acquired data before and after drug addition
at a -50-mV holding potential in 2-s sweeps every 4 s for the times indicated in the ®gures.
Mini-postsynaptic currents were analysed using Axograph 4 (Axon Instruments). IPSCs
and excitatory postsynaptic currents (EPSCs) were distinguished on the basis of their
decay constants; in addition, picrotoxin (100 mM) blocked all IPSCs. POMC neurons
receive a low EPSC tone, and the frequency was not modulated by any of the treatments
described here.
Immunostaining for light and electron microscopy
We carried out double immunocytochemistry for NPY and POMC using different colour
diaminobenzidine (DAB) chromogens on ®xed mouse hypothalami according to
published protocols
27
. For electron microscopy, pre-embedding immunostaining for b-
endorphin was done with an ABC Elite kit (Vector Laboratories) and a DAB reaction,
followed by post-embedding labelling of GABA and NPY using rabbit anti-GABA, 1:1000
(v/v), and gold-conjugated (10-nm) goat anti-rabbit IgG or sheep anti-NPY and gold-
conjugated (25-nm) goat anti-sheep IgG. Sections were contrasted with saturated uranyl
acetate (10 min) and lead citrate (20±30 s), and examined using a Philips CM-10 electron
microscope.
Received 19 December 2000; accepted 12 March 2001.
1. Zhang, Y. et al. Positional cloning of the mouse obese gene and its human homologue. Nature 372,
425±432 (1994); erratum ibid. 374, 479 (1995).
2. Farooqi, I. S. et al. Effects of recombinant leptin therapy in a child with congenital leptin de®ciency.
N. Engl. J. Med. 341, 879±884 (1999).
3. Camp®eld, L. A., Smith, F. J., Guisez, Y., Devos, R. & Burn, P. Recombinant mouse OB protein:
evidence for a peripheral signal linking adiposity and central neural networks. Science 269, 546±549
(1995).
4. Kim, M. S. et al. The central melanocortin system affects the hypothalamo-pituitary thyroid axis and
may mediate the effect of leptin. J. Clin. Invest. 105, 1005±1011 (2000).
5. Haynes, W. G., Morgan, D. A., Walsh, S. A., Mark, A. L. & Sivitz, W. I. Receptor-mediated regional
sympathetic nerve activation by leptin. J. Clin. Invest. 100, 270±278 (1997).
6. Haynes, W. G., Morgan, D. A., Djalali, A., Sivitz, W. I. & Mark, A. L. Interactions between the
melanocortin system and leptin in control of sympathetic nerve traf®c. Hypertension 33, 542±547
(1999).
7. Hakansson, M. L., Brown, H., Ghilardi, N., Skoda, R. C. & Meister, B. Leptin receptor immunor-
eactivity in chemically de®ned target neurons of the hypothalamus. J. Neurosci. 18, 559±572 (1998).
8. Kalra, S. P. et al. Interacting appetite-regulating pathways in the hypothalamic regulation of body
weight. Endocr. Rev. 20, 68±100 (1999).
9. Cone, R. D. The central melanocortin system and energy homeostasis. Trends Endocrinol. Metab. 10,
211±216 (1999).
10. Elias, C. F. et al. Leptin differentially regulates NPY and POMC neurons projecting to the lateral
hypothalamic area. Neuron 23, 775±786 (1999).
11. Glaum, S. R. et al. Leptin, the obese gene product, rapidly modulates synaptic transmission in the
hypothalamus. Mol. Pharmacol. 50, 230±235 (1996).
12. Spanswick, D., Smith, M. A., Groppi, V. E., Logan, S. D. & Ashford, M. L. Leptin inhibits hypothalamic
neurons by activation of ATP-sensitive potassium channels. Nature 390, 521±525 (1997).
13. Lee, K., Dixon, A. K., Richardson, P. J. & Pinnock, R. D. Glucose-receptive neurones in the rat
ventromedial hypothalamus express KATP channels composed of Kir6.1 and SUR1 subunits.
J. Physiol. (Lond.) 515, 439±452 (1999).
14. Shiraishi, T., Sasaki, K., Niijima, A. & Oomura, Y. Leptin effects on feeding-related hypothalamic and
peripheral neuronal activities in normal and obese rats. Nutrition 15, 576±579 (1999).
15. Bagnol, D. et al. Anatomy of an endogenous antagonist: relationship between Agouti-related protein
and proopiomelanocortin in brain. J. Neurosci. [online] (cited 25 Aug. 99) ,http://www.jneurosci.
org/cgi/content/full/19/18/RC26. (1999).
16. Butler, A. A. et al. A unique metabolic syndrome causes obesity in the melanocortin-3 receptor-
de®cient mouse. Endocrinology 141, 3518±3521 (2000).
17. Young, J. I. et al. Authentic cell-speci®c and developmentally regulated expression of pro-opiome-
lanocortin genomic fragments in hypothalamic and hindbrain neurons of transgenic mice. J. Neurosci.
18, 6631±6640 (1998).
18. Franklin, K. B. J. & Paxinos, G. The Mouse Brain in Stereotaxic Coordinates (Academic, San Diego,
1997).
19. Kelly, M. J., Loose, M. D. & Ronnekleiv, O. K. Opioids hyperpolarize b-endorphin neurons via
m-receptor activation of a potassium conductance. Neuroendocrinology 52, 268±275 (1990).
20. Slugg, R. M., Hayward, M. D., Ronnekleiv, O. K., Low, M. J. & Kelly, M. J. Effect of the m-opioid
agonist DAMGO on medial basal hypothalamic neurons in b-endorphin knock-out mice.
Neuroendocrinology 72, 208±217 (2000).
21. Powis, J. E., Bains, J. S. & Ferguson, A. V. Leptin depolarizes rat hypothalamic paraventricular nucleus
neurons. Am. J. Physiol. 274, R1468±R1472 (1998).
22. Horvath, T. L., Bechmann, I., Naftolin, F., Kalra, S. P. & Leranth, C. Heterogeneity in the neuropeptide
Y-containing neurons of the rat arcuate nucleus: GABAergic and non-GABAergic subpopulations.
Brain Res. 756, 283±286 (1997).
23. Broberger, C., Landry, M., Wong, H., Walsh, J. N. & Hokfelt, T. Subtypes Y1 and Y2 of the
neuropeptide Y receptor are respectively expressed in pro-opiomelanocortin- and neuropeptide-Y-
containing neurons of the rat hypothalamic arcuate nucleus. Neuroendocrinology 66, 393±408 (1997).
24. King, P. J., Widdowson, P. S., Doods, H. N. & Williams, G. Regulation of neuropeptide Y release by
neuropeptide Y receptor ligands and calcium channel antagonists in hypothalamic slices.
J. Neurochem. 73, 641±646 (1999).
25. Grieco, P., Balse, P. M., Weinberg, D., MacNeil, T. & Hruby, V. J.
D-Amino acid scan of gamma-
melanocyte-stimulating hormone: importance of Trp(8) on human MC3 receptor selectivity. J. Med.
Chem. 43, 4998±5002 (2000).
26. Csiffary, A., Gorcs, T. J. & Palkovits, M. Neuropeptide Y innervation of ACTH-immunoreactive
neurons in the arcuate nucleus of rats: a correlated light and electron microscopic double
immunolabeling study. Brain Res. 506, 215±222 (1990).
27. Horvath, T. L., Naftolin, F. & Leranth, C. GABAergic and catecholaminergic innervation of
mediobasal hypothalamic beta-endorphin cells projecting to the medial preoptic area. Neuroscience
51, 391±399 (1992).
28. Cowley, M. A. et al. Integration of NPY, AGRP, and melanocortin signals in the hypothalamic
paraventricular nucleus: evidence of a cellular basis for the adipostat. Neuron 24, 155±163 (1999).
29. Halaas, J. L. et al. Physiological response to long-term peripheral and central leptin infusion in lean
and obese mice. Proc. Natl Acad. Sci. USA 94, 8878±8883 (1997).
Acknowledgements
We wish to thank V. J. Hruby for the D-Trp
8
-gMSH, O. K. Ronnekliev, R. G. Allen and
M. R. Brown for antisera and J. T. Williams and J. M. Brundege for advice. This work was
supported by the NIH, a Fogarty International Research Collaborative Award, the
International Scholar Program of the Howard Hughes Medical Institute, and Agencia
Nacional de Promocio
Â
n Cienti®ca y Technolo
Â
gica.
Correspondence and requests for materials should be addressed to R.D.C.
(e-mail: cone@ohsu.edu) or M.J.L. (e-mail: low@ohsu.edu).
.................................................................
Calmodulin bifurcates the local
Ca
2+
signal that modulates
P/Q-type Ca
2+
channels
Carla D. DeMaria*, Tuck Wah Soong
²
, Badr A. Alseikhan*,
Rebecca S. Alvania* & David T. Yue*
* The Johns Hopkins University School of Medicine, Departments of Biomedical
Engineering and Neuroscience, Program in Molecular and Cellular Systems
Physiology, 720 Rutland Avenue, Baltimore, Maryland 21205, USA
²
National Neuroscience Institute, 11 Jalan Tan Tuck Seng, Singapore 308433, and
Department of Physiology, National University of Singapore
..............................................................................................................................................
Acute modulation of P/Q-type (a
1A
) calcium channels by neuro-
nal activity-dependent changes in intracellular Ca
2+
concentra-
tion may contribute to short-term synaptic plasticity
1±3
,
potentially enriching the neurocomputational capabilities of
the brain
4,5
. An unconventional mechanism for such channel
modulation has been proposed
6,7
in which calmodulin (CaM)
may exert two opposing effects on individual channels, initially
promoting (`facilitation') and then inhibiting (`inactivation')
channel opening. Here we report that such dual regulation
arises from surprising Ca
2+
-transduction capabilities of CaM.
First, although facilitation and inactivation are two competing
processes, both require Ca
2+
-CaM binding to a single `IQ-like'
domain on the carboxy tail of a
1A
8
; a previously identi®ed `CBD'
CaM-binding site
6,7
has no detectable role. Second, expression of a
CaM mutant with impairment of all four of its Ca
2+
-binding sites
(CaM
1234
) eliminates both forms of modulation. This result
con®rms that CaM is the Ca
2+
sensor for channel regulation,
and indicates that CaM may associate with the channel even
before local Ca
2+
concentration rises. Finally, the bifunctional
capability of CaM arises from bifurcation of Ca
2+
signalling by the
lobes of CaM: Ca
2+
binding to the amino-terminal lobe selectively
initiates channel inactivation, whereas Ca
2+
sensing by the carboxy-
terminal lobe induces facilitation. Such lobe-speci®c detection
provides a compact means to decode local Ca
2+
signals in two
ways, and to separately initiate distinct actions on a single
molecular complex.
To simplify the dissection of the molecular mechanisms, we
studied recombinant P/Q-type (a
1A
/b
2a
/a
2
d) channels expressed
in mammalian HEK293 cells. Figure 1a shows that Ca
2+
-dependent
© 2001 Macmillan Magazines Ltd
Citations
More filters
Journal ArticleDOI

Central nervous system control of food intake and body weight

TL;DR: This new information provides a biological context within which to consider the global obesity epidemic and identifies numerous potential avenues for therapeutic intervention and future research.
Journal ArticleDOI

Gut hormone PYY 3-36 physiologically inhibits food intake

TL;DR: In this paper, the authors showed that post-prandial elevation of PYY3-36 may act through the arcuate nucleus Y2R to inhibit feeding in a gut-hypothalamic pathway.
Journal Article

BRIEF COMMUNICATION ARISING: Gut hormone PYY3-36 physiologically inhibits food intake

TL;DR: The authors showed that post-prandial elevation of PYY3-36 may act through the arcuate nucleus Y2R to inhibit feeding in a gut-hypothalamic pathway.
Journal ArticleDOI

Inhibition of food intake in obese subjects by peptide YY3-36.

TL;DR: It is found that obese subjects were not resistant to the anorectic effects of PYY, and endogenous PYY levels were low in obese subjects, suggesting that PYY deficiency may contribute to the pathogenesis of obesity.
References
More filters
Book

The Mouse Brain in Stereotaxic Coordinates

TL;DR: The 3rd edition of this atlas is now in more practical 14"x11" format for convenient lab use and includes a CD of all plates and diagrams, as well as Adobe Illustrator files of the diagrams, and a variety of additional useful material.
Journal ArticleDOI

Positional cloning of the mouse obese gene and its human homologue

TL;DR: The ob gene product may function as part of a signalling pathway from adipose tissue that acts to regulate the size of the body fat depot.
Journal ArticleDOI

Targeted gene expression as a means of altering cell fates and generating dominant phenotypes.

TL;DR: The GAL4 system, a system for targeted gene expression that allows the selective activation of any cloned gene in a wide variety of tissue- and cell-specific patterns, has been designed and used to expand the domain of embryonic expression of the homeobox protein even-skipped.
Journal ArticleDOI

Recombinant mouse OB protein: evidence for a peripheral signal linking adiposity and central neural networks.

TL;DR: The behavioral effects after brain administration suggest that OB protein can act directly on neuronal networks that control feeding and energy balance in ob/ob and diet-induced obese mice.
Journal ArticleDOI

Effects of recombinant leptin therapy in a child with congenital leptin deficiency.

TL;DR: The administration of leptin corrects their obesity by reducing their food intake and increasing their energy expenditure and these mice also have hyperinsulinemia, corticosterone excess, and infertility, which also are reversed by treatment with leptin.
Related Papers (5)
Frequently Asked Questions (1)
Q1. What are the contributions in this paper?

In this paper, the adjusted error rates ( a9 ) were P = 0.004 and 0.01, respectively, for the 13 subsequent planned pairwise comparisons in Fig. 4a and for the 5 subsequent planned comparisons inFig. 4b.