Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2015 Jul 10.
Published in final edited form as: Vitam Horm. 2012;89:91–109. doi: 10.1016/B978-0-12-394623-2.00006-8

Energy Expenditure: Role of Orexin

Jennifer A Teske *,, Vijayakumar Mavanji
PMCID: PMC4498671  NIHMSID: NIHMS704788  PMID: 22640610

Abstract

The orexins/hypocretins are endogenous, modulatory and multifunctional neuropeptides with prominent influence on several physiological processes. The influence of orexins on energy expenditure is highlighted with focus on orexin action on individual components of energy expenditure. As orexin stabilizes and maintains normal states of arousal and the sleep/wake cycle, we also highlight orexin mediation of sleep and how sleep interacts with energy expenditure.

I. Introduction

The hypocretins (hypocretin 1 and 2) are neuropeptides with significant amino acid homology to the secretin/incretin protein family that were described first by de Lecea et al. (1998) after their earlier report documented isolation of its precursor, preprohypocretin (Gautvik et al., 1996). One month later, Sakurai et al. reported identification of the same neuropeptide and associated their two G-protein-coupled receptors, however, referred to the neuropeptides as orexins (orexin A and B, OXA and OXB) due to their observation of hyperphagia after central administration (Sakurai et al., 1998). Unlike other neuropeptides, synthesis of the precursor protein, preprohypocretin or prepro-orexin, is confined to discrete regions in the lateral and lateral/perifornical hypothalamus (Gautvik et al., 1996). However, orexin protein is ubiquitously expressed in several brain nuclei, in all levels of the central nervous system and within discrete peripheral locations (Chen et al., 1999; Cutler et al., 1999; Date et al., 2000; Nambu et al., 1999; Peyron et al., 1998). Like orexin protein distribution, orexin receptors (orexin receptor 1 and 2, OX1R and OX2R) are ubiquitously distributed centrally and in discrete peripheral locations (Cluderay et al., 2002; Hervieu et al., 2001; Johren et al., 2001; Marcus et al., 2001; Trivedi et al., 1998). Of particular note, orexin receptors have been identified in human adipose tissue, and thus targeting orexin receptor-expressing adipose tissue may be an attractive therapeutic tool for metabolism-related pathologies including obesity and diabetes (Digby et al., 2006). Despite the near concurrent location of orexin terminals and receptors, OX1R and OX2R are unequally distributed within central and peripheral sites (Cluderay et al., 2002; Digby et al., 2006; Hervieu et al., 2001; Marcus et al., 2001; Trivedi et al., 1998). For instance, within discrete brain sites such as the locus coeruleus, OX1R is abundantly expressed while OX2R is relatively void. In contrast, OX2R is highly expressed in the hypothalamic paraventricular nucleus despite null expression of OX1R.

Unlike the expression patterns of prepro-orexin, orexin terminals, and their receptors, less is known regarding the mechanism of action or the intracellular signaling cascades following orexin receptor activation. Regarding the mechanism of action, it is clear that OX1R and OX2R have dissimilar affinity to OXA and OXB such that OX1R has 10-fold greater affinity for OXA relative to OXB, while OX2R appears to have equal affinity for OXA and OXB (Sakurai et al., 1998). Although the significance of the dissimilar binding affinity is not readily apparent, the disparate behavioral effects of orexins on energy expenditure-related processes may uncover this physiological relevance. It is also clear that orexin receptor binding can activate various G-alpha subunit proteins including Gq, Gs, and Gi/o, which increases the complexity of understanding their signaling cascades (Bernard et al., 2006; Holmqvist et al., 2005; Karteris et al., 2001, 2005; Randeva et al., 2001). An early report documented that orexin receptor stimulation increased extracellular calcium influx (Ammoun et al., 2003; van den Pol et al., 1998). Further, neuronal excitation results from activation of L-type calcium channels (Kohlmeier et al., 2004), T-type calcium channels (Zhang et al., 2009), transient receptor potential cation channel subtypes C1 and C3 (Larsson et al., 2005), nonselective cation channels (Yang and Ferguson, 2003), sodium–calcium exchange channel transporter (Burdakov et al., 2003; Eriksson et al., 2001), as well as the suppression of potassium conductance through G-protein inwardly rectifying potassium channels (Hoang et al., 2003; Ivanov and Aston-Jones, 2000). Moreover, orexin stimulates release of glutamate and GABA (van den Pol et al., 1998), and stimulation of orexin receptors by OXA and OXB activates p38 mitogen-activated phosphate kinase and extracellular signal-regulated kinases (ERK1/2) (Ammoun et al., 2006; Tang et al., 2008) by way of multiple signaling mechanisms. Moreover, orexin neurons are energy sensors sensitive to ATP levels (Liu et al., 2011) and lactate (Parsons and Hirasawa, 2010) and have been differentially described based on synaptic input and their behavioral, morphological, and electrophysiological profile (Harris et al., 2005; Horvath and Gao, 2005). A recent review suggests that orexin neurons function within a local network comprising both coexpressing neurotransmitters within orexin neurons (dynorphin, glutamate, nociceptin/orphanin FQ) as well as neurons devoid of orexin in the lateral hypothalamus that express melanin concentrating hormone, thyrotropin-releasing hormone, corticotropin-releasing hormone, neurotensin, galanin, and GABA (Burt et al., 2011). Based on the mirrored distribution between orexin fibers and receptors, the multisynaptic input to the lateral hypothalamus and the local network within orexin neurons, it is not surprising that orexins mediate multiple physiological processes and their role as a physiological integrator could not be made more clear (de Lecea et al., 1998; Sakurai et al., 1998; Sutcliffe and de Lecea, 2000; Willie et al., 2001).

A. Components of energy expenditure

Total energy expenditure can be partitioned into several components (D’Alessio et al., 1988; Donahoo et al., 2004; Joosen and Westerterp, 2006; Ravussin and Bogardus, 1992; Ravussin et al., 1986). However, the relative contribution of each component to total energy expenditure is largely dependent upon the interindividual variability of each component. For example, energy expenditure from physical activity is the most variable (Ravussin et al., 1986). The primary component of total energy expenditure comprising 60–70%, basal metabolism, is defined by the Webster’s Medical Dictionary as “the turnover of energy in a fasting and resting organism using energy solely to maintain vital cellular activity, respiration, and circulation as measured by the basal metabolic rate” (BMR). Diet-induced thermogenesis, comprising 10–15% of total energy expenditure, is due to the energy required to digest, absorb, and store food. Adaptive thermogenesis or the energy required to thermoregulate and respond to changes in the environmental temperature comprises 10–15% of total energy expenditure. The final and most variable component of total energy expenditure, comprising 6–10%, is due to physical activity thermogenesis or the work derived from all forms of physical activity, postural maintenance, and muscular contraction. This includes both physical activity due to exercise and all other types of physical activity excluding exercise (e.g., spontaneous physical activity or nonexercise activity) (Levine et al., 1999; Ravussin et al., 1986).

B. Methodological considerations: Energy expenditure and sleep measurements in rodents in the context of obesity research

Orexin modulates food intake, BMR, physical activity, arousal, the sleep/ wake cycle, and hormone regulation and thus has profound influences on energy expenditure. Therefore, it is important to reflect upon common methods currently used to measure energy expenditure, indirect calorimetry. With this method, rodents are placed in a calorimetry chamber and “reference” air of a known percentage of oxygen and carbon dioxide flows into the chamber. The rodent within the calorimetry chamber consumes oxygen and expires carbon dioxide, which combines with the air flowing into the chamber. Through periodic sampling, energy expenditure is calculated based on the difference between the percentage of oxygen and carbon dioxide between the reference and sampled air. Although indirect calorimetry is used to measure total energy expenditure, the method and analysis of energy expenditure are problematic when comparing energy expenditure in rodents within the context of obesity research (Arch et al., 2006; Butler and Kozak, 2010; Rauh et al., 1990). First, it is currently impossible to partition total energy expenditure into its relative components since an instantaneous change in basal metabolism, physical activity, temperature, or diet-induced thermogenesis cannot be detected in real time. Instead, BMR is commonly considered to be the lowest metabolic rate during the resting phase or derived from a series of metabolic rates with the lowest standard deviation during the resting phase for the animal. Adaptive thermogenesis is considered to be negligible at thermoneutral, which seems germane to consider when comparing energy expenditure between animals of varying body mass, surface area, and volume. Laboratory rodents are commonly housed at temperatures below their thermoneutral (Overton, 2010). Therefore, adaptive thermogenesis and BMR would be greater in rodents with smaller surfaces area to volume ratios. Thus, the derived or calculated BMR may overestimate the true BMR for small animals housed below thermoneutral in an indirect calorimeter, since the calculated BMR will include the energy required for adaptive thermogenesis. Also, while it is possible to measure physical activity and energy expenditure concurrently, physical activity-induced energy expenditure cannot be directly measured. Finally, the metabolic rate during sleep is reported infrequently in indirect calorimetry studies possibly due to the technical difficulties associated with quantifying and measuring sleep directly. Sleep can be classified broadly into rapid eye movement (REM) sleep and non-rapid eye movement (NREM or slow wave sleep (SWS)) sleep and each has distinct electrographic waveforms in electroencephalogram (EEG) and electromyogram (EMG) signals (Greene and Siegel, 2004). Quantification of sleep and wake first requires continuous recordings of EEG and EMG signals and then requires the EEG/EMG waveforms to be scored and classified as REM sleep, SWS, or wake. Pairing methodology to concurrently time stamp and differentiate sleep and wake states by EEG/EMG recordings and energy expenditure with indirect calorimetry would be an improvement over methods that estimate sleeping metabolic rate. Despite that pairing methodologies are imperfect, as mentioned previously, comparing patterns of sleep, wake, and energy expenditure in parallel will lead to a more accurate estimation of sleeping metabolic rate. This has physiological significance for human obesity given the effect of sleep deprivation on energy expenditure and body mass. Therefore, we focus on orexin modulation of total energy expenditure in addition to the components of energy expenditure including sympathetic outflow, thermo-regulation, physical activity, and sleep; yet acknowledge that many neurotransmitter systems and neuropeptides contribute to the regulation of energy expenditure.

C. Total energy expenditure and metabolic rate

The earliest report of orexin modulation of total energy expenditure was reported shortly after their discovery (Lubkin and Stricker-Krongrad, 1998) and confirmed in later studies (Asakawa et al., 2002; Semjonous et al., 2009; Wang et al., 2001). Mice with cannule directed toward the third ventricle were infused with a single dose of OXA in the light or dark phase among food-deprived or ad libitum-fed mice (Lubkin and Stricker-Krongrad, 1998). Interestingly, OXA stimulated metabolic rate and the duration of action was longer after the dark phase injection, suggesting that other components of energy expenditure may have been influenced. It is plausible that energy expenditure due to increased physical activity or sympathetic activity but not diet-induced thermogenesis contributed to the prolonged increase in BMR, since OXA stimulates both physical activity and sympathetic activity and the control-treated mice had greater food intake than the OXA-treated mice. As the site of action for an injectate infused into the brain ventricles is largely unknown, others tested the effects of OXA after infusion into specific brain sites to identify potential sites of action. Wang et al. found OXA stimulated whole body oxygen consumption in anesthetized rats after infusion into the arcuate nucleus only (Wang et al., 2003). There was no increase after infusion into the locus coeruleus, paraventricular nucleus of thalamus or several hypothalamic nuclei including the medial preoptic area, paraventricular, dorsomedial, ventromedial, and the lateral hypothalamic area. In contrast, in awake freely moving rats, OXA increased whole body energy expenditure after infusion into the hypothalamic paraventricular nucleus (Kiwaki et al., 2004; Novak et al., 2006) and the lateral hypothalamus (Teske et al., 2006). As physical activity would be absent in anesthetized rats, it is plausible that the positive effect of OXA on energy expenditure after infusion in the hypothalamic paraventricular nucleus and lateral hypothalamus in the latter studies may be due in part to increased energy expenditure associated with physical activity.

D. Components of total energy expenditure

1. Sympathetic outflow and thermoregulation

Orexin stimulates thermoregulatory and cardiovascular systems (Ferguson and Samson, 2003; Samson et al., 2005; Shirasaka et al., 2002; Szekely et al., 2002), which would contribute to the increase in basal metabolism following central orexin noted previously. Orexin is largely sympathoexcitatory. Studies in rodents document orexin-stimulated elevations in blood pressure and heart rate (Chen et al., 2000; Monda et al., 2001; Samson et al., 1999; Shirasaka et al., 1999; Wang et al., 2001) and sympathetic outflow indicated by increased renal sympathetic nerve activity, plasma epinephrine, nor-adrenaline release, and firing rate of sympathetic nerves (Hirota et al., 2001; Matsumura et al., 2001; Monda et al., 2001, 2003, 2004; Shirasaka et al., 1999). Despite that peripheral orexin receptors have been identified intravenous orexin infusion failed to increase heart rate (Chen et al., 2000), which suggests that central sites of action such as the rostral ventrolateral medulla (Chen et al., 2000), nucleus of the solitary tract (de Oliveira and Ciriello, 2003; Smith et al., 2002), arcuate nucleus (Wang et al., 2003), hypothalamic paraventricular nucleus (Sato-Suzuki et al., 2002), and the diagonal band of Broca (Monda et al., 2004) may be largely responsible for the tachycardic response since OXA readily transverses the blood–brain barrier (Kastin and Akerstrom, 1999). However, interestingly, both heart rate and blood pressure are increased after infusion into the rostral ventrolateral medulla (Chen et al., 2000) and nucleus of the solitary tract (de Oliveira and Ciriello, 2003; Smith et al., 2002). In contrast, OXA has been shown to have negative effects as OXA in the nucleus ambiguous (de Oliveira and Ciriello, 2003) and the subfornical organ reduced heart rate (Smith et al., 2007), and OXA in the nucleus ambiguous (de Oliveira and Ciriello, 2003) had no effect on blood pressure but intra-subfornical organ OXA (Smith et al., 2007) reduced blood pressure. The significance of the brain site of infusion is underscored by comparing the pressor and tachycardic response to OXA across brain sites.

Orexins role in thermoregulation is exemplified by the stimulatory effect of OXA on the interscapular brown adipose tissue (iBAT), a thermosensitive organ rich in uncoupling protein one, which is largely responsible for adaptive thermogenesis in rodents. Despite that iBAT is lacking in adult humans, the relevance of brown adipose tissue to human thermogenesis has been hotly debated and recent studies suggest relevance to human obesity (Sellayah et al., 2011). Increased uncoupling protein one activity favors heat production rather than ATP production by shuttling protons across the inner mitochondrial membrane to dissipate the proton-motive force driving oxidative phosphorylation. Stimulation of sympathetic nerves innervating the iBAT stimulates colonic temperature, and cold exposure stimulates iBAT activity, which directly contrasts the effect of housing rodents above thermoneutral, which reduces iBAT activity. An early study demonstrated that OXA increased firing rate of sympathetic nerves innervating the iBAT (Monda et al., 2001) and iBAT temperature (Monda et al., 2001), both of which would be expected to increase thermogenesis and is consistent with OXA stimulation of colonic temperature (Monda et al., 2001, 2004; Wang et al., 2003; Zheng et al., 2005). Recently, Morrison and colleagues proposed a neuroanatomical basis for OXA modulation of thermoregulation through the rostral raphe pallidus and lateral parapyramidal area, which are sites of BAT sympathetic premotor neurons. Orexin A injection into the rostral raphe pallidus and lateral parapyramidal area robustly increased iBAT sympathetic outflow, expired carbon dioxide and both core and iBAT temperature in anesthetized rats (Tupone et al., 2011). Although this suggests that OXA-modulation of temperature does in fact increase energy expenditure, heart rate was also increased. Contrary to the stimulatory effect of OXA on thermoregulation, OXA failed to increase temperature after chronic infusion (Haynes et al., 1999); however, it is plausible that the structural stability of OXA at body temperature may underlie this discrepancy. Thus, the cardiovascular, sympathetic, and thermoregulatory action of orexins support the stimulatory action of orexins on basal metabolism and total energy expenditure.

2. Physical activity

Physical activity contributes to energy expenditure and the promotion of physical activity by OXA is an irrefutable effect. Irrespective of the central site of microinfusion, OXA stimulates physical activity including locomotion, rearing, grooming, and burrowing behaviors after ventricular (Hagan et al., 1999; Ida et al., 1999) or intraparenchymal infusion. Further, a dose-dependent increase in physical activity is paralleled by elevated energy expenditure (Kiwaki et al., 2004). While the magnitude of OXA-induced physical activity varies with respect to the location of the microinjection (Kotz et al., 2008), the overall stimulatory effect remains consistent. Thus far, OXA infusion into the following brain sites has been shown to reliably stimulate physical activity: lateral hypothalamus, hypothalamic paraventricular nucleus, substantia nigra, tuberomammillary nucleus, dorsal raphe, nucleus accumbens, medial preoptic area, and locus coeruleus (Espana et al., 2001; Kiwaki et al., 2004; Kotz et al., 2002, 2006, 2008; Novak and Levine, 2009; Novak et al., 2006; Teske et al., 2006, 2010; Thorpe and Kotz, 2005). Likewise, OXA infused into the medial preoptic area and medial septum elicits grooming (Espana et al., 2001). Moreover, direct increases in muscle tone or EMG activity in specific muscles, which would be indicative of muscle tone, was observed following OXA microinfusion into the locus coeruleus (Kiyashchenko et al., 2001), the alpha gigantocellular reticular nucleus in the medioventral medullary region (Mileykovskiy et al., 2002), trigeminal motor nucleus, and hypoglossal motor nucleus (Peever et al., 2003). Interestingly, orexin into the pontine inhibitory area (Kiyashchenko et al., 2001) and the ventral gigantocellular reticular nucleus in the medioventral medullary region inhibited hindlimb muscle tone (Mileykovskiy et al., 2002). Despite that inhibition of hindlimb muscle tone would seem to contraindicate the increases in physical activity after OXA, parallel activation and suppression of muscles in opposition are required for movement, which would be essential for normal locomotion.

3. Sleep

The role of orexin in sleep/wake regulation and subsequent influence on energy expenditure is exemplified in the pathological sleep condition narcolepsy due to the loss of orexin-containing neurons or orexin receptors in humans and animal models (Chemelli et al., 1999; Lin et al., 1999; Nishino et al., 2000; Siegel, 1999). The symptomatology of narcolepsy, characterized by the inability to consolidate sleep or wake into long bouts, results in sleep fragmentation and sleep behavioral state instability (Mochizuki et al., 2004; Nishino et al., 2000; Sakurai, 2005; Willie et al., 2003). These symptoms are dampened by exogenous OXA in narcoleptic animals (Mieda et al., 2004), which highlights the importance of orexin in the maintenance and stabilization of sleep behavior.

Electrophysiological and molecular studies further support a role for orexin in the maintenance and stability of normal sleep and wakefulness. Orexin neurons display phasic firing patterns during wakefulness yet are silent during SWS (Lee et al., 2005; Mileykovskiy et al., 2005). In parallel, orexin levels display circadian rhythmicity and gradually increase in anticipation of the active phase and decrease prior to and during the rest phase (Fujiki et al., 2001; Lee et al., 2005; Yoshida et al., 2001). Not surprisingly, orexin levels also increase following sleep deprivation (Allard et al., 2007; Pedrazzoli et al., 2004; Wu et al., 2002). Orexin neuron activity during wakefulness is also modulated by the ascending arousal network including the sleep/wake regulating monoamines and acetylcholine (Li et al., 2002; Saper et al., 2010; Yamanaka et al., 2003) such that aminergic and cholinergic arousal circuits sustain orexin cells during wake (Estabrooke et al., 2001; Takahashi et al., 2008). These circuits in turn send inhibitory input to the ventrolateral preoptic area sleep-active neurons of the hypothalamus, and thereby further maintain wakefulness (Gallopin et al., 2000). Interestingly, orexin neurons also promote wake by actively hindering REM sleep by inhibiting the oral pontine reticular nucleus, a structure involved in the generation of REM sleep (Nunez et al., 2006). Thus, the absence of orexin results in reduced activity throughout the arousal network and culminates in an inappropriately low threshold for the transition into sleep from wake, which results in narcolepsy characterized by behavioral instability.

Orexins essential role in maintaining and stabilizing wake is consistent with previously mentioned actions of orexins, including the stimulation of total energy expenditure by promoting elevations in physical activity, body temperature, and sympathetic outflow. One of the earliest investigations on sleep found that between SWS and REM sleep, oxygen consumption was lowest during SWS and highest during REM sleep (Brebbia and Altshuler, 1965). Interestingly, studies in narcoleptic animals and humans suggest that narcolepsy-associated weight gain may be primarily due to reduced availability of orexin and behavioral instability, rather than hypersomnia (excessive active period sleep), as patients with idiopathic hypersomnia had a lower body weight (Kok et al., 2003). Like humans, mice lacking orexin weighed more than wild-type mice, yet they had reduced energy intake and physical activity (Hara et al., 2001). Thus, reduced metabolic rate and energy expenditure resulting from the absence of long wake bouts likely contributes to the increased body weight in orexin-deficient humans and animals. Furthermore, behavioral studies show that orexin antagonism increases sleep fragmentation (Beuckmann et al., 2004; Brisbare-Roch et al., 2007; Chen et al., 2006; Gerashchenko et al., 2001, 2003; Llewellyn-Smith et al., 2003; Thakkar et al., 1999) while central administration of OXA increases arousal and wakefulness and reduces both NREM and REM sleep (Bourgin et al., 2000; Espana et al., 2001; Hagan et al., 1999; John et al., 2000; Methippara et al., 2000; Piper et al., 2000; Rodgers et al., 2002; Xi et al., 2001), which is blunted by preinfusion with the OX1R antagonist SB 334867 (Smith et al., 2003). In a parallel manner, optogenetic silencing of orexin neurons induces SWS (Tsunematsu et al., 2011), while optogenetic activation of orexin neurons prompts mice to wake from sleep (Adamantidis et al., 2007; Carter et al., 2009). Thus, several studies highlight the importance of orexin in sleep/wake and observations from persons with narcolepsy and animals models of narcolepsy underscore its physiological significance.

Apart from the stimulating energy expenditure by promoting wake and activity, orexin modulates body temperature during sleep (Mochizuki et al., 2006). Orexin may have indirect effects on sleep and thermoregulation as OXA-induced physical activity during the day may increase body temperature and the accumulation of metabolic end products that promote sleep and heat loss in a circadian manner (Alam et al., 1996; McGinty and Szymusiak, 2001). At the neural level, orexin neurons reciprocally innervate the preoptic area, which promotes sleep and heat loss (Alam et al., 1995; Kumar, 2004) and thermal stimuli to the preoptic area strongly modulates sleep propensity and EEG delta activity (Kumar, 2004). Orexin terminals project to critical thermoregulatory brain sites including the tuberomammillary neurons, locus coeruleus, basal forebrain, pontine inhibitory area dorsomedial hypothalamus, ventromedial hypothalamus, posterior hypothalamus, periaqueductal gray, dorsal raphe nucleus, and intermediolateral column of the spinal cord (Kornum et al., 2011; Mochizuki et al., 2006). Thus, orexins are neuroanatomically positioned to modulate thermoregulation and contribute to heat loss during sleep (Ohno and Sakurai, 2008).

Overall heat loss, due to reduced production and increased loss, is an essential aspect of sleep (Mochizuki et al., 2006), which underscores the energy conservation function proposed for sleep. Circadian modulation of body temperature may also alter sleep propensity (Kumar, 2004), which would be expected to have energetic consequences. In parallel, a moderate fall in body temperature may be necessary for good-quality sleep (Gilbert et al., 2004). When compared to persons with normal sleep quality, the normal decline in body temperature is blunted in individuals with poor sleep quality (Lushington et al., 2000; Pierangeli et al., 2001; Zepelin and McDonald, 1987). Thus, it is feasible that altered thermoregulation also contributes to sleep fragmentation in narcolepsy (Mochizuki et al., 2006). Indeed, Mochizuki et al. showed parallel 24 h rhythms of physical activity, wake and body temperature in wild-type mice (Mochizuki et al., 2006). This demonstrates an association between vigilance states and energy expenditure, which was less prominent in orexin knockout mice. Importantly, while both wild-type and orexin knockout mice exhibited declines in body temperature during sleep, temperature declined to a lesser extent during sleep in orexin knockout mice. Likewise, elevations in body temperature during active wake were minor in orexin knockout mice compared the elevations in wild-type mice. Hence, the blunted decline in body temperature may be due to the intermittent awakenings during sleep (Mochizuki et al., 2006) while the dampened increase in body temperature may be due to shorter wake bouts and physical activity during wake in the orexin knockout mice (Mochizuki et al., 2004). In sum, these data reiterate a role for orexin in heat loss during sleep, possibly through sleep stabilization (Mochizuki et al., 2006).

It is plausible that orexin effects on sleep/wake are also modulated by the autonomic nervous system. During wake, orexin inhibits the heat loss-inducing parasympathetic nervous system (Dergacheva et al., 2005) and promotes sympathetic nervous system activity. In contrast, during sleep, orexin neuron activity and orexin levels are low while parasympathetic activity predominates (Kuo et al., 2008). As noted above, exogenous orexin activates the sympathetic nervous system and orexin injection into the raphe pallidus increases sympathetic outflow to the brown adipose tissue, which increases thermogenesis (Morrison and Nakamura, 2011; Tupone et al., 2011). Conversely, mice lacking orexin have impaired brown adipose tissue thermogenesis (Sellayah et al., 2011; Vijgen et al., 2011). Thus, these data imply that orexin potently enhances energy expenditure, via stimulation of sympathetic activation and suppression of parasympathetic activity.

II. Conclusion

Orexins provide a crucial link between energy balance and arousal. The orexin neurons functionally interact with neural pathways for physical activity pathways and sleep and cardiovascular and temperature regulatory centers in the brain, which positions orexin neurons to coordinate energy expenditure and the sleep–wake cycle. Corollary to this, obesity in humans and animal models is associated with altered sleep, reduced physical activity, and plasma and hypothalamic levels of orexin, whereas weight loss is associated with increased orexin levels in the plasma and improved sleep quality. These studies suggest that reduced component-specific energy expenditure due to lack of physical activity may culminate in weight gain in narcoleptic and obese humans as well as in orexin-deficient animals. Thus, the orexin system may act as a sensor for metabolism and send signals to intra- and extrahypothalamic targets to modify behavioral outputs based on energy needs.

References

  1. Adamantidis AR, Zhang F, Aravanis AM, Deisseroth K, de Lecea L. Neural substrates of awakening probed with optogenetic control of hypocretin neurons. Nature. 2007;450:420–424. doi: 10.1038/nature06310. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Alam MN, McGinty D, Szymusiak R. Neuronal discharge of preoptic/ anterior hypothalamic thermosensitive neurons: Relation to NREM sleep. Am J Physiol. 1995;269:R1240–R1249. doi: 10.1152/ajpregu.1995.269.5.R1240. [DOI] [PubMed] [Google Scholar]
  3. Alam MN, McGinty D, Szymusiak R. Preoptic/anterior hypothalamic neurons: Thermosensitivity in wakefulness and non rapid eye movement sleep. Brain Res. 1996;718:76–82. doi: 10.1016/0006-8993(96)00035-2. [DOI] [PubMed] [Google Scholar]
  4. Allard JS, Tizabi Y, Shaffery JP, Manaye K. Effects of rapid eye movement sleep deprivation on hypocretin neurons in the hypothalamus of a rat model of depression. Neuropeptides. 2007;41:329–337. doi: 10.1016/j.npep.2007.04.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Ammoun S, Holmqvist T, Shariatmadari R, Oonk HB, Detheux M, Parmentier M, Akerman KE, Kukkonen JP. Distinct recognition of OX1 and OX2 receptors by orexin peptides. J Pharmacol Exp Ther. 2003;305:507–514. doi: 10.1124/jpet.102.048025. [DOI] [PubMed] [Google Scholar]
  6. Ammoun S, Lindholm D, Wootz H, Akerman KE, Kukkonen JP. G-protein-coupled OX1 orexin/hcrtr-1 hypocretin receptors induce caspase-dependent and -independent cell death through p38 mitogen-/stress-activated protein kinase. J Biol Chem. 2006;281:834–842. doi: 10.1074/jbc.M508603200. [DOI] [PubMed] [Google Scholar]
  7. Arch JR, Hislop D, Wang SJ, Speakman JR. Some mathematical and technical issues in the measurement and interpretation of open-circuit indirect calorimetry in small animals. Int J Obes (Lond) 2006;30:1322–1331. doi: 10.1038/sj.ijo.0803280. [DOI] [PubMed] [Google Scholar]
  8. Asakawa A, Inui A, Goto K, Yuzuriha H, Takimoto Y, Inui T, Katsuura G, Fujino MA, Meguid MM, Kasuga M. Effects of agouti-related protein, orexin and melanin-concentrating hormone on oxygen consumption in mice. Int J Mol Med. 2002;10:523–525. [PubMed] [Google Scholar]
  9. Bernard R, Lydic R, Baghdoyan HA. Hypocretin (orexin) receptor subtypes differentially enhance acetylcholine release and activate g protein subtypes in rat pontine reticular formation. J Pharmacol Exp Ther. 2006;317:163–171. doi: 10.1124/jpet.105.097071. [DOI] [PubMed] [Google Scholar]
  10. Beuckmann CT, Sinton CM, Williams SC, Richardson JA, Hammer RE, Sakurai T, Yanagisawa M. Expression of a poly-glutamine-ataxin-3 transgene in orexin neurons induces narcolepsy-cataplexy in the rat. J Neurosci. 2004;24:4469–4477. doi: 10.1523/JNEUROSCI.5560-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Bourgin P, Huitron-Resendiz S, Spier AD, Fabre V, Morte B, Criado JR, Sutcliffe JG, Henriksen SJ, de Lecea L. Hypocretin-1 modulates rapid eye movement sleep through activation of locus coeruleus neurons. J Neurosci. 2000;20:7760–7765. doi: 10.1523/JNEUROSCI.20-20-07760.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Brebbia DR, Altshuler KZ. Oxygen consumption rate and electroencephalographic stage of sleep. Science. 1965;150:1621–1623. doi: 10.1126/science.150.3703.1621. [DOI] [PubMed] [Google Scholar]
  13. Brisbare-Roch C, Dingemanse J, Koberstein R, Hoever P, Aissaoui H, Flores S, Mueller C, Nayler O, van Gerven J, de Haas SL, Hess P, Qiu C, et al. Promotion of sleep by targeting the orexin system in rats, dogs and humans. Nat Med. 2007;13:150–155. doi: 10.1038/nm1544. [DOI] [PubMed] [Google Scholar]
  14. Burdakov D, Liss B, Ashcroft FM. Orexin excites GABAergic neurons of the arcuate nucleus by activating the sodium–calcium exchanger. J Neurosci. 2003;23:4951–4957. doi: 10.1523/JNEUROSCI.23-12-04951.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. 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. 2011;301:R572–R580. doi: 10.1152/ajpregu.00674.2010. [DOI] [PubMed] [Google Scholar]
  16. Butler AA, Kozak LP. A recurring problem with the analysis of energy expenditure in genetic models expressing lean and obese phenotypes. Diabetes. 2010;59:323–329. doi: 10.2337/db09-1471. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Carter ME, Adamantidis A, Ohtsu H, Deisseroth K, de Lecea L. Sleep homeostasis modulates hypocretin-mediated sleep-to-wake transitions. J Neurosci. 2009;29:10939–10949. doi: 10.1523/JNEUROSCI.1205-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Chemelli RM, Willie JT, Sinton CM, Elmquist JK, Scammell T, Lee C, Richardson JA, Williams SC, Xiong Y, Kisanuki Y, Fitch TE, Nakazato M, et al. Narcolepsy in orexin knockout mice: Molecular genetics of sleep regulation. Cell. 1999;98:437–451. doi: 10.1016/s0092-8674(00)81973-x. [DOI] [PubMed] [Google Scholar]
  19. Chen CT, Dun SL, Kwok EH, Dun NJ, Chang JK. Orexin A-like immunoreactivity in the rat brain. Neurosci Lett. 1999;260:161–164. doi: 10.1016/s0304-3940(98)00977-x. [DOI] [PubMed] [Google Scholar]
  20. Chen CT, Hwang LL, Chang JK, Dun NJ. Pressor effects of orexins injected intracisternally and to rostral ventrolateral medulla of anesthetized rats. Am J Physiol Regul Integr Comp Physiol. 2000;278:R692–R697. doi: 10.1152/ajpregu.2000.278.3.R692. [DOI] [PubMed] [Google Scholar]
  21. Chen L, Thakkar MM, Winston S, Bolortuya Y, Basheer R, McCarley RW. REM sleep changes in rats induced by siRNA-mediated orexin knockdown. Eur J Neurosci. 2006;24:2039–2048. doi: 10.1111/j.1460-9568.2006.05058.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Cluderay JE, Harrison DC, Hervieu GJ. Protein distribution of the orexin-2 receptor in the rat central nervous system. Regul Pept. 2002;104:131–144. doi: 10.1016/s0167-0115(01)00357-3. [DOI] [PubMed] [Google Scholar]
  23. Cutler DJ, Morris R, Sheridhar V, Wattam TA, Holmes S, Patel S, Arch JR, Wilson S, Buckingham RE, Evans ML, Leslie RA, Williams G. Differential distribution of orexin-A and orexin-B immunoreactivity in the rat brain and spinal cord. Peptides. 1999;20:1455–1470. doi: 10.1016/s0196-9781(99)00157-6. [DOI] [PubMed] [Google Scholar]
  24. D’Alessio DA, Kavle EC, Mozzoli MA, Smalley KJ, Polansky M, Kendrick ZV, Owen LR, Bushman MC, Boden G, Owen OE. Thermic effect of food in lean and obese men. J Clin Invest. 1988;81:1781–1789. doi: 10.1172/JCI113520. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Date Y, Mondal MS, Matsukura S, Nakazato M. Distribution of orexin-A and orexin-B (hypocretins) in the rat spinal cord. Neurosci Lett. 2000;288:87–90. doi: 10.1016/s0304-3940(00)01195-2. [DOI] [PubMed] [Google Scholar]
  26. de Lecea L, Kilduff TS, Peyron C, Gao X, Foye PE, Danielson PE, Fukuhara C, Battenberg EL, Gautvik VT, Bartlett FS, 2nd, Frankel WN, van den Pol AN, et al. The hypocretins: Hypothalamus-specific peptides with neuroexcitatory activity. Proc Natl Acad Sci U S A. 1998;95:322–327. doi: 10.1073/pnas.95.1.322. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. de Oliveira CV, Ciriello J. Cardiovascular responses to hypocretin-1 in nucleus ambiguus of the ovariectomized female rat. Brain Res. 2003;986:148–156. doi: 10.1016/s0006-8993(03)03226-8. [DOI] [PubMed] [Google Scholar]
  28. Dergacheva O, Wang X, Huang ZG, Bouairi E, Stephens C, Gorini C, Mendelowitz D. Hypocretin-1 (orexin-A) facilitates inhibitory and diminishes excitatory synaptic pathways to cardiac vagal neurons in the nucleus ambiguus. J Pharmacol Exp Ther. 2005;314:1322–1327. doi: 10.1124/jpet.105.086421. [DOI] [PubMed] [Google Scholar]
  29. Digby JE, Chen J, Tang JY, Lehnert H, Matthews RN, Randeva HS. Orexin receptor expression in human adipose tissue: Effects of orexin-A and orexin-B. J Endocrinol. 2006;191:129–136. doi: 10.1677/joe.1.06886. [DOI] [PubMed] [Google Scholar]
  30. Donahoo WT, Levine JA, Melanson EL. Variability in energy expenditure and its components. Curr Opin Clin Nutr Metab Care. 2004;7:599–605. doi: 10.1097/00075197-200411000-00003. [DOI] [PubMed] [Google Scholar]
  31. Eriksson KS, Sergeeva O, Brown RE, Haas HL. Orexin/hypocretin excites the histaminergic neurons of the tuberomammillary nucleus. J Neurosci. 2001;21:9273–9279. doi: 10.1523/JNEUROSCI.21-23-09273.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Espana RA, Baldo BA, Kelley AE, Berridge CW. Wake-promoting and sleep-suppressing actions of hypocretin (orexin): Basal forebrain sites of action. Neuroscience. 2001;106:699–715. doi: 10.1016/s0306-4522(01)00319-0. [DOI] [PubMed] [Google Scholar]
  33. Estabrooke IV, McCarthy MT, Ko E, Chou TC, Chemelli RM, Yanagisawa M, Saper CB, Scammell TE. Fos expression in orexin neurons varies with behavioral state. J Neurosci. 2001;21:1656–1662. doi: 10.1523/JNEUROSCI.21-05-01656.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Ferguson AV, Samson WK. The orexin/hypocretin system: A critical regulator of neuroendocrine and autonomic function. Front Neuroendocrinol. 2003;24:141–150. doi: 10.1016/s0091-3022(03)00028-1. [DOI] [PubMed] [Google Scholar]
  35. Fujiki N, Yoshida Y, Ripley B, Honda K, Mignot E, Nishino S. Changes in CSF hypocretin-1 (orexin A) levels in rats across 24 hours and in response to food deprivation. Neuroreport. 2001;12:993–997. doi: 10.1097/00001756-200104170-00026. [DOI] [PubMed] [Google Scholar]
  36. Gallopin T, Fort P, Eggermann E, Cauli B, Luppi PH, Rossier J, Audinat E, Muhlethaler M, Serafin M. Identification of sleep-promoting neurons in vitro. Nature. 2000;404:992–995. doi: 10.1038/35010109. [DOI] [PubMed] [Google Scholar]
  37. Gautvik KM, de Lecea L, Gautvik VT, Danielson PE, Tranque P, Dopazo A, Bloom FE, Sutcliffe JG. Overview of the most prevalent hypothalamus-specific mRNAs, as identified by directional tag PCR subtraction. Proc Natl Acad Sci U S A. 1996;93:8733–8738. doi: 10.1073/pnas.93.16.8733. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Gerashchenko D, Kohls MD, Greco M, Waleh NS, Salin-Pascual R, Kilduff TS, Lappi DA, Shiromani PJ. Hypocretin-2-saporin lesions of the lateral hypothalamus produce narcoleptic-like sleep behavior in the rat. J Neurosci. 2001;21:7273–7283. doi: 10.1523/JNEUROSCI.21-18-07273.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Gerashchenko D, Blanco-Centurion C, Greco MA, Shiromani PJ. Effects of lateral hypothalamic lesion with the neurotoxin hypocretin-2-saporin on sleep in Long-Evans rats. Neuroscience. 2003;116:223–235. doi: 10.1016/s0306-4522(02)00575-4. [DOI] [PubMed] [Google Scholar]
  40. Gilbert SS, van den Heuvel CJ, Ferguson SA, Dawson D. Thermoregulation as a sleep signalling system. Sleep Med Rev. 2004;8:81–93. doi: 10.1016/S1087-0792(03)00023-6. [DOI] [PubMed] [Google Scholar]
  41. Greene R, Siegel J. Sleep: A functional enigma. Neuromolecular Med. 2004;5:59–68. doi: 10.1385/NMM:5:1:059. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Hagan JJ, Leslie RA, Patel S, Evans ML, Wattam TA, Holmes S, Benham CD, Taylor SG, Routledge C, Hemmati P, Munton RP, Ashmeade TE, et al. Orexin A activates locus coeruleus cell firing and increases arousal in the rat. Proc Natl Acad Sci U S A. 1999;96:10911–10916. doi: 10.1073/pnas.96.19.10911. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. 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. 2001;30:345–354. doi: 10.1016/s0896-6273(01)00293-8. [DOI] [PubMed] [Google Scholar]
  44. Harris GC, Wimmer M, Aston-Jones G. A role for lateral hypothalamic orexin neurons in reward seeking. Nature. 2005;437:556–559. doi: 10.1038/nature04071. [DOI] [PubMed] [Google Scholar]
  45. Haynes AC, Jackson B, Overend P, Buckingham RE, Wilson S, Tadayyon M, Arch JR. Effects of single and chronic intracerebroventricular administration of the orexins on feeding in the rat. Peptides. 1999;20:1099–1105. doi: 10.1016/s0196-9781(99)00105-9. [DOI] [PubMed] [Google Scholar]
  46. Hervieu GJ, Cluderay JE, Harrison DC, Roberts JC, Leslie RA. Gene expression and protein distribution of the orexin-1 receptor in the rat brain and spinal cord. Neuroscience. 2001;103:777–797. doi: 10.1016/s0306-4522(01)00033-1. [DOI] [PubMed] [Google Scholar]
  47. Hirota K, Kushikata T, Kudo M, Kudo T, Lambert DG, Matsuki A. Orexin A and B evoke noradrenaline release from rat cerebrocortical slices. Br J Pharmacol. 2001;134:1461–1466. doi: 10.1038/sj.bjp.0704409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Hoang QV, Bajic D, Yanagisawa M, Nakajima S, Nakajima Y. Effects of orexin (hypocretin) on GIRK channels. J Neurophysiol. 2003;90:693–702. doi: 10.1152/jn.00001.2003. [DOI] [PubMed] [Google Scholar]
  49. Holmqvist T, Johansson L, Ostman M, Ammoun S, Akerman KE, Kukkonen JP. OX1 orexin receptors couple to adenylyl cyclase regulation via multiple mechanisms. J Biol Chem. 2005;280:6570–6579. doi: 10.1074/jbc.M407397200. [DOI] [PubMed] [Google Scholar]
  50. Horvath TL, Gao XB. Input organization and plasticity of hypocretin neurons: Possible clues to obesity’s association with insomnia. Cell Metab. 2005;1:279–286. doi: 10.1016/j.cmet.2005.03.003. [DOI] [PubMed] [Google Scholar]
  51. Ida T, Nakahara K, Katayama T, Murakami N, Nakazato M. Effect of lateral cerebroventricular injection of the appetite-stimulating neuropeptide, orexin and neuropeptide Y, on the various behavioral activities of rats. Brain Res. 1999;821:526–529. doi: 10.1016/s0006-8993(99)01131-2. [DOI] [PubMed] [Google Scholar]
  52. Ivanov A, Aston-Jones G. Hypocretin/orexin depolarizes and decreases potassium conductance in locus coeruleus neurons. Neuroreport. 2000;11:1755–1758. doi: 10.1097/00001756-200006050-00031. [DOI] [PubMed] [Google Scholar]
  53. John J, Wu MF, Siegel JM. Systemic administration of hypocretin-1 reduces cataplexy and normalizes sleep and waking durations in narcoleptic dogs. Sleep Res Online. 2000;3:23–28. [PMC free article] [PubMed] [Google Scholar]
  54. Johren O, Neidert SJ, Kummer M, Dendorfer A, Dominiak P. Prepro-orexin and orexin receptor mRNAs are differentially expressed in peripheral tissues of male and female rats. Endocrinology. 2001;142:3324–3331. doi: 10.1210/endo.142.8.8299. [DOI] [PubMed] [Google Scholar]
  55. Joosen AM, Westerterp KR. Energy expenditure during overfeeding. Nutr Metab. 2006;3:25. doi: 10.1186/1743-7075-3-25. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Karteris E, Randeva HS, Grammatopoulos DK, Jaffe RB, Hillhouse EW. Expression and coupling characteristics of the CRH and orexin type 2 receptors in human fetal adrenals. J Clin Endocrinol Metab. 2001;86:4512–4519. doi: 10.1210/jcem.86.9.7849. [DOI] [PubMed] [Google Scholar]
  57. Karteris E, Machado RJ, Chen J, Zervou S, Hillhouse EW, Randeva HS. Food deprivation differentially modulates orexin receptor expression and signaling in rat hypothalamus and adrenal cortex. Am J Physiol Endocrinol Metab. 2005;288:E1089–E1100. doi: 10.1152/ajpendo.00351.2004. [DOI] [PubMed] [Google Scholar]
  58. Kastin AJ, Akerstrom V. Orexin A but not orexin B rapidly enters brain from blood by simple diffusion. J Pharmacol Exp Ther. 1999;289:219–223. [PubMed] [Google Scholar]
  59. Kiwaki K, Kotz CM, Wang C, Lanningham-Foster L, Levine JA. Orexin A (hypocretin 1) injected into hypothalamic paraventricular nucleus and spontaneous physical activity in rats. Am J Physiol Endocrinol Metab. 2004;286:E551–E559. doi: 10.1152/ajpendo.00126.2003. [DOI] [PubMed] [Google Scholar]
  60. Kiyashchenko LI, Mileykovskiy BY, Lai YY, Siegel JM. Increased and decreased muscle tone with orexin (hypocretin) microinjections in the locus coeruleus and pontine inhibitory area. J Neurophysiol. 2001;85:2008–2016. doi: 10.1152/jn.2001.85.5.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Kohlmeier KA, Inoue T, Leonard CS. Hypocretin/orexin peptide signaling in the ascending arousal system: Elevation of intracellular calcium in the mouse dorsal raphe and laterodorsal tegmentum. J Neurophysiol. 2004;92:221–235. doi: 10.1152/jn.00076.2004. [DOI] [PubMed] [Google Scholar]
  62. Kok SW, Overeem S, Visscher TL, Lammers GJ, Seidell JC, Pijl H, Meinders AE. Hypocretin deficiency in narcoleptic humans is associated with abdominal obesity. Obes Res. 2003;11:1147–1154. doi: 10.1038/oby.2003.156. [DOI] [PubMed] [Google Scholar]
  63. Kornum BR, Faraco J, Mignot E. Narcolepsy with hypocretin/orexin deficiency, infections and autoimmunity of the brain. Curr Opin Neurobiol. 2011;21:897–903. doi: 10.1016/j.conb.2011.09.003. [DOI] [PubMed] [Google Scholar]
  64. Kotz CM, Teske JA, Levine JA, Wang C. Feeding and activity induced by orexin A in the lateral hypothalamus in rats. Regul Pept. 2002;104:27–32. doi: 10.1016/s0167-0115(01)00346-9. [DOI] [PubMed] [Google Scholar]
  65. 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. 2006;142:29–36. doi: 10.1016/j.neuroscience.2006.05.028. [DOI] [PubMed] [Google Scholar]
  66. Kotz CM, Teske JA, Billington CJ. Neuroregulation of nonexercise activity thermogenesis and obesity resistance. Am J Physiol Regul Integr Comp Physiol. 2008;294:R699–R710. doi: 10.1152/ajpregu.00095.2007. [DOI] [PubMed] [Google Scholar]
  67. Kumar VM. Why the medial preoptic area is important for sleep regulation. Indian J Physiol Pharmacol. 2004;48:137–149. [PubMed] [Google Scholar]
  68. Kuo TB, Shaw FZ, Lai CJ, Yang CC. Asymmetry in sympathetic and vagal activities during sleep-wake transitions. Sleep. 2008;31:311–320. doi: 10.1093/sleep/31.3.311. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Larsson KP, Peltonen HM, Bart G, Louhivuori LM, Penttonen A, Antikainen M, Kukkonen JP, Akerman KE. Orexin-A-induced Ca2+ entry: Evidence for involvement of trpc channels and protein kinase C regulation. J Biol Chem. 2005;280:1771–1781. doi: 10.1074/jbc.M406073200. [DOI] [PubMed] [Google Scholar]
  70. Lee MG, Hassani OK, Jones BE. Discharge of identified orexin/ hypocretin neurons across the sleep-waking cycle. J Neurosci. 2005;25:6716–6720. doi: 10.1523/JNEUROSCI.1887-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Levine JA, Eberhardt NL, Jensen MD. Role of nonexercise activity thermogenesis in resistance to fat gain in humans. Science. 1999;283:212–214. doi: 10.1126/science.283.5399.212. [DOI] [PubMed] [Google Scholar]
  72. 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. 2002;36:1169–1181. doi: 10.1016/s0896-6273(02)01132-7. [DOI] [PubMed] [Google Scholar]
  73. Lin L, Faraco J, Li R, Kadotani H, Rogers W, Lin X, Qiu X, de Jong PJ, Nishino S, Mignot E. The sleep disorder canine narcolepsy is caused by a mutation in the hypocretin (orexin) receptor 2 gene. Cell. 1999;98:365–376. doi: 10.1016/s0092-8674(00)81965-0. [DOI] [PubMed] [Google Scholar]
  74. Liu ZW, Gan G, Suyama S, Gao XB. Intracellular energy status regulates activity in hypocretin/orexin neurones: A link between energy and behavioural states. J Physiol. 2011;589:4157–4166. doi: 10.1113/jphysiol.2011.212514. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Llewellyn-Smith IJ, Martin CL, Marcus JN, Yanagisawa M, Minson JB, Scammell TE. Orexin-immunoreactive inputs to rat sympathetic preganglionic neurons. Neurosci Lett. 2003;351:115–119. doi: 10.1016/s0304-3940(03)00770-5. [DOI] [PubMed] [Google Scholar]
  76. Lubkin M, Stricker-Krongrad A. Independent feeding and metabolic actions of orexins in mice. Biochem Biophys Res Commun. 1998;253:241–245. doi: 10.1006/bbrc.1998.9750. [DOI] [PubMed] [Google Scholar]
  77. Lushington K, Dawson D, Lack L. Core body temperature is elevated during constant wakefulness in elderly poor sleepers. Sleep. 2000;23:504–510. [PubMed] [Google Scholar]
  78. Marcus JN, Aschkenasi CJ, Lee CE, Chemelli RM, Saper CB, Yanagisawa M, Elmquist JK. Differential expression of orexin receptors 1 and 2 in the rat brain. J Comp Neurol. 2001;435:6–25. doi: 10.1002/cne.1190. [DOI] [PubMed] [Google Scholar]
  79. Matsumura K, Tsuchihashi T, Abe I. Central orexin-A augments sympathoadrenal outflow in conscious rabbits. Hypertension. 2001;37:1382–1387. doi: 10.1161/01.hyp.37.6.1382. [DOI] [PubMed] [Google Scholar]
  80. McGinty D, Szymusiak R. Brain structures and mechanisms involved in the generation of NREM sleep: Focus on the preoptic hypothalamus. Sleep Med Rev. 2001;5:323–342. doi: 10.1053/smrv.2001.0170. [DOI] [PubMed] [Google Scholar]
  81. Methippara MM, Alam MN, Szymusiak R, McGinty D. Effects of lateral preoptic area application of orexin-A on sleep-wakefulness. Neuroreport. 2000;11:3423–3426. doi: 10.1097/00001756-200011090-00004. [DOI] [PubMed] [Google Scholar]
  82. Mieda M, Willie JT, Hara J, Sinton CM, Sakurai T, Yanagisawa M. Orexin peptides prevent cataplexy and improve wakefulness in an orexin neuron-ablated model of narcolepsy in mice. Proc Natl Acad Sci U S A. 2004;101:4649–4654. doi: 10.1073/pnas.0400590101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Mileykovskiy BY, Kiyashchenko LI, Siegel JM. Muscle tone facilitation and inhibition after orexin-a (hypocretin-1) microinjections into the medial medulla. J Neurophysiol. 2002;87:2480–2489. doi: 10.1152/jn.2002.87.5.2480. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Mileykovskiy BY, Kiyashchenko LI, Siegel JM. Behavioral correlates of activity in identified hypocretin/orexin neurons. Neuron. 2005;46:787–798. doi: 10.1016/j.neuron.2005.04.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Mochizuki T, Crocker A, McCormack S, Yanagisawa M, Sakurai T, Scammell TE. Behavioral state instability in orexin knock-out mice. J Neurosci. 2004;24:6291–6300. doi: 10.1523/JNEUROSCI.0586-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Mochizuki T, Klerman EB, Sakurai T, Scammell TE. Elevated body temperature during sleep in orexin knockout mice. Am J Physiol Regul Integr Comp Physiol. 2006;291:R533–R540. doi: 10.1152/ajpregu.00887.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Monda M, Viggiano A, Mondola P, De Luca V. Inhibition of prostaglandin synthesis reduces hyperthermic reactions induced by hypocretin-1/orexin A. Brain Res. 2001;909:68–74. doi: 10.1016/s0006-8993(01)02606-3. [DOI] [PubMed] [Google Scholar]
  88. Monda M, Viggiano A, De Luca V. Paradoxical [correction of parodoxical] effect of orexin A: Hypophagia induced by hyperthermia. Brain Res. 2003;961:220–228. doi: 10.1016/s0006-8993(02)03953-7. [DOI] [PubMed] [Google Scholar]
  89. Monda M, Viggiano A, Viggiano A, Fuccio F, De Luca V. Injection of orexin A into the diagonal band of Broca induces sympathetic and hyperthermic reactions. Brain Res. 2004;1018:265–271. doi: 10.1016/j.brainres.2004.05.084. [DOI] [PubMed] [Google Scholar]
  90. Morrison SF, Nakamura K. Central neural pathways for thermoregulation. Front Biosci. 2011;16:74–104. doi: 10.2741/3677. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Nambu T, Sakurai T, Mizukami K, Hosoya Y, Yanagisawa M, Goto K. Distribution of orexin neurons in the adult rat brain. Brain Res. 1999;827:243–260. doi: 10.1016/s0006-8993(99)01336-0. [DOI] [PubMed] [Google Scholar]
  92. Nishino S, Ripley B, Overeem S, Lammers GJ, Mignot E. Hypocretin (orexin) deficiency in human narcolepsy. Lancet. 2000;355:39–40. doi: 10.1016/S0140-6736(99)05582-8. [DOI] [PubMed] [Google Scholar]
  93. Novak CM, Levine JA. Daily intraparaventricular orexin-A treatment induces weight loss in rats. Obesity (Silver Spring) 2009;17:1493–1498. doi: 10.1038/oby.2009.91. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Novak CM, Kotz CM, Levine JA. Central orexin sensitivity, physical activity, and obesity in diet-induced obese and diet-resistant rats. Am J Physiol Endo-crinol Metab. 2006;290:E396–E403. doi: 10.1152/ajpendo.00293.2005. [DOI] [PubMed] [Google Scholar]
  95. Nunez A, Moreno-Balandran ME, Rodrigo-Angulo ML, Garzon M, De Andres I. Relationship between the perifornical hypothalamic area and oral pontine reticular nucleus in the rat. Possible implication of the hypocretinergic projection in the control of rapid eye movement sleep. Eur J Neurosci. 2006;24:2834–2842. doi: 10.1111/j.1460-9568.2006.05159.x. [DOI] [PubMed] [Google Scholar]
  96. Ohno K, Sakurai T. Orexin neuronal circuitry: Role in the regulation of sleep and wakefulness. Front Neuroendocrinol. 2008;29:70–87. doi: 10.1016/j.yfrne.2007.08.001. [DOI] [PubMed] [Google Scholar]
  97. Overton JM. Phenotyping small animals as models for the human metabolic syndrome: Thermoneutrality matters. Int J Obes (Lond) 2010;34(Suppl 2):S53–S58. doi: 10.1038/ijo.2010.240. [DOI] [PubMed] [Google Scholar]
  98. Parsons MP, Hirasawa M. ATP-sensitive potassium channel-mediated lactate effect on orexin neurons: Implications for brain energetics during arousal. J Neurosci. 2010;30:8061–8070. doi: 10.1523/JNEUROSCI.5741-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Pedrazzoli M, D’Almeida V, Martins PJ, Machado RB, Ling L, Nishino S, Tufik S, Mignot E. Increased hypocretin-1 levels in cerebrospinal fluid after REM sleep deprivation. Brain Res. 2004;995:1–6. doi: 10.1016/j.brainres.2003.09.032. [DOI] [PubMed] [Google Scholar]
  100. Peever JH, Lai YY, Siegel JM. Excitatory effects of hypocretin-1 (orexin-A) in the trigeminal motor nucleus are reversed by NMDA antagonism. J Neurophysiol. 2003;89:2591–2600. doi: 10.1152/jn.00968.2002. [DOI] [PubMed] [Google Scholar]
  101. 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. 1998;18:9996–10015. doi: 10.1523/JNEUROSCI.18-23-09996.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Pierangeli G, Provini F, Maltoni P, Barletta G, Contin M, Lugaresi E, Montagna P, Cortelli P. Nocturnal body core temperature falls in Parkinson’s disease but not in Multiple-System Atrophy. Mov Disord. 2001;16:226–232. doi: 10.1002/mds.1039. [DOI] [PubMed] [Google Scholar]
  103. Piper DC, Upton N, Smith MI, Hunter AJ. The novel brain neuropeptide, orexin-A, modulates the sleep-wake cycle of rats. Eur J Neurosci. 2000;12:726–730. doi: 10.1046/j.1460-9568.2000.00919.x. [DOI] [PubMed] [Google Scholar]
  104. Randeva HS, Karteris E, Grammatopoulos D, Hillhouse EW. Expression of orexin-A and functional orexin type 2 receptors in the human adult adrenals: Implications for adrenal function and energy homeostasis. J Clin Endocrinol Metab. 2001;86:4808–4813. doi: 10.1210/jcem.86.10.7921. [DOI] [PubMed] [Google Scholar]
  105. Rauh RA, Senior DG, Miller WP. Delayed complete heart block complicating percutaneous transluminal coronary angioplasty. Am Heart J. 1990;120:972–975. doi: 10.1016/0002-8703(90)90218-m. [DOI] [PubMed] [Google Scholar]
  106. Ravussin E, Bogardus C. A brief overview of human energy metabolism and its relationship to essential obesity. Am J Clin Nutr. 1992;55:242S–245S. doi: 10.1093/ajcn/55.1.242s. [DOI] [PubMed] [Google Scholar]
  107. Ravussin E, Lillioja S, Anderson TE, Christin L, Bogardus C. Determinants of 24-hour energy expenditure in man. Methods and results using a respiratory chamber. J Clin Invest. 1986;78:1568–1578. doi: 10.1172/JCI112749. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Rodgers RJ, Ishii Y, Halford JC, Blundell JE. Orexins and appetite regulation. Neuropeptides. 2002;36:303–325. doi: 10.1016/s0143-4179(02)00085-9. [DOI] [PubMed] [Google Scholar]
  109. Sakurai T. Reverse pharmacology of orexin: From an orphan GPCR to integrative physiology. Regul Pept. 2005;126:3–10. doi: 10.1016/j.regpep.2004.08.006. [DOI] [PubMed] [Google Scholar]
  110. Sakurai T, Amemiya A, Ishii M, Matsuzaki I, Chemelli RM, Tanaka H, Williams SC, Richardson JA, Kozlowski GP, Wilson S, Arch JR, Buckingham RE, et al. Orexins and orexin receptors: A family of hypothalamic neuropeptides and G protein-coupled receptors that regulate feeding behavior. Cell. 1998;92:573–585. doi: 10.1016/s0092-8674(00)80949-6. [DOI] [PubMed] [Google Scholar]
  111. Samson WK, Gosnell B, Chang JK, Resch ZT, Murphy TC. Cardiovascular regulatory actions of the hypocretins in brain. Brain Res. 1999;831:248–253. doi: 10.1016/s0006-8993(99)01457-2. [DOI] [PubMed] [Google Scholar]
  112. Samson WK, Taylor MM, Ferguson AV. Non-sleep effects of hypocretin/orexin. Sleep Med Rev. 2005;9:243–252. doi: 10.1016/j.smrv.2004.07.006. [DOI] [PubMed] [Google Scholar]
  113. Saper CB, Fuller PM, Pedersen NP, Lu J, Scammell TE. Sleep state switching. Neuron. 2010;68:1023–1042. doi: 10.1016/j.neuron.2010.11.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Sato-Suzuki I, Kita I, Seki Y, Oguri M, Arita H. Cortical arousal induced by microinjection of orexins into the paraventricular nucleus of the rat. Behav Brain Res. 2002;128:169–177. doi: 10.1016/s0166-4328(01)00307-2. [DOI] [PubMed] [Google Scholar]
  115. Sellayah D, Bharaj P, Sikder D. Orexin is required for brown adipose tissue development, differentiation, and function. Cell Metab. 2011;14:478–490. doi: 10.1016/j.cmet.2011.08.010. [DOI] [PubMed] [Google Scholar]
  116. 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) 2009;33:775–785. doi: 10.1038/ijo.2009.96. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Shirasaka T, Nakazato M, Matsukura S, Takasaki M, Kannan H. Sympathetic and cardiovascular actions of orexins in conscious rats. Am J Physiol. 1999;277:R1780–R1785. doi: 10.1152/ajpregu.1999.277.6.R1780. [DOI] [PubMed] [Google Scholar]
  118. Shirasaka T, Kunitake T, Takasaki M, Kannan H. Neuronal effects of orexins: Relevant to sympathetic and cardiovascular functions. Regul Pept. 2002;104:91–95. doi: 10.1016/s0167-0115(01)00352-4. [DOI] [PubMed] [Google Scholar]
  119. Siegel JM. Narcolepsy: A key role for hypocretins (orexins) Cell. 1999;98:409–412. doi: 10.1016/s0092-8674(00)81969-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Smith PM, Connolly BC, Ferguson AV. Microinjection of orexin into the rat nucleus tractus solitarius causes increases in blood pressure. Brain Res. 2002;950:261–267. doi: 10.1016/s0006-8993(02)03048-2. [DOI] [PubMed] [Google Scholar]
  121. Smith MI, Piper DC, Duxon MS, Upton N. Evidence implicating a role for orexin-1 receptor modulation of paradoxical sleep in the rat. Neurosci Lett. 2003;341:256–258. doi: 10.1016/s0304-3940(03)00066-1. [DOI] [PubMed] [Google Scholar]
  122. Smith PM, Samson WK, Ferguson AV. Cardiovascular actions of orexin-A in the rat subfornical organ. J Neuroendocrinol. 2007;19:7–13. doi: 10.1111/j.1365-2826.2006.01497.x. [DOI] [PubMed] [Google Scholar]
  123. Sutcliffe JG, de Lecea L. The hypocretins: Excitatory neuromodulatory peptides for multiple homeostatic systems, including sleep and feeding. J Neurosci Res. 2000;62:161–168. doi: 10.1002/1097-4547(20001015)62:2<161::AID-JNR1>3.0.CO;2-1. [DOI] [PubMed] [Google Scholar]
  124. Szekely M, Petervari E, Balasko M, Hernadi I, Uzsoki B. Effects of orexins on energy balance and thermoregulation. Regul Pept. 2002;104:47–53. doi: 10.1016/s0167-0115(01)00348-2. [DOI] [PubMed] [Google Scholar]
  125. Takahashi K, Lin JS, Sakai K. Neuronal activity of orexin and non-orexin waking-active neurons during wake-sleep states in the mouse. Neuroscience. 2008;153:860–870. doi: 10.1016/j.neuroscience.2008.02.058. [DOI] [PubMed] [Google Scholar]
  126. Tang J, Chen J, Ramanjaneya M, Punn A, Conner AC, Randeva HS. The signalling profile of recombinant human orexin-2 receptor. Cell Signal. 2008;20:1651–1661. doi: 10.1016/j.cellsig.2008.05.010. [DOI] [PubMed] [Google Scholar]
  127. 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. 2006;291:R889–R899. doi: 10.1152/ajpregu.00536.2005. [DOI] [PubMed] [Google Scholar]
  128. Teske JA, Billington CJ, Kotz CM. Hypocretin/orexin and energy expenditure. Acta Physiol (Oxf) 2010;198:303–312. doi: 10.1111/j.1748-1716.2010.02075.x. [DOI] [PubMed] [Google Scholar]
  129. Thakkar MM, Ramesh V, Cape EG, Winston S, Strecker RE, McCarley RW. REM sleep enhancement and behavioral cataplexy following orexin (hypocretin)-II receptor antisense perfusion in the pontine reticular formation. Sleep Res Online. 1999;2:112–120. [PubMed] [Google Scholar]
  130. Thorpe AJ, Kotz CM. Orexin A in the nucleus accumbens stimulates feeding and locomotor activity. Brain Res. 2005;1050:156–162. doi: 10.1016/j.brainres.2005.05.045. [DOI] [PubMed] [Google Scholar]
  131. Trivedi P, Yu H, MacNeil DJ, Van der Ploeg LH, Guan XM. Distribution of orexin receptor mRNA in the rat brain. FEBS Lett. 1998;438:71–75. doi: 10.1016/s0014-5793(98)01266-6. [DOI] [PubMed] [Google Scholar]
  132. Tsunematsu T, Kilduff TS, Boyden ES, Takahashi S, Tominaga M, Yamanaka A. Acute optogenetic silencing of orexin/hypocretin neurons induces slow-wave sleep in mice. J Neurosci. 2011;31:10529–10539. doi: 10.1523/JNEUROSCI.0784-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Tupone D, Madden CJ, Cano G, Morrison SF. An orexinergic projection from perifornical hypothalamus to raphe pallidus increases rat brown adipose tissue thermogenesis. J Neurosci. 2011;31:15944–15955. doi: 10.1523/JNEUROSCI.3909-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. van den Pol AN, Gao XB, Obrietan K, Kilduff TS, Belousov AB. Presynaptic and postsynaptic actions and modulation of neuroendocrine neurons by a new hypothalamic peptide, hypocretin/orexin. J Neurosci. 1998;18:7962–7971. doi: 10.1523/JNEUROSCI.18-19-07962.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  135. Vijgen GH, Bouvy ND, Teule GJ, Brans B, Schrauwen P, van Marken Lichtenbelt WD. Brown adipose tissue in morbidly obese subjects. PLoS One. 2011;6:e17247. doi: 10.1371/journal.pone.0017247. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Wang J, Osaka T, Inoue S. Energy expenditure by intracerebroventricular administration of orexin to anesthetized rats. Neurosci Lett. 2001;315:49–52. doi: 10.1016/s0304-3940(01)02322-9. [DOI] [PubMed] [Google Scholar]
  137. Wang J, Osaka T, Inoue S. Orexin-A-sensitive site for energy expenditure localized in the arcuate nucleus of the hypothalamus. Brain Res. 2003;971:128–134. doi: 10.1016/s0006-8993(03)02437-5. [DOI] [PubMed] [Google Scholar]
  138. Willie JT, Chemelli RM, Sinton CM, Yanagisawa M. To eat or to sleep? Orexin in the regulation of feeding and wakefulness. Annu Rev Neurosci. 2001;24:429–458. doi: 10.1146/annurev.neuro.24.1.429. [DOI] [PubMed] [Google Scholar]
  139. Willie JT, Chemelli RM, Sinton CM, Tokita S, Williams SC, Kisanuki YY, Marcus JN, Lee C, Elmquist JK, Kohlmeier KA, Leonard CS, Richardson JA, et al. Distinct narcolepsy syndromes in orexin receptor-2 and orexin null mice: Molecular genetic dissection of Non-REM and REM sleep regulatory processes. Neuron. 2003;38:715–730. doi: 10.1016/s0896-6273(03)00330-1. [DOI] [PubMed] [Google Scholar]
  140. Wu MF, John J, Maidment N, Lam HA, Siegel JM. Hypocretin release in normal and narcoleptic dogs after food and sleep deprivation, eating, and movement. Am J Physiol Regul Integr Comp Physiol. 2002;283:R1079–R1086. doi: 10.1152/ajpregu.00207.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  141. Xi MC, Morales FR, Chase MH. Effects on sleep and wakefulness of the injection of hypocretin-1 (orexin-A) into the laterodorsal tegmental nucleus of the cat. Brain Res. 2001;901:259–264. doi: 10.1016/s0006-8993(01)02317-4. [DOI] [PubMed] [Google Scholar]
  142. Yamanaka A, Muraki Y, Tsujino N, Goto K, Sakurai T. Regulation of orexin neurons by the monoaminergic and cholinergic systems. Biochem Biophys Res Commun. 2003;303:120–129. doi: 10.1016/s0006-291x(03)00299-7. [DOI] [PubMed] [Google Scholar]
  143. Yang B, Ferguson AV. Orexin-A depolarizes nucleus tractus solitarius neurons through effects on nonselective cationic and K+ conductances. J Neurophysiol. 2003;89:2167–2175. doi: 10.1152/jn.01088.2002. [DOI] [PubMed] [Google Scholar]
  144. Yoshida Y, Fujiki N, Nakajima T, Ripley B, Matsumura H, Yoneda H, Mignot E, Nishino S. Fluctuation of extracellular hypocretin-1 (orexin A) levels in the rat in relation to the light-dark cycle and sleep-wake activities. Eur J Neurosci. 2001;14:1075–1081. doi: 10.1046/j.0953-816x.2001.01725.x. [DOI] [PubMed] [Google Scholar]
  145. Zepelin H, McDonald CS. Age differences in autonomic variables during sleep. J Gerontol. 1987;42:142–146. doi: 10.1093/geronj/42.2.142. [DOI] [PubMed] [Google Scholar]
  146. Zhang L, Renaud LP, Kolaj M. Properties of a T-type Ca2+ channel-activated slow after hyperpolarization in thalamic paraventricular nucleus and other thalamic midline neurons. J Neurophysiol. 2009;101:2741–2750. doi: 10.1152/jn.91183.2008. [DOI] [PubMed] [Google Scholar]
  147. Zheng H, Patterson LM, Berthoud HR. Orexin-A projections to the caudal medulla and orexin-induced c-Fos expression, food intake, and autonomic function. J Comp Neurol. 2005;485:127–142. doi: 10.1002/cne.20515. [DOI] [PubMed] [Google Scholar]

RESOURCES