Download Local network regulation of orexin neurons in the lateral hypothalamus

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Haemodynamic response wikipedia , lookup

Bird vocalization wikipedia , lookup

Neuroeconomics wikipedia , lookup

Neuroplasticity wikipedia , lookup

Types of artificial neural networks wikipedia , lookup

Environmental enrichment wikipedia , lookup

Long-term depression wikipedia , lookup

Adult neurogenesis wikipedia , lookup

Connectome wikipedia , lookup

Convolutional neural network wikipedia , lookup

Brain wikipedia , lookup

Biochemistry of Alzheimer's disease wikipedia , lookup

Apical dendrite wikipedia , lookup

Electrophysiology wikipedia , lookup

Neuromuscular junction wikipedia , lookup

Biological neuron model wikipedia , lookup

Artificial general intelligence wikipedia , lookup

Single-unit recording wikipedia , lookup

Axon wikipedia , lookup

Multielectrode array wikipedia , lookup

Axon guidance wikipedia , lookup

Synaptogenesis wikipedia , lookup

Nonsynaptic plasticity wikipedia , lookup

Activity-dependent plasticity wikipedia , lookup

Metastability in the brain wikipedia , lookup

Caridoid escape reaction wikipedia , lookup

Neural oscillation wikipedia , lookup

Neural coding wikipedia , lookup

Mirror neuron wikipedia , lookup

Development of the nervous system wikipedia , lookup

Stimulus (physiology) wikipedia , lookup

Endocannabinoid system wikipedia , lookup

Neurotransmitter wikipedia , lookup

Central pattern generator wikipedia , lookup

Hypothalamus wikipedia , lookup

Nervous system network models wikipedia , lookup

Neuroanatomy wikipedia , lookup

Chemical synapse wikipedia , lookup

Molecular neuroscience wikipedia , lookup

Premovement neuronal activity wikipedia , lookup

Feature detection (nervous system) wikipedia , lookup

Optogenetics wikipedia , lookup

Circumventricular organs wikipedia , lookup

Neurotoxin wikipedia , lookup

Synaptic gating wikipedia , lookup

Pre-Bötzinger complex wikipedia , lookup

Clinical neurochemistry wikipedia , lookup

Channelrhodopsin wikipedia , lookup

Neuropsychopharmacology wikipedia , lookup

Transcript
Am J Physiol Regul Integr Comp Physiol 301: R572–R580, 2011.
First published June 22, 2011; doi:10.1152/ajpregu.00674.2010.
Review
Local network regulation of orexin neurons in the lateral hypothalamus
Julia Burt, Christian O. Alberto, Matthew P. Parsons, and Michiru Hirasawa
Division of Biomedical Sciences, Faculty of Medicine, Memorial University, St. John’s, Newfoundland, Canada
Submitted 12 October 2010; accepted in final form 16 June 2011
food intake; MCH neuron; sleep
THE LATERAL HYPOTHALAMUS (LH) is the most extensively interconnected area of the hypothalamus, allowing it to control and
convey a variety of essential autonomic and somatomotor
functions. Neuroanatomical studies have demonstrated direct
projections from the LH to other hypothalamic areas, cortical/
limbic areas, and the autonomic and motor system of the
brainstem (101, 102, 104). Such extensive connectivity is
thought to represent the anatomical underpinning that supports
sleep-wake regulation (10, 90, 106), energy homeostasis, as
well as cognitive, reward-related, and emotion-related functions (8, 24, 103).
There are a number of neuronal populations that have been
identified within this hypothalamic region. To a significant
degree, the function of the LH can be attributed to orexin
neurons that synthesize orexin A and B (also called hypocretin-1 and -2), 33 and 28 amino acid peptides, respectively,
cleaved from the precursor protein prepro-orexin (22, 96).
Orexin effects are mediated by two subtypes of orexin receptors, OX1R and OX2R, which have extensive, yet distinct,
expression patterns in the brain (118). Orexin neurons are
almost exclusively localized in the LH and adjacent periforni-
Address for reprint requests and other correspondence: M. Hirasawa, Division
of BioMedical Sciences, Faculty of Medicine, Memorial Univ., 300 Prince Philip
Dr., St. John’s, NL, A1B 3V6, Canada (e-mail: [email protected]).
R572
cal area (PFA) in animal and human brains and have wideranging projections (14, 30, 90, 121), including the LH/PFA
itself, where these neurons make synaptic contacts onto one
another (41). Physiological functions of the orexin peptides
include stabilization of wakefulness (97, 105), energy homeostasis (83, 96, 112), behavioral responses to food reward and
addictive drugs (7, 38), neuroendocrine and autonomic outflow
(99), and analgesia (17).
Many of orexin’s known behavioral effects, such as stimulation of food intake, wheel running, and spontaneous physical
activity can be induced by local injection of orexins into the
LH/PFA (27, 60, 61, 79, 111, 113, 116, 125). An orexin
injection into the LH/PFA also results in Fos expression, a
marker for neuronal activation (79), suggesting that endogenous orexin release within this area has an excitatory effect,
which may lead to an amplification of excitation by further
activating other orexin neurons. The idea of orexin neurons
working in concert explains how a relatively small population
of neurons (90) can coordinate diverse physiological functions.
Furthermore, many previous studies, mainly using in vitro
electrophysiological and histological approaches, collectively
suggest complex interactions that take place among orexin
neurons and other cells types within the LH/PFA. These
interactions can ultimately determine the sensitivity to and
integration of incoming signals and the levels of outgoing
signals.
0363-6119/11 Copyright © 2011 the American Physiological Society
http://www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.32.246 on June 17, 2017
Burt J, Alberto CO, Parsons MP, Hirasawa M. Local network regulation
of orexin neurons in the lateral hypothalamus. Am J Physiol Regul Integr Comp
Physiol 301: R572–R580, 2011. First published June 22, 2011;
doi:10.1152/ajpregu.00674.2010.—Obesity and inadequate sleep are among the
most common causes of health problems in modern society. Thus, the discovery
that orexin (hypocretin) neurons play a pivotal role in sleep/wake regulation, energy
balance, and consummatory behaviors has sparked immense interest in understanding the regulatory mechanisms of these neurons. The local network consisting of
neurons and astrocytes within the lateral hypothalamus and perifornical area
(LH/PFA), where orexin neurons reside, shapes the output of orexin neurons and
the LH/PFA. Orexin neurons not only send projections to remote brain areas but
also contribute to the local network where they release multiple neurotransmitters
to modulate its activity. These neurotransmitters have opposing actions, whose
balance is determined by the amount released and postsynaptic receptor desensitization. Modulation and negative feedback regulation of excitatory glutamatergic
inputs as well as release of astrocyte-derived factors, such as lactate and ATP, can
also affect the excitability of orexin neurons. Furthermore, distinct populations of
LH/PFA neurons express neurotransmitters with known electrophysiological actions on orexin neurons, such as melanin-concentrating hormone, corticotropinreleasing factor, thyrotropin-releasing hormone, neurotensin, and GABA. These
LH/PFA-specific mechanisms may be important for fine tuning the firing activity of
orexin neurons to maintain optimal levels of prolonged output to sustain wakefulness and stimulate consummatory behaviors. Building on these exciting findings
should shed further light onto the cellular mechanisms of energy balance and
sleep-wake regulation.
Review
NEUROTRANSMITTER INTERACTIONS IN THE LATERAL HYPOTHALAMUS
R573
Fig. 1. Regulation of orexin neurons by local neurotransmitters. The activity of
orexin neurons is controlled by positive and negative feedback mechanisms
mediated by neurotransmitters released by lateral hypothalamus/perifornical
area (LH/PFA) neurons. Orexin neurons corelease excitatory neurotransmitters
orexin (Orx) and glutamate (Glut), and inhibitory transmitters dynorphin (Dyn)
and nociceptin/orphanin FQ (N/OFQ), all of which can directly affect postsynaptic orexin neurons (A) or indirectly by modulating glutamate release at
excitatory synapses (B). The balance between the excitatory and inhibitory
effects determines the activity levels of the postsynaptic cell. C: at excitatory
synapses, glutamate acts on presynaptic autoreceptors to inhibit glutamate
release, completing a short, negative feedback loop. In addition, depolarization
of orexin neurons induces release of endocannabinoids (eCB), which in turn
act as retrograde messengers to inhibit excitatory inputs (D). Reciprocal
communications between orexin neurons and MCH/GABA neurons (E) as well
as GABAergic interneurons (F) also represent local feedback mechanisms.
Excitation of Orexin Neurons
Orexin neurons have intrinsic features and the local environment that can promote a long-lasting firing activity. This
may be because neuropeptide release from dense core vesicles
needs prolonged depolarization (21) and/or because of its
primary role in the maintenance of wakefulness and homeostasis etc., where sustained output may be more relevant than
precise timing.
Orexin neurons are intrinsically in a depolarized state (28),
largely due to a constitutively active nonselective cation current mediated by transcient receptor potential C channels (20).
Furthermore, the local network of orexin neurons allows them
to sustain an active state and/or to recruit a larger number of
orexin neurons, utilizing excitatory neurotransmitters, such as
orexins and glutamate (1, 94, 117). Orexin A or B activate
OX2Rs, which in turn open nonselective cation channels to
depolarize orexin neurons (137) (Fig. 1A) and increase presynaptic glutamate release without affecting GABA transmission
(66, 137) (Fig. 1B). Notably, orexin perikarya receive numerous excitatory inputs that greatly outnumber inhibitory synAJP-Regul Integr Comp Physiol • VOL
Fig. 2. Regulation of orexin (Orx) neurons by astrocytes. Activated astrocytes
release lactate (Lac) and protons (H⫹) through monocarboxylate transporters
(MCT). Orexin neurons metabolize astrocyte-derived lactate, not glucose
itself, as an energy substrate to sustain firing activity. In addition, a pH drop,
resulting from the MCT activity, stimulates these neurons. ATP can also be
released into the extracellular space by astrocytes or neurons, which has a
direct excitatory effect on orexin neurons. Adenosine can be hydrolyzed from
ATP or released from neurons, which has a direct and indirect (synaptic) effect
leading to inhibition of orexin neurons. Glut, glutamate.
301 • SEPTEMBER 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.32.246 on June 17, 2017
apses as shown by ultrastructural and electrophysiological
studies (56), which provide the structural basis for fast glutamatergic transmission that supplies a major excitatory drive
mediated by ␣-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) and N-methyl-D-aspartate (NMDA), but not
kainate receptors (4, 66, 92).
Glutamatergic transmission may also stimulate orexin neurons indirectly in an unconventional manner (Fig. 2). Glutamate is known to stimulate lactate production in astrocytes,
which in turn release lactate into the extracellular space
through monocarboxylate transporters (MCTs) (45, 89). Since
MCTs cotransport protons along with lactate, lactate release
accompanies a local decline in extracellular pH (45). A decrease in pH would result in depolarization of orexin neurons
(129). Moreover, orexin neurons utilize astrocyte-derived lactate, but not glucose, as an energy substrate to maintain
spontaneous firing activity (86). In conditions where lactate
supply is insufficient or ATP production is impaired, ATPsensitive K⫹ channels become active and hyperpolarize these
neurons (86). Therefore, the MCT-mediated astrocyte-neuronlactate shuttle can mediate the excitatory action of glutamatergic transmission.
Activation of astrocytes or neurons can also induce release
of ATP, which can, in turn, act as a transmitter molecule (26,
42) (Fig. 2). Extracellular ATP directly depolarizes orexin
neurons via ionotropic P2X receptors expressed on the plasma
Review
R574
NEUROTRANSMITTER INTERACTIONS IN THE LATERAL HYPOTHALAMUS
Self-Regulatory Mechanisms
Excitatory positive-feedback mechanisms can conceivably
continue until exhaustion without regulatory mechanisms that
keep them in check. Indeed, many negative feedback pathways
exist that regulate orexin neurons upon their excitation, including inhibitory neurotransmitters released by the same neurons.
Dynorphin, a neuropeptide coexpressed by orexin neurons
(18), directly hyperpolarizes these neurons by a ␬-opioid receptor-mediated activation of G protein-dependent inwardly
rectifying potassium channels and suppression of Ca2⫹ currents (67) (Fig. 1A). Interestingly, the dynorphin effect desensitizes faster than that of orexins, which results in a timedependent shift in the balance between inhibitory (dynorphin)
and excitatory (orexin) influence over the neurons (67). Another neuropeptide recently identified to be coexpressed by
orexin neurons is nociceptin/orphanin FQ (N/OFQ) (73).
N/OFQ inhibits orexin neurons through activation of K⫹ currents and inhibition of Ca2⫹ currents via N/OFQ peptide
receptors (135). This, unlike dynorphin, is a long-lasting effect
that can last for over an hour (M. P. Parsons, unpublished
observation).
Glutamate, a classical neurotransmitter, can be expected to
be released with less intense activity compared with peptide
transmitters (21). Thus, it is tempting to speculate that the
intensity and duration of firing activity dictate the release and
postsynaptic response to the four coneurotransmitters of orexin
neurons. At low-to-moderate activity levels, orexin neurons
AJP-Regul Integr Comp Physiol • VOL
will preferentially release glutamate and increase the excitability of postsynaptic orexin neurons. With higher firing frequencies, glutamate release becomes depressed (133), while peptide
release would be favored. This will initially induce an inhibitory postsynaptic response due to the inhibitory actions of
dynorphin and N/OFQ, masking the excitatory orexin effect.
However, during a prolonged firing activity, ␬-receptors will
be desensitized and the excitatory orexin effect will eventually
prevail. Nonetheless, the excitability will be kept in check due
to the nondesensitizing N/OFQ effect. It has been postulated
that the dynamic balance of orexin neuron’s excitatory and
inhibitory neurotransmitters underlie the apparent lag between
the electrophysiological activity and functional outcome (128).
Specifically, on one hand, the timing of orexin neuron’s spiking activity precedes wakefulness (64, 75), which suggests that
orexins should affect every wake bout. On the other hand,
orexins are only effective at maintaining sustained wake bouts
and not brief wake bouts of ⬍ 1 min (25). Perhaps the initial
glutamate release is not sufficient to fully recruit and maintain
the activity of a large number of orexin neurons, and the orexin
effect needs to be unmasked from the inhibitory dynorphin for
the maintenance of wakefulness.
Excitatory synaptic inputs are also subject to negative regulation (Fig. 1B). Dynorphin acts on presynaptic excitatory
terminals to attenuate glutamate release (67, 135), while
N/OFQ inhibits both excitatory and inhibitory transmission
(135). Furthermore, synaptically released glutamate negatively
regulates the presynaptic release of glutamate and GABA
through group III metabotropic glutamate receptors (mGluRs)
(2) (Fig. 1C). This would primarily be an autoinhibitory mechanism that limits glutamatergic transmission, since the inhibitory mGluRs on excitatory terminals (autoreceptors) are tonically activated by endogenous glutamate, whereas those expressed on GABAergic terminals seem to require more intense
synaptic activation and glutamate spillover to be activated (2).
Endocannabinoids, on the other hand, are typically released by
the postsynaptic neuron upon a strong depolarization and act as
a retrograde messenger (5). In orexin neurons, postsynaptic
depolarization results in an inhibition of glutamatergic transmission (depolarization-induced suppression of excitation) but
not GABAergic transmission, subsequently hyperpolarizing
the postsynaptic cell (58) (Fig. 1D). These processes are
designed not to turn on when the excitatory synapse or the
postsynaptic cell is quiescent, thus providing negative feedback mechanisms to prevent overexcitation.
Intra-LH/PFA Neuronal Network
Several neuronal populations have been identified within the
LH/PFA that are distinct from the orexin-expressing population. There is no doubt that these neurons also play important
roles in the LH/PFA functions, including shaping the output
and/or mediating the functions of orexin neurons (Fig. 3).
Melanin-concentrating hormone neurons. Melanin-concentrating hormone (MCH) neurons are concentrated in the LH/
PFA as well as in the zona incerta, where they are densely
intermingled with orexin neurons but constitute a distinct cell
group (14, 30, 90, 121). These neurons have also been the
subject of intense research because of the crucial roles of MCH
in promoting positive energy balance (91, 95, 108), sleep (3,
51, 123), and anxiety (11, 37). MCH neurons, like orexin
301 • SEPTEMBER 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.32.246 on June 17, 2017
membrane, while having little effect on synaptic inputs (131).
Adenosine, an inhibitory purine, is released endogenously onto
orexin neurons by stimulation of afferent fibers at a frequency
as low as 10 Hz in vitro (133), and an A1-receptor antagonist
has been shown to increase wakefulness when injected locally
into the LH/PFA in vivo (114). Whether or not neurons are the
source of adenosine in this case remains unresolved, as adenosine is thought to be primarily produced by hydrolysis of
astrocyte-derived ATP in the extracellular space by ectonucleotidases (124, 142) and affect neuronal activity (88). In orexin
neurons, adenosine inhibits voltage-gated Ca2⫹ currents (69)
and suppresses firing activity by A1-receptor-mediated reduction of presynaptic glutamate release (69, 133). The ratio of the
two purines would depend on the rate of ATP release, conversion of ATP to adenosine, and uptake by nucleoside transporters (124) and would determine the polarity of the effect
(excitatory or inhibitory) on orexin neurons. This may have
important functional implications in sleep homeostasis (44,
114) and survival of LH neurons (141).
In summary, there are a number of local mechanisms that
drive orexin neurons to be in a sustained active state. These
include intrinsic ion channels (transcient receptor potential C
channels), paracrine and/or autocrine actions of excitatory
neurotransmitters released by orexin neurons, and neighboring
astrocytes that can amplify excitatory inputs and further stimulate orexin neurons by releasing lactate, protons, and ATP.
Astrocytes may also mediate the propagation of excitatory
signals among neurons that are not physically coupled (85,
140). These mechanisms may be important for orexin neurons
in exerting their physiological functions that require prolonged
output, such as maintaining wakefulness and stimulating consummatory behaviors.
Review
NEUROTRANSMITTER INTERACTIONS IN THE LATERAL HYPOTHALAMUS
neurons, project widely within the central nervous system (10,
14, 16, 90) and form reciprocal connections with orexin neurons within the LH/PFA (41). MCH neurons are also known to
express GABA (31, 50) as well as anorexic peptides nesfatin-1
(34, 35) and cocaine- and amphetamine-regulated transcript
(13, 29, 49). It remains to be seen whether and how these
peptides, which are functionally opposite to MCH (in terms of
feeding effect), exert electrophysiological effects on LH/PFA
neurons.
Communication between orexin and MCH neurons is bidirectional (Fig. 1E). Orexin A and B directly induce depolarization of MCH neurons (67, 122) and stimulate presynaptic
glutamate release (122), whereas dynorphin and N/OFQ have
direct hyperpolarizing effects, each through activation of
GIRK channels (67, 87). Unlike orexin neurons, not only
dynorphin (67) but also N/OFQ, induce an effect that desensitizes with repeated applications (87), while the orexin effect
does not desensitize (67). Together, the effect of coreleased
excitatory and inhibitory peptides from orexin neurons onto
MCH neurons can be expected to shift toward excitation with
AJP-Regul Integr Comp Physiol • VOL
prolonged activity, because the sustained excitatory orexin
effect outlasts the desensitizing dynorphin and N/OFQ effects.
MCH neurons also receive glutamatergic inputs, some of
which may be from orexin neurons. Fast EPSCs are mediated
by NMDA and non-NMDA receptors (122), whereas postsynaptic group I mGluRs provide another excitatory pathway to
MCH neurons that induces a slow depolarization mediated by
Na⫹/Ca2⫹ exchanger and potentiation of NMDA currents (57).
Thus, somewhat akin to orexin neurons, MCH neurons are
regulated by a balance between excitatory (orexins and glutamate) and inhibitory (dynorphin and N/OFQ) neurotransmitters
originating from orexin neurons, a balance influenced by the
quantity and time course of each transmitter release.
On the other hand, MCH released onto orexin neurons can
act as a gatekeeper to prevent excess excitatory inputs. MCH
does not have any apparent presynaptic effect on its own, but
can attenuate presynaptic glutamate release induced by activation of orexin and D1-like receptors (93). At the postsynaptic
side, MCH downregulates AMPA receptors in orexin neurons
(93). Therefore, in response to the excitatory input from orexin
neurons, MCH neurons send a negative feedback signal via
inhibitory neurotransmitters MCH and GABA. Such mechanisms may be underlying the reciprocal firing activities of
MCH neurons and orexin neurons observed in vivo (51).
MCH neurons can also self-regulate their activity levels. The
MCH peptide does not affect the resting membrane potential of
MCH neurons, but inhibits voltage-gated Ca2⫹ channels (36).
Thus, it is possible that MCH can negatively regulate its own
release via modulation of Ca2⫹ channels expressed at the axon
terminals. GABA also conveys important information to MCH
neurons, because inhibitory inputs are dominant over excitatory inputs, evident from spontaneous EPSCs being relatively
scarce compared with inhibitory postsynaptic currents (58, 68)
in stark contrast to orexin neurons that receive dominant
excitatory inputs (56). In neonate rodents, GABAA currents are
depolarizing until about postnatal days 8 –9, providing the
main excitatory inputs to MCH neurons (68). In mature MCH
neurons, the GABAA current becomes inhibitory and mediates
a tonic inhibitory tone by endogenous GABA (58, 68). In
addition, the GABAB receptor agonist baclofen has been observed to hyperpolarize a subpopulation of MCH neurons (n ⫽
2 out of 6 cells examined; C. O. Alberto, unpublished data).
Endocannabinoids are also released by depolarization of
MCH neurons, which inhibit presynaptic GABA release (depolarization-induced suppression of inhibition). This works as
a positive feedback, since the postsynaptic cell is consequently
disinhibited from a tonic GABAA tone (58, 59). Although
endocannabinoids also inhibit glutamate release to MCH neurons, depolarization-induced suppression of inhibition seems to
have an overwhelming influence (58). Since depolarizationinduced endocannabinoid release is dependent on Ca2⫹ currents (59), a concurrent MCH release (although likely from
different subcellular locations i.e., dendrites vs. terminals respectively), resulting in an inhibition of Ca2⫹ channels may put
a break on the positive feedback by endocannabinoids.
Leptin receptor-expressing GABAergic neurons. Another
distinct neuronal population that exists in the LH/PFA is the
leptin receptor-expressing (LepRb⫹) neurons (65). These neurons are excited by leptin, use GABA as a neurotransmitter
(65), and make synaptic contacts with orexin neurons but not
MCH neurons (70). This suggests that leptin not only inhibits
301 • SEPTEMBER 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.32.246 on June 17, 2017
Fig. 3. Local network within the lateral hypothalamus/perifornical area (LH/PFA).
A number of neuronal populations have been identified within the LH/PFA in
addition to orexin (Orx) neurons, including those expressing melaninconcentrating hormone (MCH), thyrotropin-releasing hormone (TRH), corticotropin-releasing factor (CRF), neurotensin (NT), galanin (Gal) and leptin receptor
(LepRb⫹)-expressing GABAergic neurons. With the exception of galanin, whose
electrophysiological effect remains unknown, these neurotransmitters have been
demonstrated to modulate the excitability of orexin neurons. Therefore, these
neurons, along with astrocytes, constitute a local network that can fine-tune the
activity levels of this brain area through complex interactions. Further investigation is necessary to fully elucidate the intricate mechanisms that control the
functional output of the LH/PFA in physiological and pathological conditions.
R575
Review
R576
NEUROTRANSMITTER INTERACTIONS IN THE LATERAL HYPOTHALAMUS
AJP-Regul Integr Comp Physiol • VOL
lar level, TRH has been found to directly depolarize orexin
neurons by activating nonselective cation channels (39, 47).
However, TRH also increases the firing activity of presynaptic
GABAergic interneurons, increasing the inhibitory synaptic
influence (47). Therefore, the location of synaptic release of
TRH would likely determine whether TRH excites or inhibits
orexin neurons.
It is interesting that these neuropeptides that excite orexin
neurons are known to promote energy expenditure and inhibit
food intake [CRF (107), TRH (63), and neurotensin (72)].
Some of these effects may indeed be mediated by orexin
neurons; for example the TRH effect on locomotor activity is
reduced in orexin neuron-ablated mice (47). Unlike other
orexigenic neuropeptides that inhibit energy expenditure to
promote positive energy balance, orexins not only induce
feeding but also spontaneous physical activity (61, 81, 110,
115), sympathetic outflow (32, 33, 99, 100, 109), thermogenesis (84, 138, 139), and oxygen consumption (107). In fact, the
role of orexins in energy expenditure may be more substantial
than in food intake, because, in the absence of orexin signaling,
animals become obese despite accompanying hypophagia (46,
48). Thus, it seems plausible that orexin neurons mediate the
effects of catabolic neuropeptides expressed in the LH/PFA.
Likewise, orexins (9, 76, 126, 135), CRF (62), and TRH (12)
have been shown to have antinociceptive properties. Therefore,
orexin neurons could possibly mediate the effects of CRF and
TRH. On the other hand, N/OFQ is known to block opioidmediated stress-induced antinociception (78), which involves
suppression of orexin neurons and the downstream analgesic
effect (135).
Perspectives and Significance
Previous studies have revealed intricate interactions within
the network of neurons and astrocytes that may fine tune the
output of the LH/PFA. A given neuron can release multiple
neurotransmitters to other LH/PFA neurons, some with opposing actions on the postsynaptic cell, and the impact of each
neurotransmitter can change in a time-dependent manner.
What remains largely unknown is whether or not these cotransmitters are regulated in parallel, with respect to synthesis,
storage (are they packaged in the same vesicle?), release
(amount and site of release), and receptor expression (number
and subcellular localization).
Many of the neurotransmitters expressed in the LH/PFA are
not exclusive to this brain area, with the exception of orexins
and MCH that are largely confined. Future studies should
investigate the roles of LH/PFA-specific expression of neurotransmitters to fully elucidate the characteristics of local
neurochemical interactions as well as specific physiological
functions of the LH/PFA. Site-specific knockdown of individual neurotransmitters may prove to be useful in this regard.
Not all LH/PFA neurons make synaptic contacts onto one
another simply because of their proximity, as seen with MCH
neurons not receiving synaptic inputs from the nearby LepRb⫹
neurons (70). Nonetheless, unlike glutamate or GABA, whose
diffusion is restricted, neuropeptides can act as volume transmitters even in the absence of synaptic specialization between
the pre- and postsynaptic cell (143) and can be expected to play
an important role within the local circuitry.
301 • SEPTEMBER 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.32.246 on June 17, 2017
orexin neurons directly, as demonstrated as a reversible hyperpolarization in isolated orexin neurons (136), but also indirectly by activating these inhibitory LepRb⫹ neurons. Leptin
also increases the transcription factor STAT3 immunoreactivity in orexin neurons (43), and local action of leptin within the
LH results in a robust orexin mRNA expression after 26 h (71).
Thus, leptin induces an acute reversible inhibition as well as a
slow, long-term stimulation of the orexin system.
At inhibitory synapses to orexin neurons, GABAA receptors
mediate fast inhibitory postsynaptic currents (66), whereas
GABAB receptors mediate a slow hyperpolarization (134).
GABAB agonists also inhibit excitatory and inhibitory transmission presynaptically (134). Genetic knockout of GABAB
receptors on orexin neurons indirectly activates inhibitory
interneurons, which in turn increase the membrane conductance of orexin neurons and shunt other synaptic inputs (74).
Therefore, it seems that there is reciprocal communication
between orexin and GABAergic interneurons, where orexin
neurons activate interneurons, which in turn send negative
feedback to orexin neurons (Fig. 1F). Whether or not LepRb⫹
GABAergic neurons play a part in this reciprocal network
remains to be elucidated.
Obviously, local interneurons are not the sole source of
glutamatergic and GABAergic synaptic inputs to orexin neurons; major projections from other brain areas are also mediated by these important neurotransmitters. Sources of glutamatergic afferents, in addition to orexin neurons, include those
originating from energy balance-related arcuate proopiomelanocortin neurons (53), arousal-related basal forebrain (52), and
lateral parabrachial neurons (82), as well as the dorsomedial
hypothalamus that relays circadian rhythms from the suprachiasmatic nucleus to the LH/PFA (19, 23). MCH and LepRb⫹
neurons are one source of GABAergic inputs to LH/PFA
neurons, but other GABAergic sources include the sleep-wakerelated preoptic area (98) and basal forebrain (52), energy
balance-related neuropeptide Y, proopiomelanocortin neurons
of the arcuate nucleus (14, 30, 53–55), and emotion-related
central amygdaloid nucleus (80). Thus, to understand the
cellular mechanisms underlying the essential functions mediated by the LH/PFA, it would be important to determine how
excitatory and inhibitory synaptic inputs from various brain
areas are integrated by orexin neurons.
Other peptidergic neurons. Other neuropeptides have also
been detected in cells within the LH/PFA (Fig. 3). Galaninpositive neurons have relatively small cell bodies and are
separate from the orexin-expressing population (43, 132), although it has been shown that some orexin neurons may
coexpress galanin (43). Galanin has no effect on Ca2⫹ signaling in orexin neurons (119); however, no study has examined
the electrophysiological effect.
Corticotrophin-releasing factor (CRF) neurons located in the
LH/PFA are also distinct from orexin and MCH neurons but
coexpress neurotensin (127). CRF depolarizes a subpopulation
of orexin neurons (25%) via CRF1 receptors. The effect is
reversible but nondesensitizing as long as the ligand is present
(130). Neurotensin has been shown to activate Ca2⫹ signaling
(119).
A significant number of neurons expressing thyrotropinreleasing hormone (TRH) are also found in the LH/PFA
according to the Allen Institute for Brain Science Atlas (http://
mouse.brain-map.org/brain/gene/71016631.html). At the cellu-
Review
NEUROTRANSMITTER INTERACTIONS IN THE LATERAL HYPOTHALAMUS
These local interactions may also affect the sensitivity of
LH/PFA neurons to incoming projections from other brain
areas (15). The functioning of the LH/PFA circuitry is further
complicated by functional plasticity and dynamic synaptic
remodeling of orexin neurons that occurs in response to changing physiological needs, as seen following sleep (40, 77, 92,
120) and food deprivation (56) and in relation to circadian
rhythm (6). It is highly likely that the intra-LH/PFA network
activity discussed here would be altered as a result. Further
research on neurotransmitter effects and synaptic integration in
various physiological and pathological contexts will help provide a clearer picture of the regulatory mechanisms for orexin
neurons and the LH/PFA.
14.
15.
16.
17.
The authors thank Dr. Jackie Vanderluit for the critical reading of the
manuscript.
18.
GRANTS
19.
This work was supported by funding from the Canadian Institutes of Health
Research.
20.
DISCLOSURES
No conflicts of interest, financial or otherwise, are declared by the author(s).
21.
REFERENCES
22.
1. Abrahamson EE, Leak RK, Moore RY. The suprachiasmatic nucleus
projects to posterior hypothalamic arousal systems. Neuroreport 12:
435–440, 2001.
2. Acuna-Goycolea C, Li Y, van Den Pol AN. Group III metabotropic
glutamate receptors maintain tonic inhibition of excitatory synaptic input
to hypocretin/orexin neurons. J Neurosci 24: 3013–3022, 2004.
3. Ahnaou A, Drinkenburg WH, Bouwknecht JA, Alcazar J, Steckler
T, Dautzenberg FM. Blocking melanin-concentrating hormone MCH1
receptor affects rat sleep-wake architecture. Eur J Pharmacol 579:
177–188, 2008.
4. Alberto CO, Hirasawa M. AMPA receptor-mediated miniature EPSCs
have heterogeneous time courses in orexin neurons. Biochem Biophys
Res Commun 400: 707–712, 2010.
5. Alger BE. Endocannabinoids: getting the message across. Proc Natl
Acad Sci USA 101: 8512–8513, 2004.
6. Appelbaum L, Wang G, Yokogawa T, Skariah GM, Smith SJ,
Mourrain P, Mignot E. Circadian and homeostatic regulation of structural synaptic plasticity in hypocretin neurons. Neuron 68: 87–98, 2010.
7. Aston-Jones G, Smith RJ, Sartor GC, Moorman DE, Massi L,
Tahsili-Fahadan P, Richardson KA. Lateral hypothalamic orexin/
hypocretin neurons: a role in reward-seeking and addiction. Brain Res
1314: 74 –90, 2010.
8. Berthoud HR. Multiple neural systems controlling food intake and body
weight. Neurosci Biobehav Rev 26: 393–428, 2002.
9. Bingham S, Davey PT, Babbs AJ, Irving EA, Sammons MJ, Wyles
M, Jeffrey P, Cutler L, Riba I, Johns A, Porter RA, Upton N, Hunter
AJ, Parsons AA. Orexin-A, an hypothalamic peptide with analgesic
properties. Pain 92: 81–90, 2001.
10. Bittencourt JC, Presse F, Arias C, Peto C, Vaughan J, Nahon JL,
Vale W, Sawchenko PE. The melanin-concentrating hormone system of
the rat brain: an immuno- and hybridization histochemical characterization. J Comp Neurol 319: 218 –245, 1992.
11. Borowsky B, Durkin MM, Ogozalek K, Marzabadi MR, DeLeon J,
Lagu B, Heurich R, Lichtblau H, Shaposhnik Z, Daniewska I,
Blackburn TP, Branchek TA, Gerald C, Vaysse PJ, Forray C.
Antidepressant, anxiolytic and anorectic effects of a melanin-concentrating hormone-1 receptor antagonist. Nat Med 8: 825–830, 2002.
12. Boschi G, Desiles M, Reny V, Rips R, Wrigglesworth S. Antinociceptive properties of thyrotropin releasing hormone in mice: comparison
with morphine. Br J Pharmacol 79: 85–92, 1983.
13. Broberger C. Hypothalamic cocaine- and amphetamine-regulated transcript (CART) neurons: histochemical relationship to thyrotropin-releasAJP-Regul Integr Comp Physiol • VOL
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
ing hormone, melanin-concentrating hormone, orexin/hypocretin and
neuropeptide Y. Brain Res 848: 101–113, 1999.
Broberger C, de Lecea L, Sutcliffe JG, Hokfelt T. Hypocretin/orexinand melanin-concentrating hormone-expressing cells form distinct populations in the rodent lateral hypothalamus: relationship to the neuropeptide Y and agouti gene-related protein systems. J Comp Neurol 402:
460 –474, 1998.
Carter ME, Adamantidis A, Ohtsu H, Deisseroth K, de Lecea L.
Sleep Homeostasis Modulates Hypocretin-Mediated Sleep-to-Wake
Transitions. J Neurosci 29: 10939 –10949, 2009.
Chemelli RM, Willie JT, Sinton CM, Elmquist JK, Scammell T, Lee
C, Richardson JA, Williams SC, Xiong Y, Kisanuki Y, Fitch TE,
Nakazato M, Hammer RE, Saper CB, Yanagisawa M. Narcolepsy in
orexin knockout mice: molecular genetics of sleep regulation. Cell 98:
437–451, 1999.
Chiou LC, Lee HJ, Ho YC, Chen SP, Liao YY, Ma CH, Fan PC, Fuh
JL, Wang SJ. Orexins/hypocretins: pain regulation and cellular actions.
Curr Pharm Des 16: 3089 –3100, 2010.
Chou TC, Lee CE, Lu J, Elmquist JK, Hara J, Willie JT, Beuckmann
CT, Chemelli RM, Sakurai T, Yanagisawa M, Saper CB, Scammell
TE. Orexin (hypocretin) neurons contain dynorphin. J Neurosci 21:
RC168, 2001.
Chou TC, Scammell TE, Gooley JJ, Gaus SE, Saper CB, Lu J.
Critical role of dorsomedial hypothalamic nucleus in a wide range of
behavioral circadian rhythms. J Neurosci 23: 10691–10702, 2003.
Cvetkovic-Lopes V, Eggermann E, Uschakov A, Grivel J, Bayer L,
Jones BE, Serafin M, Muhlethaler M. Rat hypocretin/orexin neurons
are maintained in a depolarized state by TRPC channels. PLoS One 5:
e15673, 2010.
De Camilli P, Jahn R. Pathways to regulated exocytosis in neurons.
Annu Rev Physiol 52: 625–645, 1990.
de Lecea L, Kilduff TS, Peyron C, Gao X, Foye PE, Danielson PE,
Fukuhara C, Battenberg EL, Gautvik VT, Bartlett FS, Frankel WN,
van Den Pol AN, Bloom FE, Gautvik KM, Sutcliffe JG. The hypocretins: hypothalamus-specific peptides with neuroexcitatory activity.
Proc Natl Acad Sci USA 95: 322–327, 1998.
Deurveilher S, Semba K. Indirect projections from the suprachiasmatic
nucleus to major arousal-promoting cell groups in rat: implications for
the circadian control of behavioural state. Neuroscience 130: 165–183,
2005.
DiLeone RJ, Georgescu D, Nestler EJ. Lateral hypothalamic neuropeptides in reward and drug addiction. Life Sci 73: 759 –768, 2003.
Diniz Behn CG, Kopell N, Brown EN, Mochizuki T, Scammell TE.
Delayed orexin signaling consolidates wakefulness and sleep: physiology
and modeling. J Neurophysiol 99: 3090 –3103, 2008.
Druzin M, Haage D, Malinina E, Johansson S. Dual and opposing
roles of presynaptic Ca2⫹ influx for spontaneous GABA release from rat
medial preoptic nerve terminals. J Physiol 542: 131–146, 2002.
Dube MG, Kalra SP, Kalra PS. Food intake elicited by central administration of orexins/hypocretins: identification of hypothalamic sites of
action. Brain Res 842: 473–477, 1999.
Eggermann E, Bayer L, Serafin M, Saint-Mleux B, Bernheim L,
Machard D, Jones BE, Muhlethaler M. The wake-promoting hypocretin-orexin neurons are in an intrinsic state of membrane depolarization. J
Neurosci 23: 1557–1562, 2003.
Elias CF, Lee CE, Kelly JF, Ahima RS, Kuhar M, Saper CB,
Elmquist JK. Characterization of CART neurons in the rat and human
hypothalamus. J Comp Neurol 432: 1–19, 2001.
Elias CF, Saper CB, Maratos-Flier E, Tritos NA, Lee C, Kelly J,
Tatro JB, Hoffman GE, Ollmann MM, Barsh GS, Sakurai T, Yanagisawa M, Elmquist JK. Chemically defined projections linking the
mediobasal hypothalamus and the lateral hypothalamic area. J Comp
Neurol 402: 442–459, 1998.
Elias CF, Sita LV, Zambon BK, Oliveira ER, Vasconcelos LA,
Bittencourt JC. Melanin-concentrating hormone projections to areas
involved in somatomotor responses. J Chem Neuroanat 35: 188 –201,
2008.
Ferguson AV, Samson WK. The orexin/hypocretin system: a critical
regulator of neuroendocrine and autonomic function. Front Neuroendocrinol 24: 141–150, 2003.
Follwell MJ, Ferguson AV. Cellular mechanisms of orexin actions on
paraventricular nucleus neurones in rat hypothalamus. J Physiol 545:
855–867, 2002.
301 • SEPTEMBER 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.32.246 on June 17, 2017
ACKNOWLEDGMENTS
R577
Review
R578
NEUROTRANSMITTER INTERACTIONS IN THE LATERAL HYPOTHALAMUS
AJP-Regul Integr Comp Physiol • VOL
54. Horvath TL, Bechmann I, Naftolin F, Kalra SP, 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.
55. Horvath TL, Diano S, van Den Pol AN. Synaptic interaction between
hypocretin (orexin) and neuropeptide Y cells in the rodent and primate
hypothalamus: a novel circuit implicated in metabolic and endocrine
regulations. J Neurosci 19: 1072–1087, 1999.
56. Horvath TL, Gao XB. Input organization and plasticity of hypocretin
neurons: possible clues to obesity’s association with insomnia. Cell
Metab 1: 279 –286, 2005.
57. Huang H, van Den Pol AN. Rapid direct excitation and long-lasting
enhancement of NMDA response by group I metabotropic glutamate
receptor activation of hypothalamic melanin-concentrating hormone neurons. J Neurosci 27: 11560 –11572, 2007.
58. Huang H, Acuna-Goycolea C, Li Y, Cheng HM, Obrietan K, van den
Pol AN. Cannabinoids excite hypothalamic melanin-concentrating hormone but inhibit hypocretin/orexin neurons: implications for cannabinoid
actions on food intake and cognitive arousal. J Neurosci 27: 4870 –4881,
2007.
59. Jo YH, Chen YJ, Chua SC Jr, Talmage DA, Role LW. Integration of
endocannabinoid and leptin signaling in an appetite-related neural circuit.
Neuron 48: 1055–1066, 2005.
60. Kotz CM, Teske JA, Levine JA, Wang C. Feeding and activity induced
by orexin A in the lateral hypothalamus in rats. Regul Pept 104: 27–32,
2002.
61. Kotz CM, Wang C, Teske JA, Thorpe AJ, Novak CM, Kiwaki K,
Levine JA. Orexin a mediation of time spent moving in rats: neural
mechanisms. Neuroscience 142: 29 –36, 2006.
62. Lariviere WR, Melzack R. The role of corticotropin-releasing factor in
pain and analgesia. Pain 84: 1–12, 2000.
63. Lechan RM, Fekete C. The TRH neuron: a hypothalamic integrator of
energy metabolism. Prog Brain Res 153: 209 –235, 2006.
64. Lee MG, Hassani OK, Jones BE. Discharge of identified orexin/
hypocretin neurons across the sleep-waking cycle. J Neurosci 25: 6716 –
6720, 2005.
65. Leinninger GM, Jo YH, Leshan RL, Louis GW, Yang H, Barrera
JG, Wilson H, Opland DM, Faouzi MA, Gong Y, Jones JC, Rhodes
CJ, Chua S Jr, Diano S, Horvath TL, Seeley RJ, Becker JB, Munzberg H, Myers MG Jr. Leptin acts via leptin receptor-expressing
lateral hypothalamic neurons to modulate the mesolimbic dopamine
system and suppress feeding. Cell Metab 10: 89 –98, 2009.
66. Li Y, Gao XB, Sakurai T, van Den Pol AN. Hypocretin/orexin excites
hypocretin neurons via a local glutamate neuron-A potential mechanism
for orchestrating the hypothalamic arousal system. Neuron 36: 1169 –
1181, 2002.
67. Li Y, van Den Pol AN. Differential target-dependent actions of coexpressed inhibitory dynorphin and excitatory hypocretin/orexin neuropeptides. J Neurosci 26: 13037–13047, 2006.
68. Li Y, van den Pol AN. Enhanced excitatory input to melanin concentrating hormone neurons during developmental period of high food intake
is mediated by GABA. J Neurosci 29: 15195–15204, 2009.
69. Liu ZW, Gao XB. Adenosine inhibits activity of hypocretin/orexin
neurons by the A1-receptor in the lateral hypothalamus: a possible
sleep-promoting effect. J Neurophysiol 97: 837–848, 2007.
70. Louis GW, Leinninger GM, Rhodes CJ, Myers MG Jr. Direct
innervation and modulation of orexin neurons by lateral hypothalamic
LepRb neurons. J Neurosci 30: 11278 –11287, 2010.
71. Louis GW, Leinninger GM, Rhodes CJ, Myers MG Jr. Direct
innervation and modulation of orexin neurons by lateral hypothalamic
LepRb Neurons. J Neurosci 30: 11278 –11287, 2010.
72. Luttinger D, King RA, Sheppard D, Strupp J, Nemeroff CB, Prange
J. The effect of neurotensin on food consumption in the rat. Eur J
Pharmacol 81: 499 –503, 1982.
73. Maolood N, Meister B. Nociceptin/orphanin FQ peptide in hypothalamic neurones associated with the control of feeding behaviour. J
Neuroendocrinol 22: 75–82, 2010.
74. Matsuki T, Nomiyama M, Takahira H, Hirashima N, Kunita S,
Takahashi S, Yagami K, Kilduff TS, Bettler B, Yanagisawa M,
Sakurai T. Selective loss of GABAB receptors in orexin-producing
neurons results in disrupted sleep/wakefulness architecture. Proc Natl
Acad Sci USA 106: 4459 –4464, 2009.
301 • SEPTEMBER 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.32.246 on June 17, 2017
34. Foo KS, Brismar H, Broberger C. Distribution and neuropeptide
coexistence of nucleobindin-2 mRNA/nesfatin-like immunoreactivity in
the rat CNS. Neuroscience 156: 563–579, 2008.
35. Fort P, Salvert D, Hanriot L, Jego S, Shimizu H, Hashimoto K, Mori
M, Luppi PH. The satiety molecule nesfatin-1 is co-expressed with
melanin concentrating hormone in tuberal hypothalamic neurons of the
rat. Neuroscience 155: 174 –181, 2008.
36. Gao XB, Ghosh PK, van Den Pol AN. Neurons synthesizing melaninconcentrating hormone identified by selective reporter gene expression
after transfection in vitro: transmitter responses. J Neurophysiol 90:
3978 –3985, 2003.
37. Georgescu D, Sears RM, Hommel JD, Barrot M, Bolanos CA, Marsh
DJ, Bednarek MA, Bibb JA, Maratos-Flier E, Nestler EJ, DiLeone
RJ. The hypothalamic neuropeptide melanin-concentrating hormone acts
in the nucleus accumbens to modulate feeding behavior and forced-swim
performance. J Neurosci 25: 2933–2940, 2005.
38. Georgescu D, Zachariou V, Barrot M, Mieda M, Willie JT, Eisch AJ,
Yanagisawa M, Nestler EJ, DiLeone RJ. Involvement of the lateral
hypothalamic peptide orexin in morphine dependence and withdrawal. J
Neurosci 23: 3106 –3111, 2003.
39. Gonzalez JA, Horjales-Araujo E, Fugger L, Broberger C, Burdakov
D. Stimulation of orexin/hypocretin neurones by thyrotropin-releasing
hormone. J Physiol 587: 1179 –1186, 2009.
40. Grivel J, Cvetkovic V, Bayer L, Machard D, Tobler I, Muhlethaler
M, Serafin M. The wake-promoting hypocretin/orexin neurons change
their response to noradrenaline after sleep deprivation. J Neurosci 25:
4127–4130, 2005.
41. Guan JL, Uehara K, Lu S, Wang QP, Funahashi H, Sakurai T,
Yanagizawa M, Shioda S. Reciprocal synaptic relationships between
orexin- and melanin-concentrating hormone-containing neurons in the rat
lateral hypothalamus: a novel circuit implicated in feeding regulation. Int
J Obes Relat Metab Disord 26: 1523–1532, 2002.
42. Guthrie PB, Knappenberger J, Segal M, Bennett MVL, Charles AC,
Kater SB. ATP released from astrocytes mediates glial calcium waves.
J Neurosci 19: 520 –528, 1999.
43. Hakansson M, de Lecea L, Sutcliffe JG, Yanagisawa M, Meister B.
Leptin receptor- and STAT3-immunoreactivities in hypocretin/orexin
neurones of the lateral hypothalamus. J Neuroendocrinol 11: 653–663,
1999.
44. Halassa MM, Florian C, Fellin T, Munoz JR, Lee SY, Abel T,
Haydon PG, Frank MG. Astrocytic modulation of sleep homeostasis
and cognitive consequences of sleep loss. Neuron 61: 213–219, 2009.
45. Halestrap AP, Price NT. The proton-linked monocarboxylate transporter (MCT) family: structure, function and regulation. Biochem J 343:
281–299, 1999.
46. Hara J, Beuckmann CT, Nambu T, Willie JT, Chemelli RM, Sinton
CM, Sugiyama F, Yagami K, Goto K, Yanagisawa M, Sakurai T.
Genetic ablation of orexin neurons in mice results in narcolepsy, hypophagia, and obesity. Neuron 30: 345–354, 2001.
47. Hara J, Gerashchenko D, Wisor JP, Sakurai T, Xie X, Kilduff TS.
Thyrotropin-releasing hormone increases behavioral arousal through
modulation of hypocretin/orexin neurons. J Neurosci 29: 3705–3714,
2009.
48. Hara J, Yanagisawa M, Sakurai T. Difference in obesity phenotype
between orexin-knockout mice and orexin neuron-deficient mice with
same genetic background and environmental conditions. Neurosci Lett
380: 239 –242, 2005.
49. Harthoorn LF. Projection-dependent differentiation of melanin-concentrating hormone-containing neurons. Cell Mol Neurobiol 27: 49 –55,
2007.
50. Harthoorn LF, Sane A, Nethe M, Van Heerikhuize JJ. Multi-transcriptional profiling of melanin-concentrating hormone and orexin-containing neurons. Cell Mol Neurobiol 25: 1209 –1223, 2005.
51. Hassani OK, Lee MG, Jones BE. Melanin-concentrating hormone
neurons discharge in a reciprocal manner to orexin neurons across the
sleep-wake cycle. Proc Natl Acad Sci USA 106: 2418 –2422, 2009.
52. Henny P, Jones BE. Innervation of orexin/hypocretin neurons by
GABAergic, glutamatergic or cholinergic basal forebrain terminals evidenced by immunostaining for presynaptic vesicular transporter and
postsynaptic scaffolding proteins. J Comp Neurol 499: 645–661, 2006.
53. Hentges ST, Otero-Corchon V, Pennock RL, King CM, Low MJ.
Proopiomelanocortin expression in both GABA and glutamate neurons. J
Neurosci 29: 13684 –13690, 2009.
Review
NEUROTRANSMITTER INTERACTIONS IN THE LATERAL HYPOTHALAMUS
AJP-Regul Integr Comp Physiol • VOL
97.
98.
99.
100.
101.
102.
103.
104.
105.
106.
107.
108.
109.
110.
111.
112.
113.
114.
115.
116.
117.
118.
119.
WS, Terrett JA, Elshourbagy NA, Bergsma DJ, Yanagisawa M.
Orexins and orexin receptors: a family of hypothalamic neuropeptides
and G protein-coupled receptors that regulate feeding behavior. Cell 92:
573–585, 1998.
Sakurai T, Mieda M, Tsujino N. The orexin system: roles in sleep/wake
regulation. Ann NY Acad Sci 1200: 149 –161, 2010.
Sakurai T, Nagata R, Yamanaka A, Kawamura H, Tsujino N,
Muraki Y, Kageyama H, Kunita S, Takahashi S, Goto K, Koyama Y,
Shioda S, Yanagisawa M. Input of orexin/hypocretin neurons revealed
by a genetically encoded tracer in mice. Neuron 46: 297–308, 2005.
Samson WK, Taylor MM, Ferguson AV. Non-sleep effects of hypocretin/orexin. Sleep Med Rev 9: 243–252, 2005.
Samson WK, Taylor MM, Follwell M, Ferguson AV. Orexin actions
in hypothalamic paraventricular nucleus: physiological consequences
and cellular correlates. Regul Pept 104: 97–103, 2002.
Saper CB. Organization of cerebral cortical afferent systems in the rat.
II. Hypothalamocortical projections. J Comp Neurol 237: 21–46, 1985.
Saper CB, Akil H, Watson SJ. Lateral hypothalamic innervation of the
cerebral cortex: immunoreactive staining for a peptide resembling but
immunochemically distinct from pituitary/arcuate ␣-melanocyte stimulating hormone. Brain Res Bull 16: 107–120, 1986.
Saper CB, Chou TC, Elmquist JK. The need to feed: homeostatic and
hedonic control of eating. Neuron 36: 199 –211, 2002.
Saper CB, Swanson LW, Cowan WM. An autoradiographic study of
the efferent connections of the lateral hypothalamic area in the rat. J
Comp Neurol 183: 689 –706, 1979.
Saper CB, Fuller PM, Pedersen NP, Lu J, Scammell TE. Sleep state
switching. Neuron 68: 1023–1042, 2010.
Saper CB, Scammell TE, Lu J. Hypothalamic regulation of sleep and
circadian rhythms. Nature 437: 1257–1263, 2005.
Semjonous NM, Smith KL, Parkinson JR, Gunner DJ, Liu YL,
Murphy KG, Ghatei MA, Bloom SR, Small CJ. Coordinated changes
in energy intake and expenditure following hypothalamic administration
of neuropeptides involved in energy balance. Int J Obes (Lond) 33:
775–785, 2009.
Shimada M, Tritos NA, Lowell BB, Flier JS, Maratos-Flier E. Mice
lacking melanin-concentrating hormone are hypophagic and lean. Nature
396: 670 –674, 1998.
Smith PM, Connolly BC, Ferguson AV. Microinjection of orexin into
the rat nucleus tractus solitarius causes increases in blood pressure. Brain
Res 950: 261–267, 2002.
Sunter D, Morgan I, Edwards CM, Dakin CL, Murphy KG, Gardiner J, Taheri S, Rayes E, Bloom SR. Orexins: effects on behavior and
localisation of orexin receptor 2 messenger ribonucleic acid in the rat
brainstem. Brain Res 907: 27–34, 2001.
Sweet DC, Levine AS, Billington CJ, Kotz CM. Feeding response to
central orexins. Brain Res 821: 535–538, 1999.
Teske JA, Billington CJ, Kotz CM. Hypocretin/orexin and energy
expenditure. Acta Physiologica 198: 303–312, 2010.
Teske JA, Levine AS, Kuskowski M, Levine JA, Kotz CM. Elevated
hypothalamic orexin signaling, sensitivity to orexin A and spontaneous
physical activity in obesity resistant rats. Am J Physiol Regul Integr
Comp Physiol 291: R889 –R899, 2006.
Thakkar MM, Engemann SC, Walsh KM, Sahota PK. Adenosine and
the homeostatic control of sleep: effects of A1-receptor blockade in the
perifornical lateral hypothalamus on sleep-wakefulness. Neuroscience
153: 875–880, 2008.
Thorpe AJ, Kotz CM. Orexin A in the nucleus accumbens stimulates
feeding and locomotor activity. Brain Res 1050: 156 –162, 2005.
Thorpe AJ, Mullett MA, Wang C, Kotz CM. Peptides that regulate
food intake: regional, metabolic, and circadian specificity of lateral
hypothalamic orexin A feeding stimulation. Am J Physiol Regul Integr
Comp Physiol 284: R1409 –R1417, 2003.
Torrealba F, Yanagisawa M, Saper CB. Colocalization of orexin a and
glutamate immunoreactivity in axon terminals in the tuberomammillary
nucleus in rats. Neuroscience 119: 1033–1044, 2003.
Trivedi P, Yu H, MacNeil DJ, Van der Ploeg LH, Guan XM.
Distribution of orexin receptor mRNA in the rat brain. FEBS Lett 438:
71–75, 1998.
Tsujino N, Yamanaka A, Ichiki K, Muraki Y, Kilduff TS, Yagami K,
Takahashi S, Goto K, Sakurai T. Cholecystokinin activates orexin/
hypocretin neurons through the cholecystokinin A receptor. J Neurosci
25: 7459 –7469, 2005.
301 • SEPTEMBER 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.32.246 on June 17, 2017
75. Mileykovskiy BY, Kiyashchenko LI, Siegel JM. Behavioral correlates
of activity in identified hypocretin/orexin neurons. Neuron 46: 787–798,
2005.
76. Mobarakeh JI, Takahashi K, Sakurada S, Nishino S, Watanabe H,
Kato M, Yanai K. Enhanced antinociception by intracerebroventricularly and intrathecally administered orexin A and B (hypocretin-1 and -2)
in mice. Peptides 26: 767–777, 2005.
77. Modirrousta M, Mainville L, Jones BE. Orexin and MCH neurons
express c-Fos differently after sleep deprivation vs. recovery and bear
different adrenergic receptors. Eur J Neurosci 21: 2807–2816, 2005.
78. Mogil JS, Grisel JE, Reinscheid RK, Civelli O, Belknap JK, Grandy
DK. Orphanin FQ is a functional anti-opioid peptide. Neuroscience 75:
333–337, 1996.
79. Mullett MA, Billington CJ, Levine AS, Kotz CM. Hypocretin I in the
lateral hypothalamus activates key feeding-regulatory brain sites. Neuroreport 11: 103–108, 2000.
80. Nakamura S, Tsumori T, Yokota S, Oka T, Yasui Y. Amygdaloid
axons innervate melanin-concentrating hormone- and orexin-containing
neurons in the mouse lateral hypothalamus. Brain Res 1278: 66 –74,
2009.
81. Nakamura T, Uramura K, Nambu T, Yada T, Goto K, Yanagisawa
M, Sakurai T. Orexin-induced hyperlocomotion and stereotypy are
mediated by the dopaminergic system. Brain Res 873: 181–187, 2000.
82. Niu JG, Yokota S, Tsumori T, Qin Y, Yasui Y. Glutamatergic lateral
parabrachial neurons innervate orexin-containing hypothalamic neurons
in the rat. Brain Res 1358: 110 –122, 2010.
83. Ohno K, Sakurai T. Orexin neuronal circuitry: role in the regulation of
sleep and wakefulness. Front Neuroendocrinol 29: 70 –87, 2008.
84. Oldfield BJ, Giles ME, Watson A, Anderson C, Colvill LM, McKinley
MJ. The neurochemical characterisation of hypothalamic pathways projecting polysynaptically to brown adipose tissue in the rat. Neuroscience
110: 515–526, 2002.
85. Parri HR, Gould TM, Crunelli V. Spontaneous astrocytic Ca2⫹ oscillations in situ drive NMDAR-mediated neuronal excitation. Nat Neurosci
4: 803–812, 2001.
86. Parsons MP, Hirasawa M. ATP-sensitive potassium channel-mediated
lactate effect on orexin neurons: implications for brain energetics during
arousal. J Neurosci 30: 8061–8070, 2010.
87. Parsons MP, Hirasawa M. GIRK channel-mediated inhibition of melanin-concentrating hormone neurons by nociceptin/orphanin FQ. J Neurophysiol 105: 1179 –1184, 2011.
88. Pascual O, Casper KB, Kubera C, Zhang J, Revilla-Sanchez R, Sul
JY, Takano H, Moss SJ, McCarthy K, Haydon PG. Astrocytic
purinergic signaling coordinates synaptic networks. Science 310: 113–
116, 2005.
89. Pellerin L, Pellegri G, Bittar PG, Charnay Y, Bouras C, Martin JL,
Stella N, Magistretti PJ. Evidence supporting the existence of an
activity-dependent astrocyte-neuron lactate shuttle. Dev Neurosci 20:
291–299, 1998.
90. Peyron C, Tighe DK, van Den Pol AN, de Lecea L, Heller HC,
Sutcliffe JG, Kilduff TS. Neurons containing hypocretin (orexin) project to multiple neuronal systems. J Neurosci 18: 9996 –10015, 1998.
91. Qu D, Ludwig DS, Gammeltoft S, Piper M, Pelleymounter MA,
Cullen MJ, Mathes WF, Przypek R, Kanarek R, Maratos-Flier E. A
role for melanin-concentrating hormone in the central regulation of
feeding behaviour. Nature 380: 243–247, 1996.
92. Rao Y, Liu ZW, Borok E, Rabenstein RL, Shanabrough M, Lu M,
Picciotto MR, Horvath TL, Gao XB. Prolonged wakefulness induces
experience-dependent synaptic plasticity in mouse hypocretin/orexin
neurons. J Clin Invest 117: 4022–4033, 2007.
93. Rao Y, Lu M, Ge F, Marsh DJ, Qian S, Wang AH, Picciotto MR, Gao
XB. Regulation of synaptic efficacy in hypocretin/orexin-containing
neurons by melanin concentrating hormone in the lateral hypothalamus.
J Neurosci 28: 9101–9110, 2008.
94. Rosin DL, Weston MC, Sevigny CP, Stornetta RL, Guyenet PG.
Hypothalamic orexin (hypocretin) neurons express vesicular glutamate
transporters VGLUT1 or VGLUT2. J Comp Neurol 465: 593–603, 2003.
95. Rossi M, Beak SA, Choi SJ, Small CJ, Morgan DG, Ghatei MA,
Smith DM, Bloom SR. Investigation of the feeding effects of melanin
concentrating hormone on food intake–action independent of galanin and
the melanocortin receptors. Brain Res 846: 164 –170, 1999.
96. Sakurai T, Amemiya A, Ishii M, Matsuzaki I, Chemelli RM, Tanaka
H, Williams SC, Richardson JA, Kozlowski GP, Wilson S, Arch JR,
Buckingham RE, Haynes AC, Carr SA, Annan RS, McNulty DE, Liu
R579
Review
R580
NEUROTRANSMITTER INTERACTIONS IN THE LATERAL HYPOTHALAMUS
AJP-Regul Integr Comp Physiol • VOL
133.
134.
135.
136.
137.
138.
139.
140.
141.
142.
143.
hypertriglyceridemia. Am J Physiol Regul Integr Comp Physiol 284:
R1454 –R1465, 2003.
Xia J, Chen F, YEJ, Yan J, Wang H, Duan S, Hu Z. Activitydependent release of adenosine inhibits the glutamatergic synaptic transmission and plasticity in the hypothalamic hypocretin/orexin neurons.
Neuroscience 162: 980 –988, 2009.
Xie X, Crowder TL, Yamanaka A, Morairty SR, Lewinter RD,
Sakurai T, Kilduff TS. GABAB receptor-mediated modulation of hypocretin/orexin neurones in mouse hypothalamus. J Physiol 574: 399 –414,
2006.
Xie X, Wisor JP, Hara J, Crowder TL, LeWinter R, Khroyan TV,
Yamanaka A, Diano S, Horvath TL, Sakurai T, Toll L, Kilduff TS.
Hypocretin/orexin and nociceptin/orphanin FQ coordinately regulate
analgesia in a mouse model of stress-induced analgesia. J Clin Invest
118: 2471–2481, 2008.
Yamanaka A, Beuckmann CT, Willie JT, Hara J, Tsujino N, Mieda
M, Tominaga M, Yagami K, Sugiyama F, Goto K, Yanagisawa M,
Sakurai T. Hypothalamic orexin neurons regulate arousal according to
energy balance in mice. Neuron 38: 701–713, 2003.
Yamanaka A, Tabuchi S, Tsunematsu T, Fukazawa Y, Tominaga M.
Orexin directly excites orexin neurons through orexin 2 receptor. J
Neurosci 30: 12642–12652, 2010.
Yasuda T, Masaki T, Kakuma T, Hara M, Nawata T, Katsuragi I,
Yoshimatsu H. Dual regulatory effects of orexins on sympathetic nerve
activity innervating brown adipose tissue in rats. Endocrinology 146:
2744 –2748, 2005.
Yoshimichi G, Yoshimatsu H, Masaki T, Sakata T. Orexin-A regulates body temperature in coordination with arousal status. Exp Biol Med
(Maywood) 226: 468 –476, 2001.
Zhang JM, Wang HK, Ye CQ, Ge W, Chen Y, Jiang ZL, Wu CP, Poo
MM, Duan S. ATP released by astrocytes mediates glutamatergic
activity-dependent heterosynaptic suppression. Neuron 40: 971–982,
2003.
Zhang X, Zhang M, Laties AM, Mitchell CH. Balance of purines may
determine life or death of retinal ganglion cells as A3 adenosine receptors
prevent loss following P2X7 receptor stimulation. J Neurochem 98:
566 –575, 2006.
Zimmermann H, Braun N. Extracellular metabolism of nucleotides in
the nervous system. J Auton Pharmacol 16: 397–400, 1996.
Zoli M, Jansson A, Syková E, Agnati LF, Fuxe K. Volume transmission in the CNS and its relevance for neuropsychopharmacology. Trends
Pharmacol Sci 20: 142–150, 1999.
301 • SEPTEMBER 2011 •
www.ajpregu.org
Downloaded from http://ajpregu.physiology.org/ by 10.220.32.246 on June 17, 2017
120. Uschakov A, Grivel J, Cvetkovic-Lopes V, Bayer L, Bernheim L,
Jones BE, Muhlethaler M, Serafin M. Sleep-deprivation regulates ␣-2
adrenergic responses of rat hypocretin/orexin neurons. PLoS One 6:
e16672, 2011.
121. van Den Pol AN. Hypothalamic hypocretin (orexin): robust innervation
of the spinal cord. J Neurosci 19: 3171–3182, 1999.
122. van Den Pol AN, Acuna-Goycolea C, Clark KR, Ghosh PK. Physiological properties of hypothalamic MCH neurons identified with selective
expression of reporter gene after recombinant virus infection. Neuron 42:
635–652, 2004.
123. Verret L, Goutagny R, Fort P, Cagnon L, Salvert D, Leger L,
Boissard R, Salin P, Peyron C, Luppi PH. A role of melanin-concentrating hormone producing neurons in the central regulation of paradoxical sleep. BMC Neurosci 4: 19, 2003.
124. Wall M, Dale N. Activity-dependent release of adenosine: a critical
re-evaluation of mechanism. Curr Neuropharmacol 6: 329 –337, 2008.
125. Wang C, Kotz CM. Urocortin in the lateral septal area modulates
feeding induced by orexin A in the lateral hypothalamus. Am J Physiol
Regul Integr Comp Physiol 283: R358 –R367, 2002.
126. Watanabe S, Kuwaki T, Yanagisawa M, Fukuda Y, Shimoyama M.
Persistent pain and stress activate pain-inhibitory orexin pathways. Neuroreport 16: 5–8, 2005.
127. Watts AG, Sanchez-Watts G. Rapid and preferential activation of Fos
protein in hypocretin/orexin neurons following the reversal of dehydration-anorexia. J Comp Neurol 502: 768 –782, 2007.
128. Williams KS, Diniz Behn CG. Dynamic interactions between orexin
and dynorphin may delay onset of functional orexin effects: a modeling
study. J Biol Rhythms 26: 171–181, 2011.
129. Williams RH, Jensen LT, Verkhratsky A, Fugger L, Burdakov D.
Control of hypothalamic orexin neurons by acid and CO2. Proc Natl
Acad Sci USA 104: 10685–10690, 2007.
130. Winsky-Sommerer R, Yamanaka A, Diano S, Borok E, Roberts AJ,
Sakurai T, Kilduff TS, Horvath TL, de Lecea L. Interaction between
the corticotropin-releasing factor system and hypocretins (orexins): a
novel circuit mediating stress response. J Neurosci 24: 11439 –11448,
2004.
131. Wollmann G, Acuna-Goycolea C, van Den Pol AN. Direct excitation
of hypocretin/orexin cells by extracellular ATP at P2X receptors. J
Neurophysiol 94: 2195–2206, 2005.
132. Wortley KE, Chang GQ, Davydova Z, Leibowitz SF. Peptides that
regulate food intake: orexin gene expression is increased during states of