Abstract

The hypocretin/orexin (Hcrt/Orx) system in the perifornical lateral hypothalamus has been recognized as a critical node in a complex network of neuronal systems controlling both physiology and behavior in vertebrates. Our understanding of the Hcrt/Orx system and its array of functions and actions has grown exponentially in merely 2 decades. This review will examine the latest progress in discerning the roles played by the Hcrt/Orx system in regulating homeostatic functions and in executing instinctive and learned behaviors. Furthermore, the gaps that currently exist in our knowledge of sex-related differences in this field of study are discussed.

ESSENTIAL POINTS
  1. Although hypocretin/orexin (Hcrt/orx) was originally implicated as an appetite-promoting peptide in animals, it is now well established that the Hcrt/orx system regulates a wide array of physiological homeostatic processes (such as energy balance, sleep-wake cycle, and others) and controls complex behaviors (such as reward-seeking and motivation, cognition, stress coping, aggression, and others) through overlapping neuronal circuits in the brain.

  2. Dysregulated Hcrt/orx system resulting from genetic (such as a deficiency in Hcrt/orx peptides or receptors) and environmental factors (such as obesity) leads to neurological and psychiatric diseases and conditions (such as narcolepsy, deficiency in reward-seeking and stress coping, depression and anxiety, panic attack, and others) in animal models and human subjects.

  3. The expression of Hcrt/orx peptides and receptors exhibits sex differences through hormonal and non-hormonal factors. In addition, direct regulation of male and female reproductive functions and maternal behaviors by Hcrt/orx have been emerging. Nevertheless, our understanding of physiological homeostasis and complex behaviors regulated by the Hcrt/orx system has been mostly achieved from investigations in male animals.

  4. The substantial knowledge gaps in the understanding of sex differences of Hcrt/orx-mediated regulation of homeostatic functions and complex behaviors at molecular, cellular and systems levels in animals and humans need to be addressed.

All animals on planet Earth must forage for and consume energy, respond to internal and external stimuli, acclimate to their environment, and produce offspring to secure survival. Within the brain of all vertebrate animals, the hypothalamus (as part of the limbic system), is well known as a link between the central nervous and endocrine systems via the pituitary gland. During the latter half of the 20th century, the hypothalamus was shown to be a critical brain area in vertebrates that regulates essential functions necessary for the survival of individuals and their species (1). This groundbreaking work established the basis for our current understanding of prominent hypothalamic structures. However, the discovery of the hypocretin/orexin (Hcrt/Orx) system slightly more than 2 decades ago and the elucidation of its role in the regulation of brain functions has advanced our understanding of neurobiological processes within the hypothalamus in an unprecedented way. These processes underlie complex behaviors in animals as well as neurological and psychiatric conditions in humans. As summarized in Fig. 1, we now know that the Hcrt/Orx system participates in the homeostatic regulation of many basic physiological functions, such as energy balance, electrolyte-fluid balance, the sleep-wake cycle and circadian rhythm, body temperature and thermoregulation, cardiovascular and respiratory activity, neuro-immune responses, and so on (2-3). In addition to regulating these basic physiological processes, the Hcrt/Orx system integrates information and, thus, controls many complex behaviors in animals, including cognition and learning, motivated and reward-seeking behaviors, stress responses to natural and social stressors, “fight or flight” response, and emotional regulation (3-5). Dysregulations and deficiencies in 1 or several physiological functions and/or complex behaviors mediated by the Hcrt/Orx system can lead to many diverse diseases and conditions in humans, including obesity and eating disorders (Prader-Willi syndrome), addiction to illicit drugs and other natural and social reinforcers, sleep disorders and insomnia, depression and anxiety, and schizophrenia (3, 6).

A diagram depicting major functions of the Hcrt/Orx system in animals. Brain circuits regulating physiological homeostasis and controlling complex behavior converge onto the Hcrt/Orx system, which becomes an essential hub in the brain to mediate the interplay between physiological homeostasis and complex behavior in animals. A dysregulated Hcrt/Orx system resulting from altered homeostasis (such as obesity) may contribute to the development of neurological and psychiatric diseases/conditions (eg, lowered arousal levels, weakened responses to salient stimuli) in animals as reported recently (49).
Figure 1.

A diagram depicting major functions of the Hcrt/Orx system in animals. Brain circuits regulating physiological homeostasis and controlling complex behavior converge onto the Hcrt/Orx system, which becomes an essential hub in the brain to mediate the interplay between physiological homeostasis and complex behavior in animals. A dysregulated Hcrt/Orx system resulting from altered homeostasis (such as obesity) may contribute to the development of neurological and psychiatric diseases/conditions (eg, lowered arousal levels, weakened responses to salient stimuli) in animals as reported recently (49).

The neuropeptide hypocretin (orexin) was discovered by 2 groups of researchers independently at the end of the last century (7, 8). Interestingly, a novel clone named “clone 35” was shown to encode a pre-prohormone exclusively expressed in the hypothalamus of rats; it was identified through directional tag PCR subtraction as early as 1996 (9). Eventually, 2 novel peptidergic hormones named hypocretin 1 and 2, specifically synthesized in neurons of the perifornical/lateral hypothalamus, were found to be the products of this pre-prohormone through proteolytic cleavage (7). The naming of hypocretin is the result of this neuropeptide with specific expression in the hypothalamus and aminoacidic similarities to the peptide secretin. At the same time, Sakurai et al (1998) discovered 2 G protein-coupled receptors and their natural ligands that are exclusively expressed in the lateral hypothalamus and promote food intake in animals. Therefore, they named these new peptides as orexins (orexin-a and orexin-b) (8). We now know that hypocretin and orexin are the same peptidergic hormone and that the orexigenic effect merely accounts for a small fraction of the functions governed by this peptide and may be secondary to its arousal- and motivational-promoting effects (Fig. 1).

Genetic Determination and Diversity in Hcrt Cells

The expression of the Hcrt/Orx peptide starts early in development and the neuropeptides dynorphin and nociceptin/orphanin FQ coexist with Hcrt/Orx in LHA neurons (10-12). Genetic analyses have revealed many genes that regulate the expression of Hcrt/Orx under normal and pathological conditions. It was initially shown that genes such as the IGF binding protein 3 (Igfbp3), transcript encoding Tribbles homolog 2 (Trib2), and transcription factor Lhx9 were likely regulators of the Hcrt gene expression, whereas the potassium channel subunit gene Kcnh4a regulated activities in Hcrt/Orx cells in animals. This may account for the loss of function of the Hcrt/Orx system in narcoleptic patients and animal models, although they might not be specific to Hcrt/Orx cells (13-16). Later reports showed that a growing number of genes, including Ahr1, Pcsk1, Pdyn, Peg3, Plagl1, Lhx9, Nek7, Nr2f2, Nptx2, Prrx1, Rfx4, and Six6, were specific markers or fate determinants of Hcrt cells (17, 18). It has been estimated that there are about a few thousand Hcrt/Orx cells in rodents (such as proximately 2000 in male mice) and 50 000 to 90 000 in humans (19-22). The numbers of Hcrt/Orx cells fluctuate throughout the life span of animals. For instance, in C57B6 mice, numbers of Hcrt/Orx cells in both sexes were stable in the first 400 days of life, then started to slowly decline between 400 and 800 days, and with an accelerated rate of loss between 800 and 1000 days of age, at least in males (22). It is also worth noting that male mice exhibits 15% to 20% more Hcrt/Orx-immunopositive cells than females (22). Despite the small number of Hcrt/Orx cells in the brain, nerve fibers containing Hcrt/Orx and distributions of Hcrt/Orx receptors were identified throughout the central nervous system and even in peripheral organs (19, 23-27). Therefore, it is conceivable that Hcrt/Orx cells likely exhibit a heterogeneity in terms of molecular and cellular markers, broad and diverse projections of nerve fibers in the brain, and regulation of a wide array of functions in animals. An emerging body of evidence strongly suggests this possibility. Hcrt/Orx cells are glutamatergic in nature as indicated by early and most recent reports (17, 28-30). However, one-half of these cells also express gamma-aminobutyric acid (GABA) synthetic enzyme gene Gad1 (30). The role of Gad1 in Hcrt/Orx cells is not yet clear because Hcrt/Orx cells do not express an intact vesicular releasing machinery for GABA as typically seen in GABAergic neurons (30); it is intriguing and vitally important to understand the functional implications of the existence of GABA in Hcrt cells. Interestingly, Hcrt/Orx cells can be divided into 2 populations depending on the expression of sex-specific genes Ddx3y and Xist, suggesting possible sex-related dimorphism in these cells (17). The gene Ddx3y, which is located on the Y chromosome and is responsible for male fertility and spermatogenesis, plays a role in neuronal differentiation during development in animals (31). The gene Xist is located on the X chromosome and is responsible for X-inactivation in early development (32). However, it is not clear what effects these genes may have on Hcrt/Orx cells in male and female animals and whether their presence contributes to sexual dimorphism in a hypocretin-dependent way in animals.

Basic Properties of the Hcrt/Orx System

The Hcrt/Orx peptides were demonstrated to be excitatory when they were initially discovered (7). At the cellular level, Hcrt/Orx enhances synaptic transmission and neuronal activity, elevates intracellular calcium levels, and promotes synaptic potentiation in various neuronal types in the brain (33-37). The cellular effects of Hcrt/Orx are mediated by 2 G protein-coupled receptors, which are mediated by the Gs, Gq, and Gi pathways (8, 38). The PLC and PKC pathways downstream to the activation of the Gq protein likely contribute to the excitatory effects of Hcrt/Orx in neurons (38). In addition to G proteins, Hcrt/Orx receptors may also be coupled to β-arrestin, protein tyrosine phosphatase SHP-2, dynein light chain Tctex type 3, and other proteins to transduce broad effects of Hcrt/Orx (38).

Hcrt/Orx-expressing neurons themselves possess many distinctive characteristics compared with other neuronal types in the brain. First, at the single-cell level, Hcrt/Orx cells can be divided into 2 populations (D and H type), depending on their distinct electrophysiological properties (39). H-type cells had a larger 4-AP-sensitive, low-activation threshold A-current than D-type cells (39). H-type cells were sustained hyperpolarized by an elevation of extracellular glucose levels in slices, whereas D-type cells only exhibited a transient hyperpolarization (39). However, it has not been exactly clear whether and how the proposed H-type and D-type cells mediated a diverse array of functions and behaviors in animals. By using conventional and perforated whole-cell patch clamp recordings, Liu and colleagues showed that the intracellular ATP levels were generally higher than in other types of central neurons in the brain and that the intracellular levels of ATP were lower in Hcrt/Orx neurons in sleeping mice than in sleep-deprived animals (40). The cellular responses to changes in internal (intercellular ATP levels) and external cues (glucose, lactate, and amino acid levels) constitute the basis of how the Hcrt/Orx system may serve as a sensor of energy availability in the brain (40-44).

Second, at the synaptic level, Horvath and Gao (2005) showed that the frequency of miniature excitatory postsynaptic currents recorded at the soma of Hcrt/Orx cells was about 10-fold larger than that of miniature inhibitory postsynaptic currents (45). Consistently, there were more asymmetric (putatively excitatory) than symmetric (putatively inhibitory) synapses on cell bodies of Hcrt/Orx neurons (45). These results strongly indicate that their cell bodies are predominately under the innervation of excitatory (likely glutamatergic) but not inhibitory (likely GABAergic) synapses (45, 46). This unique synaptic architecture is distinctively different from other long projection neurons such as pyramidal neurons in the neocortex, in which the numbers of inhibitory (GABAergic) synapses predominate over numbers of excitatory (glutamatergic) synapses at the soma of these neurons; there are more inhibitory synapses than excitatory synapses on soma of pyramidal neurons in the neocortex (47). This unique arrangement of excitatory and inhibitory synapses on Hcrt/Orx neurons is consistent with the reports that the blockade of ionotropic glutamatergic but not GABAA-mediated transmission onto Hcrt/Orx cells with selective glutamatergic antagonists significantly attenuated the spontaneous action potential firing in these neurons (46, 48). Our recent data suggest that predominately excitatory inputs onto Hcrt/Orx cells might be weakened in obese animals compared with their control counterparts (49). Specifically, in male diet-induced obese (DIO) mice, there were relatively more inhibitory inputs onto Hcrt/Orx neuronal soma than in controls fed with a normal diet (49). The release probability of glutamatergic synapses was decreased and expression of spike-timing dependent plasticity at glutamatergic synapses was impaired in Hcrt/Orx cells in DIO mice (49).

Among the glutamatergic synapses on the Hcrt/Orx cells, there are fewer “silent” synapses among them, in that only NMDA-type receptors (NMDARs) but not AMPA-type glutamate receptors (AMPARs) are expressed (50, 51). Our results indicated that the AMPAR-mediated EPSC is significantly larger (> 2-fold) than that mediated by NMDARs (the AMPAR/NMDAR ratio is larger than 1.0) (52-54), strikingly different from what has been reported for the AMPAR/NMDAR ratio (being less than 1.0) in CA1 pyramidal neurons of the hippocampus and dopamine neurons in the ventral tegmental area (VTA) (51, 55). Therefore, it is reasonable to postulate that a small, excitatory input mediated by glutamate may lead to proportionately more activated glutamatergic synapses on Hcrt/Orx neurons than on other central neurons such as CA1 pyramidal neurons; the reason being that silent synapses containing only NMDARs require a sufficient depolarization to remove the Mg2+ blockade of NMDARs to conduct glutamatergic transmission (51) and thus the synaptic transmission would “fail” at these synapses when the postsynaptic component is not adequately excited by other factors, such as a concurrent activation of AMPARs. Therefore, it is expected that glutamatergic transmission onto Hcrt/Orx neurons might be highly efficient without silent synapses, a necessity for the Hcrt/Orx system to mediate its functions.

In addition, GluR2-lacking calcium-permeable AMPARs (Cp-AMPARs) are expressed in Hcrt/Orx cells under basal conditions as indicated by our studies (54). The NMDAR-mediated calcium influx is critical to synaptic plasticity (such as LTP) in central neurons (56). Cp-AMPAR is also recognized as a critical promoter to the development of synaptic plasticity in the brain. In pyramidal cells in the CA1 region of the hippocampus, Cp-AMPARs are not incorporated at postsynaptic sites under basal conditions, but only inserted transiently to postsynaptic sites during the induction phase of expression of NMDAR-dependent LTP (57). In central neurons, such as interneurons of the basolateral amygdala, where Cp-AMPARs are constantly expressed under basal conditions, the expression of LTP is NMDAR-independent and mediated by Cp-AMPARs (58). The existence of Cp-AMPARs at glutamatergic synapses on Hcrt cells implicate a distinctive route for calcium influx during glutamate-induced excitation and expression of synaptic plasticity in these cells.

At the circuit level, emerging evidence has presented a more complicated network of neuronal systems that interact with Hcrt/Orx neurons locally in the LHA than previously reported (Fig. 2). It was originally described that Hcrt/Orx cells had reciprocal synaptic connections with neighboring melanin-concentrating hormone (MCH) containing neurons (59, 60). Exogeneous Hcrt/Orx could directly activate MCH cells, implying a functional role of Hcrt/Orx released from presynaptic boutons onto MCH cells (60). However, exogeneous MCH did not directly modulate activity in Hcrt/Orx cells (53), which is consistent with a report that Hcrt/Orx cells do not express MCH receptor 1 (MCHR1), the only MCH receptor subtype in rodents (61). Because the Hcrt/Orx system could be indirectly suppressed by exogenous MCH in wild-type animals and up-regulated in male MCHR1-deficient mice (53), it is likely that Hcrt/Orx cells are regulated indirectly by MCH through a population of MCHR1-expressing cells. It was originally reported that Hcrt/Orx cells were inhibited by leptin (62). Later, it was shown that Hcrt/Orx cells do not express the long form leptin receptor (LepRb) (63). Instead, a population of GABAergic neurons containing the peptide neurotensin express this specific receptor LepRb in the LHA and directly innervate Hcrt/Orx cells (63, 64). In recent years, another population of GABAergic neurons were identified to express dynorphin but not Hcrt/Orx receptors and to innervate Hcrt cells in the LHA, which provides an additional inhibitory circuit onto these cells locally (65). There was also a report that some populations of GABAergic neurons expressed Hcrt/Orx receptors and innervated MCH cells in the LHA; these cells have been proposed to provide a local inhibitory circuit onto MCH cells (66).

Local neuronal populations that interact with the Hcrt/Orx system. Hcrt/Orx cells and neighboring MCH-containing neurons receive reciprocal synaptic connections between them (59, 60). The Hcrt/Orx system is likely to be regulated indirectly by MCH and leptin through populations of MCHR1-expressing and LepRb-expressing neurons (61, 63, 64). A group of GABAergic neurons expressing dynorphin but not Hcrt/Orx receptors innervate Hcrt cells in the LHA, providing an additional inhibitory circuit onto these cells locally (65). Some other populations of GABAergic neurons expressing Hcrt/Orx receptors innervate MCH cells and may provide an inhibitory pathway onto these cells locally (66).
Figure 2.

Local neuronal populations that interact with the Hcrt/Orx system. Hcrt/Orx cells and neighboring MCH-containing neurons receive reciprocal synaptic connections between them (59, 60). The Hcrt/Orx system is likely to be regulated indirectly by MCH and leptin through populations of MCHR1-expressing and LepRb-expressing neurons (61, 63, 64). A group of GABAergic neurons expressing dynorphin but not Hcrt/Orx receptors innervate Hcrt cells in the LHA, providing an additional inhibitory circuit onto these cells locally (65). Some other populations of GABAergic neurons expressing Hcrt/Orx receptors innervate MCH cells and may provide an inhibitory pathway onto these cells locally (66).

In summary, the unique glutamatergic and GABAergic synaptic organization onto the somata of Hcrt/Orx neurons, distinctive composition of subtypes of glutamate receptors in these cells, and a complex network of interaction between Hcrt/Orx cells and neighboring neuronal populations provide a functional basis for the role that Hcrt/Orx cells may play in the regulation of critical functions in animals and humans. The predominant innervation by excitatory synapses onto the cell bodies of Hcrt neurons and their putatively high efficiency in glutamatergic transmission are likely to facilitate a rapid activation of this system upon exposure to salient stimuli and environmental cues that, in turn, affects the daily tasks and overall survival of animals. It should be clearly stated, however, that all our knowledge regarding the Hcrt/Orx system described here was predominantly obtained from studies on male animals. There are still significant gaps in our understanding of the basic properties of the Hcrt/Orx system in female animals. Importantly, the influence of estrogen- and non-estrogen-dependent factors on sex differences in the Hcrt/Orx system has not been thoroughly examined. A recent report provides emerging evidence of possible changes in Hcrt/Orx cells in naturally cycling female rodents (67). The authors suggested that both endogenous and exogenous estrogen (E2) might have an inhibitory effect on Hcrt/Orx cells in naturally cycling female rats (67). However, it was not clear which types of estrogen receptors mediated the reported changes in synaptic transmission onto the Hcrt cells by E2. In fact, there was a report that immunolabeling for androgen (AR) and estrogen (ER alpha) receptors revealed no colocalization of Hcrt/Orx with AR and few Hcrt/Orx neurons expressing ER alpha in male rodents, suggesting that hormonal regulation of Hcrt/Orx expression might be via afferents from neurons containing those receptors (68).

Homeostatic Regulation of Physiological Functions by Hcrt/Orx

Physiological functions and behaviors are closely associated because of overlapping neural circuits that mediate both processes. Specifically, the Hcrt/Orx system is among neurocircuits that regulate both homeostatic functions of the brain and complex behaviors in animals. The intake and use of energy and the maintenance of optimal arousal levels are fundamental functions that allow animals to perform daily tasks required for survival. In the following sections, we discuss new developments in the regulation of these essential physiological functions (ie, energy metabolism and the sleep-wake cycle/arousal control) and behaviors (eg, reward seeking, stress coping) by the Hcrt/Orx system with consideration given to sex dimorphism.

Bidirectional Control of Energy Balance

The intake and use of energy are fundamental functions essential to the performance of daily activities in all animals, especially vertebrates. The LHA is a classic structure that participates in the regulation of food intake in animals. The pioneering work of Hetherington and Ranson (69) and Anand and Brobeck (70, 71) led to the characterization of the LHA as a feeding center (72, 73). Although the idea of feeding center has been supported by evidence obtained from electrical stimulations, selective chemical lesions, and in vivo electrophysiological recording in animals (74-78), the exact identities of the neurons responsible for this role of the LHA in positive energy balance was not clear until the discovery of diverse neuronal populations in this region.

Hcrt/Orx promotes both energy intake and expenditure in animals. The intracerebroventricular injection of Hcrt/Orx (either Hcrt 1/Orx A or Hcrt 2/Orx B) led to a short-term increase in food consumption in male rats (8). Local administration of Hcrt 1 to the paraventricular nucleus, dorsomedial nucleus, and lateral hypothalamic/perifornical area triggered food intake in male rats (79, 80). Conversely, the application of a selective Hcrt-1 receptor antagonist, SB-334867, suppressed food intake in male and female rodents (81-83). Consistent with the effects of exogeneous Hcrt/Orx on feeding, the activity in Hcrt neurons is in close alignment with energy and nutrient states in animals. Food deprivation and acute hypoglycemia resulted in up-regulation of Hcrt/Orx mRNA, Hcrt/Orx expression and c-fos expression in Hcrt/Orx neurons in male rodents and female nonhuman primates (8, 84-89), which could be inhibited by leptin treatment (90, 91). In addition to its effects on central mechanisms, Hcrt/Orx also facilitated energy intake in peripheral organs via both central and peripheral pathways. Central administration of Hcrt/Orx increased both gastric acid secretion and gastric motor function in male rodents (92, 93). On the one hand, this is likely mediated by the dorsal motor nucleus of the vagus (DMV), a key region that controls gastric acid secretion and gut motility. Hcrt/Orx neurons innervate this area and directly excite gastrointestinal-projecting DMV neurons (19, 94). Hypoglycemia triggered c-fos expression in the DMV through the activation of Hcrt/Orx neurons and the release of Hcrt/Orx peptide in male rats (95). On the other hand, Hcrt/Orx could directly stimulate contraction of duodenal smooth muscles at the periphery in female mice (96). Hcrt/Orx directly induced depolarization in duodenal longitudinal muscle cells through activation of receptor-operated, store-operated, and voltage-gated Ca2+ channels and inhibition of K+ channels (96). Under voltage clamp, Hcrt/Orx enhanced the amplitude of voltage-gated Na+ currents and lowed the threshold for Na+ channel activation. Similar effects of Hcrt/Orx on voltage-gated L- and T-type Ca2+ channels were observed (96).

Unlike other feeding-promoting neurons, such as neuropeptide Y (NPY) and agouti-related protein (AgRP) in the arcuate nucleus (ARC) and MCH cells in the LHA, which also suppress energy expenditure (97, 98), Hcrt/Orx promotes energy expenditure in animals. Microinjection of Hcrt-1 into the third ventricle, ARC, PVN, and LHA stimulated the metabolic rate in male mice and rats (97, 99-102). The increase in energy expenditure caused by Hcrt/Orx occurred in animals with and without an increase in physical activity. For example, a direct administration of Hcrt-1 into the ARC caused an enhancement in whole-body O2 consumption in male urethane-anesthetized rats (101), suggesting that Hcrt/Orx can boost energy expenditure even without an enhancement in physical activity (equivalent to the rest state) in animals. In conscious rats, Hcrt-1 enhanced energy use by increasing physical activity. Microinjections of Hcrt/Orx into the PVN and LHA significantly increased spontaneous non-exercise-related activities accompanying with an increment of in O2 consumption and thermogenesis in male rats (102-105), suggesting that Hcrt/Orx promotes non-exercise physical activity in animals. Consistent with these results, Hcrt/Orx increased and promoted glucose uptake, insulin-induced glucose uptake, and glycogen synthesis in skeletal muscle, but not in white adipose tissue, through the ventromedial hypothalamus (VMH)-sympathetic nervous system pathway in male mice and rats (106, 107). Furthermore, Hcrt/Orx was required in mediating mobilization of brown fat tissue (108, 109), which was supported by a study showing the direct innervation by Hcrt-containing nerve fibers from the LHA onto the raphe pallidus to promote BAT thermogenesis in male rats (110). Hcrt-mutant mice expressed a higher level of preadipocyte markers, fewer mitochondria, lowered levels of BAT thermogenic proteins (such as PPAR-γ1/γ2, PGC-1α/β, and UCP-1) in brown fat cells, compared with their wild-type littermates (108). These defects led to compromised thermogenesis induced by high-fat diet and cold exposure (108, 111). Based on these results, a recent report showed that Hcrt/Orx signaling was required to enhance energy expenditure and bodyweight loss in patients treated with D2 dopamine receptor agonists (109).

On-demand Control of Arousal by Hcrt/Orx

Higher animals execute many daily tasks, from exploring habitats and foraging for food to escaping from predators, to sustain their survival in the natural environment. During the reproductive season, they compete for mating partners and produce offspring to ensure the survival of the species. The successful performance of these functions likely depends on many well-controlled homeostatic processes based on a complex array of processing and computation of inputs within the brain that encode the internal and external environments of the animal. On the one hand, the promotion of general arousal across the sleep/wake cycle in concert with other wake-promoting centers in the brain is essential for the execution of homeostatic regulation of physiological functions (such as energy metabolism, water and salt balance, cardiovascular and respiratory functions), which is required to support daily tasks necessary for survival. On the other hand, Hcrt/Orx cells may be activated specifically to mediate goal-oriented behaviors beyond the normal regulation of the daily sleep/wake cycle, such as during a food deficit, or stress-inducing situation, and so on.

The involvement of the hypothalamus in sleep/wake regulation was originally proposed by van Economo based on his pioneering observations in human patients that lesions in the posterior hypothalamus and midbrain junction led to sleepiness, whereas anterior hypothalamic inflammation led to insomnia and chorea (112). Later, studies in monkeys, rats, and cats confirmed the sleep-promoting effect of lesions and inhibition of the posterior lateral hypothalamus (113-116). However, it remained unclear how the LHA participates in sleep regulation until the discovery of the neuropeptide Hcrt/Orx in this brain area (117, 118). It has already been shown that the levels of Hcrt/Orx in the cerebrospinal fluid align with the daily sleep/wake cycle: they are low during the resting (sleep) phase and high during the active (wake) phase in male rodents and male/female monkeys (119-122). The concentrations of Hcrt/Orx detected across the sleep/wake cycle are consistent with the activity levels of these cells as revealed by c-fos expression and recordings in vivo in freely moving or head-fixed male rodents (123-125). A direct and specific stimulation of Hcrt neurons by optogenetics increased arousal during the sleep phase, whereas an acute silencing of these cells increased just the slow-wave sleep-like activity during the sleep phase (126-128). The effects of Hcrt/Orx on sleep/wake regulation and arousal maintenance are likely due to its projections to major arousal regions such as the locus coeruleus, basal forebrain, and others (129-133). Although a deficiency in Hcrt peptide and its receptors (Hcrtr2 or OXR2) led to narcolepsy or narcoleptic-like phenotypes in dogs, mice, and human patients (20, 134-137), overexpression of Hcrt/Orx led to insomnia-like phenotypes in zebrafish (138).

In addition to its role in the regulation of the sleep/wake cycle, Hcrt/Orx also acts as a potent arousal promoter beyond the normal maintenance of the sleep/wake cycle in the brain. The activation of Hcrt/Orx cells was detected to be superimposed upon their normal functions of arousal promotion during the active phase. When male rats were required to maintain a prolonged wakefulness because of sleep deprivation, Hcrt/Orx cells were activated as indicated by c-fos expression in these cells (139). When animals were exposed to novel stimuli such as food, the activity in Hcrt/Orx cells increased during the initial period of feeding during the light phase (125, 140). The levels of Hcrt/Orx in the amygdala were maximal during positive emotion, social interaction, and anger in humans (141). Most importantly, activation of Hcrt/Orx neurons was required for the initiation of new locomotion but not ongoing locomotion (142). On top of this, Hcrt/Orx cells underwent synaptic plasticity of glutamatergic (excitatory) synapses induced by sleep deprivation in an activity-dependent way in male mice (52). Interestingly, the number of inhibitory synapses on Hcrt/Orx cells was decreased by sleep deprivation and increased during rebound sleep in zebrafish (143). The results suggest that the Hcrt/Orx system is required to maintain arousal and wakefulness when circumstances demand it, even when animals should be in the resting phase during their normal sleep/wake cycle. Additionally, synaptic plasticity in excitatory and inhibitory synapses upon the Hcrt cells occurs according to the behavioral state of the animal. Overall, these results further supported the idea that the activation of Hcrt/Orx cells is necessary to promote arousal on demand. However, we cannot exclude the possibility that changes in synaptic architectures and functions in Hcrt/Orx cells may result from both circadian and activity-dependent factors. Although there is evidence that the suprachiasmatic nuclei controls the release of Hcrt/Orx in male rodents (144-146), it is still not clear whether and how the circadian clock in the suprachiasmatic nuclei and core clock within Hcrt/Orx cells regulate synaptic plasticity in these cells in a way independent to activity levels in animals. Further investigations along these directions are warranted.

Use-dependent Regulation of Complex Behaviors by Hcrt/Orx

Control of Reward-seeking and Motivational Behaviors by Hcrt/Orx

The perifornical/lateral hypothalamic area has long been known as a brain structure responsible for reward-seeking behaviors. In early studies, electrical stimulation of the LHA induced an act of marked reinforcement and robust self-administration of electric current in rodents (147-150). Drugs of abuse also induce rodents to self-administrate when applied directly to the LHA (151, 152). It is now clear that Hcrt neurons mediate the development of drug addiction in animal models and human patients. These neurons are activated when rodents are exposed to drugs of abuse. The expression of c-fos and pCREB in Hcrt neurons is enhanced by opiates, cocaine, amphetamine, and nicotine in various models of drug-seeking behavior in male animals (54, 153-157). Direct administration of Hcrt into brain reward centers (such as the VTA) promotes drug seeking behavior in animals, including an increase in the breakpoint in a progressive-ratio task for cocaine and reinstatement of extinguished seeking behavior for morphine and nicotine (154, 157-160). The disruption of Hcrt receptor-mediated pathways with the selective Hcrt-1 receptor antagonist, SB334867, or a deficiency in Hcrt peptide and receptors attenuated, blocked, and abolished drug-seeking behavior induced by morphine, cocaine, amphetamine, and nicotine in animals (37,153, 154, 160-163). In human narcoleptic patients with a deficiency in Hcrt peptide or neurons, a significantly lowered tendency to drug abuse was reported (164). The action loci of Hcrt/Orx in mediating drug-seeking behaviors are identified throughout the brain, including the VTA (37, 164), basal forebrain (165), insular cortex (162), nucleus accumbens (166), hypothalamic paraventricular nucleus (PVN) (157), bed nucleus of the stria terminalis (167), paraventricular nucleus of the thalamus (168), and the central amygdala (169). There was evidence that Hcrt/Orx mediated drug-induced plasticity in these brain areas, which may be responsible for the role it plays in these regions. For example, Hcrt/Orx potentiated the NMDAR-mediated transmission by promoting the insertion of these receptors at glutamatergic synapses in VTA dopamine neurons, which was required to facilitate the expression of cocaine-induced plasticity in these cells and in the locomotor sensitization to cocaine in male animals (37). In addition, the Hcrt system per se undergoes synaptic plasticity in male rodents developing cocaine-seeking behaviors (57, 170). The long-lasting synaptic plasticity in Hcrt cells induced by cocaine exposure likely contributes to cue-induced pursuit of cocaine and morphine as well as sleep disorders in drug addicts and animals (54, 154), which has not been well explored. Our most recent data suggest that the impaired expression of activity-dependent synaptic plasticity in Hcrt/Orx cells might underlie the deficiency in the development of reward-seeking behaviors in male obese animals (49).

Another finding regarding the role that Hcrt-mediated motivational circuits play in the regulation of animal behavior is that the Hcrt system is involved in aggressive-like behaviors in animals. Flanigan et al reported that the projections from Hcrt cells to glutamic acid decarboxylase 2-expressing cells in the lateral habenula (LHb) was necessary to mediate intra-sex aggressive behaviors in male animals (171). Because Hcrt/Orx cells receive innervation from the VMH, a brain area well-known to be responsible for the initiation of aggressive behavior in male animals, the authors proposed that the pathway from Hcrt cells to glutamic acid decarboxylase 2 cells in the LHb may provide the link between aggressive and motivational circuits in the hypothalamus (171). The role of Hcrt/Orx in mediating aggressive behaviors in maternal animals has not been identified (172).

Hcrt/Orx-mediated Stress Responses and Stress Coping

The exposure to stressors is critical to both the development of drug addiction and relapse of drug-seeking behavior. The report that the Hcrt/Orx system mediates a stress-induced relapse of cocaine-seeking behavior provided the earliest evidence of the part that Hcrt/Orx cells play in mediating stress-triggered behavioral changes in male animals (158). The Hcrt/Orx system is a potent mediator of stress-induced effects on both physiological functions and behavioral changes in male and female animals (173). Hcrt/Orx cells are activated when animals are exposed to acute and chronic stress (49, 140, 158, 174-176). We now know that Hcrt cells respond to acute stress rather rapidly with a peaked phosphorylation of CREB about 5 minutes after stress exposure (49), which is similar to the activation pattern reported in corticotropin-releasing factor (CRF) neurons in the PVN (177). Considering that the expression of Hcrt/Orx induced by acute stress was still enhanced in the absence of CRF signaling (178), it is likely that the Hcrt/Orx system could serve as a pathway independent of the CRF system in the PVN to mediate certain stress-induced effects, which has not been well explored so far.

The physiological implications of stress-induced activation of Hcrt/Orx cells are emerging. In addition to triggering drug addiction relapses (158), the Hcrt/Orx cells are responsible for coping with acute stressors (49, 179, 180). The activation of Hcrt/Orx cells was required for expressing “claustrophobic” sighing when mice were subjected to confinement in a tiny space (180). The activity levels of Hcrt/Orx cells during stress exposure (such as forced cold water swim) causally contribute to stress-coping strategy (active coping vs passive coping) in male mice (49). Although certain stressors (such as foot shock) induce fear-conditioned learning, the Hcrt/Orx system may not be involved in this fear-associated learning (178, 181) but rather in the elimination of fear-related memories (182). Along this line, the activation of Hcrt/Orx cells is required to mediate anxiety and depressive-like behaviors after stress exposure and panic attack-like behaviors in both male and female animals (67, 176, 183-185). Therefore, the manipulation of the Hcrt/Orx-mediated pathway likely has therapeutic implications. The activation of orexinergic neurons by calorie restriction after social defeat elicited an antidepressant effect, as indicated by prolonged latency to immobility and shortened duration of immobility in the forced swim test (186). This may be mediated by the innervation of the LHb by Hcrt/Orx cells in the LHA (176). Also, an activated Hcrt/Orx system likely promotes adaptive responses to predator-scent stress (187). An early intervention with ORX-A/Hcrt1 reduced the prevalence of the posttraumatic stress disorder phenotype (adaptive responses to predator-scent stress) and increased the prevalence of adaptive phenotypes in animals (187).

Sex Differences in Hcrt/Orx-governed Functions and Behaviors in Animals

Hcrt/Orx specifically regulates reproductive functions and behaviors in males and females. The expression of Hcrt/Orx and its receptors exhibits sex dimorphism (188). In male animals, Hcrt and its receptors are expressed in the male reproductive system (189). In male rats, c-Fos expression was markedly increased in Hcrt/Orx neurons during copulation, whereas a systemic administration of the orexin-1 receptor antagonist, SB 334867, impaired copulatory behavior (68). Castration decreased the Hcrt/Orx neuron count and protein levels in a time course consistent with postcastration impairments in copulatory behavior (68). This effect was mediated by the VTA (68). However, there was also a report that the Hcrt/Orx system might not be critical for male sexual performance or motivation but rather plays a role in arousal and anxiety related to sexual behavior in naive male animals (190).

In female animals, Hcrt expression is increased in the proestrus phase of the estrus cycle (191). Increases in neuronal activity of Hcrt and in HCRT-1 receptor mRNA expression have been identified in association with lactation (192). At intermediate doses, intracerebroventricular injections of Hcrt-1 elevated levels of licking and grooming of pups and the number of nursing bouts. At the highest dose, Hcrt-1 delayed latency to nurse, decreased nursing, increased time off nest, and decreased maternal aggression. IP administration of the Hcrt-1 receptor antagonist, SB-334867, caused a trend toward an increase in low-arched back nursing activity and decrease in licking and grooming of pups during high-arched back nursing. This suggests that endogenously released Hcrt/Orx may be required for the full expression of maternal behaviors independently or dependently with other neuromodulators (192). However, there was also a report that the Hcrt/Orx levels in the medial preoptic area were negatively correlated with the frequency of contact with the litter or the display of erect postures in rats, suggesting possible changes in maternal behavior induced by Hcrt/Orx in specific brain areas in rodents (193).

Except for the direct regulation of reproductive function by Hcrt/Orx in a sex-dependent manner in male and female animals, any sex dimorphism in the regulation of homeostatic functions and complex behaviors by Hcrt/Orx in animals is not entirely clear because most studies to date were performed in males.

The roles of sex-dependent factors playing in the regulation of nonreproductive functions and the development of diseases/conditions resulting from dysregulated physiological functions (such as energy metabolism) are rather complex in humans (194). For example, although obesity is more frequently diagnosed in women than men, type 2 diabetes is more likely diagnosed at lower ages or body mass index in men than women (194). This complexity is consistent with findings in animal studies that brain pathways controlling energy balance exhibit sex dimorphism. The NPY-mediated effects on energy balance have been revealed to be sex-dependent (195). On the one hand, testosterone stimulates NPY expression and release in the hypothalamic nuclei (such as the median eminence, arcuate nucleus, and VMH) in rodents (196, 197). Male rats exhibit more NPY-containing cells than females in the ARC (197). On the other hand, E2 inhibits food intake. The hypothalamic expression of NPY fluctuates across the different phases of the estrus cycle, reaching its lowest levels during proestrus when the plasma estrogen level peaks (198, 199). The cyclic changes in NPY and E2 levels are consistent with the cyclic changes in food intake and body weight in naturally cycling female animals (198, 199). A recent study has identified a sex difference in the response of NPY/AgRP neurons in the development of insulin resistance in mice with DIO. The efficacy of insulin to activate KATP channels in NPY/AgRP neurons was significantly attenuated in male but not female DIO mice, demonstrating that E2 exerted a protective role against insulin resistance based upon the response of the NPY/AgRP neurons in females (200). Data regarding sex-dependent regulation of energy balance by Hcrt/Orx have not been well established thus far. In addition, the effects of gonadal hormones on sleep regulation mediated by Hcrt/Orx and changes in cognitive performance caused by sleep loss because of a dysregulated Hcrt/Orx system are also poorly defined (188).

A growing body of evidence from clinical and preclinical studies have shown that biological processes underlying addictive behaviors are different between male and female animals (201, 202). In humans, women develop substance use disorders more rapidly after initial use and show faster relapse than men at all ages (203, 204). The subjective effects of psychostimulants on women fluctuates across the menstrual cycle (205), with the greatest effects reported when E2 levels are high (206). Consistently, exogenous E2 enhances the subjective effects of psychostimulants (such as cocaine) in women (207), whereas exogenous progesterone and its metabolite, allopregnanolone, attenuate the subjective effects of cocaine and other drugs of abuse in both men and women (208). In animal studies, female rats acquire cocaine self-administration behavior more rapidly and at lower doses than males (209-211). The acquisition of cocaine self-administration behavior was markedly reduced by ovariectomy and then restored by E2 replacement (212, 213). The effects of E2 on brain functions mediated by dopamine production and signaling in the reward pathway has been well documented (214). For instance, the basal neuronal activity (action potential firing) in vivo was higher in VTA dopamine neurons of female mice during estrus than in males or diestrus females, which causally contributed to the facilitated expression of cocaine-seeking behavior in female animals (215). The exact roles played by the Hcrt/Orx system in mediating the sex dimorphism in reward-seeking and addictive behaviors are not yet clear.

Consistent with sex differences in reward-seeking behaviors, females are twice as likely as males to experience stress-related psychiatric disorders. In adult male and female rats exposed to repeated restraint stress, an increase in Hcrt/Orx expression and activation were observed in female rats compared with males (216). Female rats exhibited an impaired habituation to repeated restraints and subsequent deficits in cognitive functions after stress exposure compared to male rats. The inhibition of Hcrt/Orx using designer receptors exclusively activated by designer drugs abolished the heightened hypothalamic-pituitary-adrenal response and reduced stress-induced cognitive impairments in female rats (216). However, it is a different story during early development. Early life stress, such as neonatal maternal deprivation, induced a more remarkable effect on male than female animals (217). Early life stress dampened the responses of Hcrt cells (decreased activation of c-Fos) to restraint stress in both male and female rats in their adulthood. Voluntary exercise during the late adolescence period reversed the deficiency in the Hcrt/Orx system and behavioral deficits in males but not females (218). The data regarding sex differences in Hcrt/Orx-mediated stress-induced actions in animals continue to emerge.

Concluding Remarks

The topic of sex differences in the regulation of physiological functions and animal behaviors has been a highly focused area of research during the past several decades and is critical to the understanding of basic biological mechanisms and the real-world practice of sex equality in human health care. Contributions to our body of knowledge with regard to the essential roles of the Hcrt/Orx system in mediating the sexual dimorphism in physiological functions, complex behaviors, diseases, and pathological conditions is accelerating. However, many gaps remain. For example, it is still not entirely clear how the Hcrt/Orx system mediates sex-related diverse effects on the brain and peripheral organs in animals and humans even while the Hcrt/Orx cells per se do not directly express estrogen and androgen receptors. It is essential to decipher the connections of the Hcrt/Orx cells and their upstream partners that do express estrogen and androgen receptors and are modulated by sex hormones. It is also unclear how the Hcrt/Orx system is regulated genetically and/or epigenetically given the possible influence of various (sex) hormone milieus that occur over the life span of male and female animals. Ultimately, bridging these gaps in our understanding of sex differences in Hcrt/Orx-mediated functions and behaviors will lead to new diagnostic and therapeutic approaches in the advancement of human health care.

Abbreviations

    Abbreviations
     
  • AgRP

    agouti-related protein

  •  
  • AMPAR

    AMPA-type glutamate receptor

  •  
  • ARC

    arcuate nucleus

  •  
  • Cp-AMPAR

    calcium-permeable AMPA-type glutamate receptor

  •  
  • CRF

    corticotropin-releasing factor

  •  
  • DIO

    diet-induced obese

  •  
  • DMV

    dorsal motor nucleus of the vagus

  •  
  • E2

    estrogen

  •  
  • GABA

    gamma-aminobutyric acid

  •  
  • Hcrt/Orx

    hypocretin/orexin

  •  
  • LepRb

    long form leptin receptor

  •  
  • LHA

    lateral hypothalamic area

  •  
  • LHb

    lateral habenula

  •  
  • MCH

    melanin-concentrating hormone

  •  
  • MCHR1

    melanin-concentrating hormone receptor 1

  •  
  • NMDAR

    NMDA receptor

  •  
  • NPY

    neuropeptide Y

  •  
  • PVN

    paraventricular nucleus

  •  
  • VMH

    ventromedial hypothalamus

  •  
  • VTA

    ventral tegmental area

Acknowledgments

The authors would like to pay tribute to their friend and colleague, Dr Anthony N van den Pol, who passed away last year. He was a trailblazer in neuroscience of the hypothalamus who contributed significantly to our understanding of the Hcrt/Orx system. The authors thank Marya Shanabrough for assistance with the manuscript. The graphic abstract was partially created with BioRender.com.

Funding

The authors’ studies have been supported by National Institutes of Health grants DA046160 (T.L.H. and X.B.G.) and DK120891 (X.B.G. and T.L.H.).

Additional Information

Disclosures: The authors have no conflict of interests related with this article to disclose.

Data Availability

Data sharing is not applicable to this article.

References

1.

Brooks
CM
.
The history of thought concerning the hypothalamus and its functions
.
Brain Res Bull.
1988
;
20
(
6
):
657
-
667
.

2.

Arrigoni
E
,
Chee
MJS
,
Fuller
PM
.
To eat or to sleep: that is a lateral hypothalamic question
.
Neuropharmacology.
2019
;
154
:
34
-
49
.

3.

Abreu
AR
,
Molosh
AI
,
Johnson
PL
,
Shekhar
A
.
Role of medial hypothalamic orexin system in panic, phobia and hypertension
.
Brain Res.
2020
;
1731
:
145942
.

4.

Tyree
SM
,
Borniger
JC
,
de Lecea
L
.
Hypocretin as a hub for arousal and motivation
.
Front Neurol.
2018
;
9
:
413
.

5.

Burdakov
D
,
Peleg-Raibstein
D
.
The hypothalamus as a primary coordinator of memory updating
.
Physiol Behav.
2020
;
223
:
112988
.

6.

Gao
XB
,
Hermes
G
.
Neural plasticity in hypocretin neurons: the basis of hypocretinergic regulation of physiological and behavioral functions in animals
.
Front Syst Neurosci.
2015
;
9
:
142
.

7.

de Lecea
L
,
Kilduff
TS
,
Peyron
C
, et al.
The hypocretins: hypothalamus-specific peptides with neuroexcitatory activity
.
Proc. Natl. Acad. Sci. USA
.
1998
;
95
:
322
-
327
.

8.

Sakurai
T
,
Amemiya
A
,
Ishii
M
, 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
.

9.

Gautvik
KM
,
de Lecea
L
,
Gautvik
VT
, et al.
Overview of the most prevalent hypothalamus-specific mRNAs, as identified by directional tag PCR subtraction
.
Proc Natl Acad Sci U S A.
1996
;
93
(
16
):
8733
-
8738
.

10.

Van Den Pol
AN
,
Patrylo
PR
,
Ghosh
PK
,
Gao
XB
.
Lateral hypothalamus: early developmental expression and response to hypocretin (orexin)
.
J Comp Neurol.
2001
;
433
(
3
):
349
-
363
.

11.

Chou
TC
,
Lee
CE
,
Lu
J
, et al.
Orexin (hypocretin) neurons contain dynorphin
.
J Neurosci
.
2001
;
21
(
19
):
RC168
.

12.

Maolood
N
,
Meister
B
.
Nociceptin/orphanin FQ peptide in hypothalamic neurones associated with the control of feeding behaviour
.
J Neuroendocrinol.
2010
;
22
(
2
):
75
-
82
.

13.

Honda
M
,
Eriksson
KS
,
Zhang
S
, et al.
IGFBP3 colocalizes with and regulates hypocretin (orexin)
.
PloS One
2009
;
4
:
e4254
.

14.

Cvetkovic-Lopes
V
,
Bayer
L
,
Dorsaz
S
, et al.
2010
.
Elevated tribbles homolog 2-specific antibody levels in narcolepsy patients
.
J Clin Invest
2010
;
120
:
713
719
.

15.

Dalal
J
,
Roh
JH
,
Maloney
SE
, et al.
Translational profiling of hypocretin neurons identifies candidate molecules for sleep regulation
.
Genes Dev.
2013
;
27
:
565
578
.

16.

Yelin-Bekerman
L
,
Elbaz
I
,
Diber
A
, et al.
Hypocretin neuron-specific transcriptome profiling identifies the sleep modulator Kcnh4a
.
Elife.
2015
;
4
:
e08638
.

17.

Mickelsen
LE
,
Bolisetty
M
,
Chimileski
BR
, et al.
Single-cell transcriptomic analysis of the lateral hypothalamic area reveals molecularly distinct populations of inhibitory and excitatory neurons
.
Nat Neurosci.
2019
;
22
(
4
):
642
-
656
.

18.

Seifinejad
A
,
Li
S
,
Mikhail
C
, et al.
Molecular codes and in vitro generation of hypocretin and melanin concentrating hormone neurons
.
Proc Natl Acad Sci U S A.
2019
;
116
(
34
):
17061
-
17070
.

19.

Peyron
C
,
Tighe
DK
,
van den Pol
AN
, et al.
Neurons containing hypocretin (orexin) project to multiple neuronal systems
.
J Neurosci.
1998
;
18
(
23
):
9996
-
10015
.

20.

Thannickal
TC
,
Moore
RY
,
Nienhuis
R
, et al.
Reduced number of hypocretin neurons in human narcolepsy
.
Neuron.
2000
;
27
(
3
):
469
-
474
.

21.

Sutcliffe
JG
,
de Lecea
L
.
The hypocretins: setting the arousal threshold
.
Nat Rev Neurosci.
2002
;
3
(
5
):
339
-
349
.

22.

Brownell
SE
,
Conti
B
.
Age- and gender-specific changes of hypocretin immunopositive neurons in C57Bl/6 mice
.
Neurosci Lett.
2010
;
472
(
1
):
29
-
32
.

23.

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
(
1-2
):
71
-
75
.

24.

Date
Y
,
Ueta
Y
,
Yamashita
H
, et al.
Orexins, orexigenic hypothalamic peptides, interact with autonomic, neuroendocrine and neuroregulatory systems
.
Proc Natl Acad Sci U S A.
1999
;
96
(
2
):
748
-
753
.

25.

van den Pol
AN
.
Hypothalamic hypocretin (orexin): robust innervation of the spinal cord
.
J Neurosci.
1999
;
19
(
8
):
3171
-
3182
.

26.

Marcus
JN
,
Aschkenasi
CJ
,
Lee
CE
, et al.
Differential expression of orexin receptors 1 and 2 in the rat brain
.
J Comp Neurol.
2001
;
435
(
1
):
6
-
25
.

27.

Adeghate
E
,
Fernandez-Cabezudo
M
,
Hameed
R
, et al.
Orexin-1 receptor co-localizes with pancreatic hormones in islet cells and modulates the outcome of streptozotocin-induced diabetes mellitus
.
Plos One.
2010
;
5
(
1
):
e8587
.

28.

Rosin
DL
,
Weston
MC
,
Sevigny
CP
,
Stornetta
RL
,
Guyenet
PG
.
Hypothalamic orexin (hypocretin) neurons express vesicular glutamate transporters VGLUT1 or VGLUT2
.
J Comp Neurol.
2003
;
465
(
4
):
593
-
603
.

29.

Torrealba
F
,
Yanagisawa
M
,
Saper
CB
.
Colocalization of orexin a and glutamate immunoreactivity in axon terminals in the tuberomammillary nucleus in rats
.
Neuroscience.
2003
;
119
(
4
):
1033
-
1044
.

30.

Mickelsen
LE
,
Kolling
FE
,
Chimileski
BR
, et al.
Neurochemical heterogeneity among lateral hypothalamic hypocretin/orexin and melanin-concentrating hormone neurons identified through single cell gene expression analysis
.
eNeuro
2017
;
4
:
1
24
.

31.

Vakilian
H
,
Mirzaei
M
,
Sharifi Tabar
M
, et al.
DDX3Y, a male-specific region of Y chromosome gene, may modulate neuronal differentiation
.
J Proteome Res.
2015
;
14
(
9
):
3474
-
3483
.

32.

Kay
GF
,
Barton
SC
,
Surani
MA
,
Rastan
S
.
Imprinting and X chromosome counting mechanisms determine Xist expression in early mouse development
.
Cell.
1994
;
77
(
5
):
639
-
650
.

33.

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
(
19
):
7962
-
7971
.

34.

Hagan
JJ
,
Leslie
RA
,
Patel
S
, 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
-
6
.

35.

Davis
SF
,
Williams
KW
,
Xu
W
,
Glatzer
NR
,
Smith
BN
.
Selective enhancement of synaptic inhibition by hypocretin (orexin) in rat vagal motor neurons: implications for autonomic regulation
.
J Neurosci.
2003
;
23
(
9
):
3844
-
3854
.

36.

Korotkova
TM
,
Sergeeva
OA
,
Eriksson
KS
,
Haas
HL
,
Brown
RE
.
Excitation of ventral tegmental area dopaminergic and nondopaminergic neurons by orexins/hypocretins
.
J Neurosci.
2003
;
23
(
1
):
7
-
11
.

37.

Borgland
SL
,
Taha
SA
,
Sarti
F
,
Fields
HL
,
Bonci
A
.
Orexin A in the VTA is critical for the induction of synaptic plasticity and behavioral sensitization to cocaine
.
Neuron.
2006
;
49
(
4
):
589
-
601
.

38.

Kukkonen
JP
,
Turunen
PM
.
Cellular signaling mechanisms of hypocretin/orexin.
In
Steiner
MA
,
Yanagisawa
M
,
Clozel
M
(eds):
The Orexin System. Basic Science and Role in Sleep Pathology. Front Neurol Neurosci
.
Basel
:
Karger
;
2021
;
45
:
91
102
.

39.

Williams
RH
,
Alexopoulos
H
,
Jensen
LT
,
Fugger
L
,
Burdakov
D
.
Adaptive sugar sensors in hypothalamic feeding circuits
.
Proc Natl Acad Sci U S A.
2008
;
105
(
33
):
11975
-
11980
.

40.

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
(
17
):
4157
-
4166
.

41.

Burdakov
D
,
Jensen
LT
,
Alexopoulos
H
, et al.
Tandem-pore K+ channels mediate inhibition of orexin neurons by glucose
.
Neuron.
2006
;
50
(
5
):
711
-
722
.

42.

Guyon
A
,
Tardy
MP
,
Rovère
C
,
Nahon
JL
,
Barhanin
J
,
Lesage
F
.
Glucose inhibition persists in hypothalamic neurons lacking tandem-pore K+ channels
.
J Neurosci.
2009
;
29
(
8
):
2528
-
2533
.

43.

Parsons
MP
,
Hirasawa
M
.
ATP-sensitive potassium channel-mediated lactate effect on orexin neurons: implications for brain energetics during arousal
.
J Neurosci.
2010
;
30
(
24
):
8061
-
8070
.

44.

Karnani
MM
,
Apergis-Schoute
J
,
Adamantidis
A
, et al.
Activation of central orexin/hypocretin neurons by dietary amino acids
.
Neuron.
2011
;
72
(
4
):
616
-
629
.

45.

Horvath
TL
,
Gao
XB
.
Input organization and plasticity of hypocretin neurons: possible clues to obesity’s association with insomnia
.
Cell Metab.
2005
;
1
(
4
):
279
-
286
.

46.

Xie
X
,
Crowder
TL
,
Yamanaka
A
, et al.
GABA(B) receptor-mediated modulation of hypocretin/orexin neurones in mouse hypothalamus
.
J Physiol.
2006
;
574
(
Pt 2
):
399
-
414
.

47.

Douglas
R
,
Markram
H
,
Martin
K
.
Neocortex.
In
The Synaptic Organization of the Brain
.
G.M.
Shepherd
, ed.
New York
:
Oxford University Press
;
2004
;
499
-
558
.

48.

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
(
6
):
1169
-
1181
.

49.

Tan
Y
,
Hang
F
,
Liu
ZW
, et al.
Impaired hypocretin/orexin system alters responses to salient stimuli in obese male mice
.
J Clin Invest
2020
;
130
:
4985
-
4998
.

50.

Perkel
DJ
,
Nicoll
RA
.
Evidence for all-or-none regulation of neurotransmitter release: implications for long-term potentiation
.
J Physiol.
1993
;
471
:
481
-
500
.

51.

Isaac
JT
,
Nicoll
RA
,
Malenka
RC
.
Evidence for silent synapses: implications for the expression of LTP
.
Neuron.
1995
;
15
(
2
):
427
-
434
.

52.

Rao
Y
,
Liu
ZW
,
Borok
E
, et al.
Prolonged wakefulness induces experience-dependent synaptic plasticity in mouse hypocretin/orexin neurons
.
J Clin Invest.
2007
;
117
(
12
):
4022
-
4033
.

53.

Rao
Y
,
Lu
M
,
Ge
F
, et al.
Regulation of synaptic efficacy in hypocretin/orexin-containing neurons by melanin concentrating hormone in the lateral hypothalamus
.
J Neurosci.
2008
;
28
(
37
):
9101
-
9110
.

54.

Rao
Y
,
Mineur
YS
,
Gan
G
, et al.
Repeated in vivo exposure of cocaine induces long-lasting synaptic plasticity in hypocretin/orexin-producing neurons in the lateral hypothalamus in mice
.
J Physiol.
2013
;
591
(
7
):
1951
-
1966
.

55.

Ungless
MA
,
Whistler
JL
,
Malenka
RC
,
Bonci
A
.
Single cocaine exposure in vivo induces long-term potentiation in dopamine neurons
.
Nature.
2001
;
411
(
6837
):
583
-
587
.

56.

Glasgow
NG
,
Siegler Retchless
B
,
Johnson
JW
.
Molecular bases of NMDA receptor subtype-dependent properties
.
J Physiol.
2015
;
593
(
1
):
83
-
95
.

57.

Plant
K
,
Pelkey
KA
,
Bortolotto
ZA
, et al.
Transient incorporation of native GluR2-lacking AMPA receptors during hippocampal long-term potentiation
.
Nat Neurosci.
2006
;
9
(
5
):
602
-
604
.

58.

Mahanty
NK
,
Sah
P
.
Calcium-permeable AMPA receptors mediate long-term potentiation in interneurons in the amygdala
.
Nature.
1998
;
394
(
6694
):
683
-
687
.

59.

Guan
JL
,
Uehara
K
,
Lu
S
, et al.
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.
2002
;
26
(
12
):
1523
-
1532
.

60.

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.
2004
;
42
(
4
):
635
-
652
.

61.

Chee
MJ
,
Pissios
P
,
Maratos-Flier
E
.
Neurochemical characterization of neurons expressing melanin-concentrating hormone receptor 1 in the mouse hypothalamus
.
J Comp Neurol.
2013
;
521
(
10
):
2208
-
2234
.

62.

Yamanaka
A
,
Beuckmann
CT
,
Willie
JT
, et al.
Hypothalamic orexin neurons regulate arousal according to energy balance in mice
.
Neuron.
2003
;
38
(
5
):
701
-
713
.

63.

Leinninger
GM
,
Jo
YH
,
Leshan
RL
, et al.
Leptin acts via leptin receptor-expressing lateral hypothalamic neurons to modulate the mesolimbic dopamine system and suppress feeding
.
Cell Metab.
2009
;
10
(
2
):
89
-
98
.

64.

Louis
GW
,
Leinninger
GM
,
Rhodes
CJ
,
Myers
MG
Jr
.
Direct innervation and modulation of orexin neurons by lateral hypothalamic LepRb neurons
.
J Neurosci.
2010
;
30
(
34
):
11278
-
11287
.

65.

Ferrari
LL
,
Park
D
,
Zhu
L
,
Palmer
MR
,
Broadhurst
RY
,
Arrigoni
E
.
Regulation of lateral hypothalamic orexin activity by local GABAergic neurons
.
J Neurosci.
2018
;
38
(
6
):
1588
-
1599
.

66.

Apergis-Schoute
J
,
Iordanidou
P
,
Faure
C
, et al.
Optogenetic evidence for inhibitory signaling from orexin to MCH neurons via local microcircuits
.
J Neurosci.
2015
;
35
(
14
):
5435
-
5441
.

67.

Tenorio-Lopes
L
,
Fournier
S
,
Henry
MS
,
Bretzner
F
,
Kinkead
R
.
Disruption of estradiol regulation of orexin neurons: a novel mechanism in excessive ventilatory response to CO2 inhalation in a female rat model of panic disorder
.
Transl Psychiatry.
2020
;
10
(
1
):
394
.

68.

Muschamp
JW
,
Dominguez
JM
,
Sato
SM
,
Shen
RY
,
Hull
EM
.
A role for hypocretin (orexin) in male sexual behavior
.
J Neurosci.
2007
;
27
(
11
):
2837
-
2845
.

69.

Hetherington
AW
,
Ranson
SW
.
Hypothalamic lesions and adiposity in the rat
.
Anat Rec.
1940
;
78
:
149
172
.

70.

Anand
BK
,
Brobeck
JR
.
Hypothalamic control of food intake in rats and cats
.
Yale J Biol Med.
1951
;
24
(
2
):
123
-
140
.

71.

Anand
BK
,
Brobeck
JR
.
Localization of a “feeding center” in the hypothalamus of the rat
.
Proc Soc Exp Biol Med.
1951
;
77
(
2
):
323
-
324
.

72.

Stellar
E
.
The physiology of motivation
.
Psychol Rev.
1954
;
61
(
1
):
5
-
22
.

73.

Sawchenko
PE
.
Toward a new neurobiology of energy balance, appetite, and obesity: the anatomists weigh in
.
J Comp Neurol.
1998
;
402
(
4
):
435
-
441
.

74.

Delgado
JM
,
Anand
BK
.
Increase of food intake induced by electrical stimulation of the lateral hypothalamus
.
Am J Physiol.
1953
;
172
(
1
):
162
-
168
.

75.

Ungerstedt
U
.
Adipsia and aphagia after 6-hydroxydopamine induced degeneration of the nigro-striatal dopamine system
.
Acta Physiol Scand Suppl.
1971
;
367
:
95
-
122
.

76.

Stricker
EM
,
Swerdloff
AF
,
Zigmond
MJ
.
Intrahypothalamic injections of kainic acid produce feeding and drinking deficits in rats
.
Brain Res.
1978
;
158
(
2
):
470
-
473
.

77.

Katafuchi
T
,
Oomura
Y
,
Yoshimatsu
H
.
Single neuron activity in the rat lateral hypothalamus during 2-deoxy-D-glucose induced and natural feeding behavior
.
Brain Res.
1985
;
359
(
1-2
):
1
-
9
.

78.

Himmi
T
,
Boyer
A
,
Orsini
JC
.
Changes in lateral hypothalamic neuronal activity accompanying hyper- and hypoglycemias
.
Physiol Behav.
1988
;
44
(
3
):
347
-
354
.

79.

Dube
MG
,
Kalra
SP
,
Kalra
PS
.
Food intake elicited by central administration of orexins/hypocretins: identification of hypothalamic sites of action
.
Brain Res.
1999
;
842
(
2
):
473
-
477
.

80.

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.
2003
;
284
(
6
):
R1409
-
R1417
.

81.

Haynes
AC
,
Jackson
B
,
Chapman
H
, et al.
A selective orexin-1 receptor antagonist reduces food consumption in male and female rats
.
Regul Pept.
2000
;
96
(
1-2
):
45
-
51
.

82.

Rodgers
RJ
,
Halford
JC
,
Nunes de Souza
RL
, et al.
SB-334867, a selective orexin-1 receptor antagonist, enhances behavioural satiety and blocks the hyperphagic effect of orexin-A in rats
.
Eur J Neurosci.
2001
;
13
(
7
):
1444
-
1452
.

83.

Baldo
BA
,
Gual-Bonilla
L
,
Sijapati
K
,
Daniel
RA
,
Landry
CF
,
Kelley
AE
.
Activation of a subpopulation of orexin/hypocretin-containing hypothalamic neurons by GABAA receptor-mediated inhibition of the nucleus accumbens shell, but not by exposure to a novel environment
.
Eur J Neurosci.
2004
;
19
(
2
):
376
-
386
.

84.

Cai
XJ
,
Widdowson
PS
,
Harrold
J
, et al.
Hypothalamic orexin expression: modulation by blood glucose and feeding
.
Diabetes.
1999
;
48
(
11
):
2132
-
2137
.

85.

Griffond
B
,
Risold
PY
,
Jacquemard
C
,
Colard
C
,
Fellman
D
.
Insulin-induced hypoglycemia increases prehypocretin (orexin) mRNA in the rat lateral hypothalamic area
.
Neurosci. Lett
.
1999
;
262
:
77
80
.

86.

Moriguchi
T
,
Sakurai
T
,
Nambu
T
,
Yanagisawa
M
,
Goto
K
.
Neurons containing orexin in the lateral hypothalamic area of the adult rat brain are activated by insulin-induced acute hypoglycemia
.
Neurosci Lett.
1999
;
264
(
1-3
):
101
-
104
.

87.

Mondal
MS
,
Nakazato
M
,
Date
Y
,
Murakami
N
,
Yanagisawa
M
,
Matsukura
S
.
Widespread distribution of orexin in rat brain and its regulation upon fasting
.
Biochem Biophys Res Commun.
1999
;
256
(
3
):
495
-
499
.

88.

Yamamoto
Y
,
Ueta
Y
,
Serino
R
,
Nomura
M
,
Shibuya
I
,
Yamashita
H
.
Effects of food restriction on the hypothalamic prepro-orexin gene expression in genetically obese mice
.
Brain Res Bull.
2000
;
51
(
6
):
515
-
521
.

89.

Diano
S
,
Horvath
B
,
Urbanski
HF
,
Sotonyi
P
,
Horvath
TL
.
Fasting activates the nonhuman primate hypocretin (orexin) system and its postsynaptic targets
.
Endocrinology.
2003
;
144
(
9
):
3774
-
3778
.

90.

Beck
B
,
Richy
S
,
Dimitrov
T
,
Stricker-Krongrad
A
.
Opposite regulation of hypothalamic orexin and neuropeptide Y receptors and peptide expressions in obese Zucker rats
.
Biochem Biophys Res Commun.
2001
;
286
(
3
):
518
-
523
.

91.

López
M
,
Seoane
L
,
García
MC
, et al.
Leptin regulation of prepro-orexin and orexin receptor mRNA levels in the hypothalamus
.
Biochem Biophys Res Commun.
2000
;
269
(
1
):
41
-
45
.

92.

Takahashi
N
,
Okumura
T
,
Yamada
H
,
Kohgo
Y
.
Stimulation of gastric acid secretion by centrally administered orexin-A in conscious rats
.
Biochem Biophys Res Commun.
1999
;
254
(
3
):
623
-
627
.

93.

Krowicki
ZK
,
Burmeister
MA
,
Berthoud
HR
,
Scullion
RT
,
Fuchs
K
,
Hornby
PJ
.
Orexins in rat dorsal motor nucleus of the vagus potently stimulate gastric motor function
.
Am J Physiol Gastrointest Liver Physiol.
2002
;
283
(
2
):
G465
-
G472
.

94.

Grabauskas
G
,
Moises
HC
.
Gastrointestinal-projecting neurones in the dorsal motor nucleus of the vagus exhibit direct and viscerotopically organized sensitivity to orexin
.
J Physiol.
2003
;
549
(
Pt 1
):
37
-
56
.

95.

Cai
XJ
,
Evans
ML
,
Lister
CA
, et al.
Hypoglycemia activates orexin neurons and selectively increases hypothalamic orexin-B levels: responses inhibited by feeding and possibly mediated by the nucleus of the solitary tract
.
Diabetes.
2001
;
50
(
1
):
105
-
112
.

96.

Squecco
R
,
Garella
R
,
Luciani
G
,
Francini
F
,
Baccari
MC
.
Muscular effects of orexin A on the mouse duodenum: mechanical and electrophysiological studies
.
J Physiol.
2011
;
589
(
Pt 21
):
5231
-
5246
.

97.

Semjonous
NM
,
Smith
KL
,
Parkinson
JR
, et al.
Coordinated changes in energy intake and expenditure following hypothalamic administration of neuropeptides involved in energy balance
.
Int J Obes (Lond).
2009
;
33
(
7
):
775
-
785
.

98.

Gao
XB
.
The role of melanin-concentrating hormone in the regulation of the sleep/wake cycle: sleep promoter or arousal modulator?
In
SR
Pandi-Perumal
,
Pablo
Torterolo
,
Jaime M.
Monti
, editors:
Melanin-Concentrating Hormone and Sleep-Molecular, Functional and Clinical Aspects
.
Switzerland
:
Springer
,
2018
:
57
-
74
.

99.

Lubkin
M
,
Stricker-Krongrad
A
.
Independent feeding and metabolic actions of orexins in mice
.
Biochem Biophys Res Commun.
1998
;
253
(
2
):
241
-
245
.

100.

Asakawa
A
,
Inui
A
,
Goto
K
, et al.
Effects of agouti-related protein, orexin and melanin-concentrating hormone on oxygen consumption in mice
.
Int J Mol Med.
2002
;
10
(
4
):
523
-
525
.

101.

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
(
1
):
128
-
134
.

102.

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
(
4
):
E551
-
E559
.

103.

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
(
1-3
):
27
-
32
.

104.

Kotz
CM
,
Teske
JA
,
Billington
CJ
.
Neuroregulation of nonexercise activity thermogenesis and obesity resistance
.
Am J Physiol Regul Integr Comp Physiol.
2008
;
294
(
3
):
R699
-
R710
.

105.

Kotz
CM
,
Wang
C
,
Teske
JA
, et al.
Orexin A mediation of time spent moving in rats: neural mechanisms
.
Neuroscience.
2006
;
142
(
1
):
29
-
36
.

106.

Shiuchi
T
,
Haque
MS
,
Okamoto
S
, et al.
Hypothalamic orexin stimulates feeding-associated glucose utilization in skeletal muscle via sympathetic nervous system
.
Cell Metab.
2009
;
10
(
6
):
466
-
480
.

107.

Murakami
M
,
Ohba
T
,
Kushikata
T
, et al.
Involvement of the orexin system in sympathetic nerve regulation
.
Biochem Biophys Res Commun.
2015
;
460
(
4
):
1076
-
1081
.

108.

Sellayah
D
,
Bharaj
P
,
Sikder
D
.
Orexin is required for brown adipose tissue development, differentiation, and function
.
Cell Metab.
2011
;
14
(
4
):
478
-
490
.

109.

Folgueira
C
,
Beiroa
D
,
Porteiro
B
, et al.
Hypothalamic dopamine signaling regulates brown fat thermogenesis
.
Nat Metab
.
2019
;
1
(
8
):
811
-
829
.

110.

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
(
44
):
15944
-
15955
.

111.

Sellayah
D
,
Sikder
D
.
Orexin receptor-1 mediates brown fat developmental differentiation
.
Adipocyte.
2012
;
1
(
1
):
58
-
63
.

112.

von Economo
C
.
Sleep as a problem of localization
.
J. Nerv. Ment. Dis.
71
:
249
259
.

113.

Ranson
SW
.
Somnolence caused by hypothalamic lesions in the monkey
.
Arch. Neurol. Psychiatry
1939
;
41
:
1
23
.

114.

Nauta
WJH
.
Hypothalamic regulation of sleep in rats: an experimental study
.
J Neurophysiol
1946
;
9
:
285
316
.

115.

Swett
CP
,
Hobson
JA
.
The effects of posterior hypothalamic lesions on behavioral and electrographic manifestations of sleep and waking in cats
.
Arch Ital Biol.
1968
;
106
(
3
):
283
-
293
.

116.

Lin
JS
,
Sakai
K
,
Vanni-Mercier
G
,
Jouvet
M
.
A critical role of the posterior hypothalamus in the mechanisms of wakefulness determined by microinjection of muscimol in freely moving cats
.
Brain Res.
1989
;
479
(
2
):
225
-
240
.

117.

de Lecea
L
,
Sutcliffe
JG
.
The hypocretins and sleep
.
Febs J.
2005
;
272
(
22
):
5675
-
5688
.

118.

Sakurai
T
.
Roles of orexin/hypocretin in regulation of sleep/wakefulness and energy homeostasis
.
Sleep Med Rev.
2005
;
9
(
4
):
231
-
241
.

119.

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
(
5
):
993
-
997
.

120.

Yoshida
Y
,
Fujiki
N
,
Nakajima
T
, et al.
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
(
7
):
1075
-
1081
.

121.

Zeitzer
JM
,
Buckmaster
CL
,
Parker
KJ
,
Hauck
CM
,
Lyons
DM
,
Mignot
E
.
Circadian and homeostatic regulation of hypocretin in a primate model: implications for the consolidation of wakefulness
.
J Neurosci.
2003
;
23
(
8
):
3555
-
3560
.

122.

Desarnaud
F
,
Murillo-Rodriguez
E
,
Lin
L
, et al.
The diurnal rhythm of hypocretin in young and old F344 rats
.
Sleep.
2004
;
27
(
5
):
851
-
856
.

123.

Estabrooke
IV
,
McCarthy
MT
,
Ko
E
, et al.
Fos expression in orexin neurons varies with behavioral state
.
J Neurosci.
2001
;
21
(
5
):
1656
-
1662
.

124.

Lee
MG
,
Hassani
OK
,
Jones
BE
.
Discharge of identified orexin/hypocretin neurons across the sleep-waking cycle
.
J Neurosci.
2005
;
25
(
28
):
6716
-
6720
.

125.

Mileykovskiy
BY
,
Kiyashchenko
LI
,
Siegel
JM
.
Behavioral correlates of activity in identified hypocretin/orexin neurons
.
Neuron.
2005
;
46
(
5
):
787
-
798
.

126.

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
(
7168
):
420
-
424
.

127.

Takahashi
S
,
Tominaga
M
,
Yamanaka
A
.
Acute optogenetic silencing of orexin/hypocretin neurons induces slow-wave sleep in mice
.
J Neurosci
.
2011
;
31
(
29
):
10529
-
39
.

128.

Li
SB
,
Nevárez
N
,
Giardino
WJ
,
de Lecea
L
.
Optical probing of orexin/hypocretin receptor antagonists
.
Sleep.
2018
;
41
(
10
):
zsy141
.

129.

Bourgin
P
,
Huitrón-Résendiz
S
,
Spier
AD
, et al.
Hypocretin-1 modulates rapid eye movement sleep through activation of locus coeruleus neurons
.
J Neurosci.
2000
;
20
(
20
):
7760
-
7765
.

130.

España
RA
,
Baldo
BA
,
Kelley
AE
,
Berridge
CW
.
Wake-promoting and sleep-suppressing actions of hypocretin (orexin): basal forebrain sites of action
.
Neuroscience.
2001
;
106
(
4
):
699
-
715
.

131.

van den Pol
AN
,
Ghosh
PK
,
Liu
RJ
,
Li
Y
,
Aghajanian
GK
,
Gao
XB
.
Hypocretin (orexin) enhances neuron activity and cell synchrony in developing mouse GFP-expressing locus coeruleus
.
J Physiol.
2002
;
541
(
Pt 1
):
169
-
185
.

132.

Yoshida
K
,
McCormack
S
,
España
RA
,
Crocker
A
,
Scammell
TE
.
Afferents to the orexin neurons of the rat brain
.
J Comp Neurol.
2006
;
494
(
5
):
845
-
861
.

133.

Feng
H
,
Wen
SY
,
Qiao
QC
, et al.
Orexin signaling modulates synchronized excitation in the sublaterodorsal tegmental nucleus to stabilize REM sleep
.
Nat Commun.
2020
;
11
(
1
):
3661
.

134.

Chemelli
RM
,
Willie
JT
,
Sinton
CM
, et al.
Narcolepsy in orexin knockout mice: molecular genetics of sleep regulation
.
Cell
.
1999
;
98
:
437
-
451
.

135.

Lin
L
,
Faraco
J
,
Li
R
, et al.
The sleep disorder canine narcolepsy is caused by a mutation in the hypocretin (orexin) receptor 2 gene
.
Cell.
1999
;
98
(
3
):
365
-
376
.

136.

Nishino
S
,
Ripley
B
,
Overeem
S
,
Lammers
GJ
,
Mignot
E
.
Hypocretin (orexin) deficiency in human narcolepsy
.
Lancet.
2000
;
355
(
9197
):
39
-
40
.

137.

Ripley
B
,
Overeem
S
,
Fujiki
N
, et al.
CSF hypocretin/orexin levels in narcolepsy and other neurological conditions
.
Neurology
2001
;
57
:
2253
-
2258
.

138.

Prober
DA
,
Rihel
J
,
Onah
AA
,
Sung
RJ
,
Schier
AF
.
Hypocretin/orexin overexpression induces an insomnia-like phenotype in zebrafish
.
J Neurosci.
2006
;
26
(
51
):
13400
-
13410
.

139.

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.
2005
;
21
(
10
):
2807
-
2816
.

140.

González
JA
,
Jensen
LT
,
Iordanidou
P
,
Strom
M
,
Fugger
L
,
Burdakov
D
.
Inhibitory interplay between orexin neurons and eating
.
Curr Biol.
2016
;
26
(
18
):
2486
-
2491
.

141.

Blouin
AM
,
Fried
I
,
Wilson
CL
, et al.
Human hypocretin and melanin-concentrating hormone levels are linked to emotion and social interaction
.
Nat Commun.
2013
;
4
:
1547
.

142.

Karnani
MM
,
Schöne
C
,
Bracey
EF
, et al.
Role of spontaneous and sensory orexin network dynamics in rapid locomotion initiation
.
Prog Neurobiol.
2020
;
187
:
101771
.

143.

Elbaz
I
,
Zada
D
,
Tovin
A
, et al.
Sleep-dependent structural synaptic plasticity of inhibitory synapses in the dendrites of hypocretin/orexin neurons
.
Mol Neurobiol.
2017
;
54
(
8
):
6581
-
6597
.

144.

Zhang
S
,
Zeitzer
JM
,
Yoshida
Y
, et al.
Lesions of the suprachiasmatic nucleus eliminate the daily rhythm of hypocretin-1 release
.
Sleep.
2004
;
27
(
4
):
619
-
627
.

145.

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.
2005
;
130
(
1
):
165
-
183
.

146.

Marston
OJ
,
Williams
RH
,
Canal
MM
,
Samuels
RE
,
Upton
N
,
Piggins
HD
.
Circadian and dark-pulse activation of orexin/hypocretin neurons
.
Mol Brain.
2008
;
1
:
19
.

147.

Olds
J
,
Milner
P
.
Positive reinforcement produced by electrical stimulation of septal area and other regions of rat brain
.
J Comp Physiol Psychol.
1954
;
47
(
6
):
419
-
427
.

148.

Olds
J
.
Self-stimulation of the brain
.
Science
1958
;
127
:
315
324
.

149.

Adams
WJ
,
Lorens
SA
,
Mitchell
CL
.
Morphine enhances lateral hypothalamic self-stimulation in the rat
.
Proc Soc Exp Biol Med.
1972
;
140
(
3
):
770
-
771
.

150.

Goodall
EB
,
Carey
RJ
.
Effects of d- versus l-amphetamine, food deprivation, and current intensity on self-stimulation of the lateral hypothalamus, substantia nigra, and medial frontal cortex of the rat
.
J Comp Physiol Psychol.
1975
;
89
(
9
):
1029
-
1045
.

151.

Olds
ME
,
Williams
KN
.
Self-administration of D-Ala2-Met-enkephalinamide at hypothalamic self-stimulation sites
.
Brain Res.
1980
;
194
(
1
):
155
-
170
.

152.

Cazala
P
,
Darracq
C
,
Saint-Marc
M
.
Self-administration of morphine into the lateral hypothalamus in the mouse
.
Brain Res.
1987
;
416
(
2
):
283
-
288
.

153.

Georgescu
D
,
Zachariou
V
,
Barrot
M
, et al.
Involvement of the lateral hypothalamic peptide orexin in morphine dependence and withdrawal
.
J Neurosci.
2003
;
23
(
8
):
3106
-
3111
.

154.

Harris
GC
,
Wimmer
M
,
Aston-Jones
G
.
A role for lateral hypothalamic orexin neurons in reward seeking
.
Nature.
2005
;
437
(
7058
):
556
-
559
.

155.

Pasumarthi
RK
,
Reznikov
LR
,
Fadel
J
.
Activation of orexin neurons by acute nicotine
.
Eur J Pharmacol.
2006
;
535
(
1-3
):
172
-
176
.

156.

McPherson
CS
,
Featherby
T
,
Krstew
E
,
Lawrence
AJ
.
Quantification of phosphorylated cAMP-response element-binding protein expression throughout the brain of amphetamine-sensitized rats: activation of hypothalamic orexin A-containing neurons
.
J Pharmacol Exp Ther.
2007
;
323
(
3
):
805
-
812
.

157.

Plaza-Zabala
A
,
Flores
Á
,
Maldonado
R
,
Berrendero
F
.
Hypocretin/orexin signaling in the hypothalamic paraventricular nucleus is essential for the expression of nicotine withdrawal
.
Biol Psychiatry.
2012
;
71
(
3
):
214
-
223
.

158.

Boutrel
B
,
Kenny
PJ
,
Specio
SE
, et al.
Role for hypocretin in mediating stress-induced reinstatement of cocaine-seeking behavior
.
Proc Natl Acad Sci U S A.
2005
;
102
(
52
):
19168
-
19173
.

159.

Hamlin
AS
,
Clemens
KJ
,
McNally
GP
.
Renewal of extinguished cocaine-seeking
.
Neuroscience.
2008
;
151
(
3
):
659
-
670
.

160.

España
RA
,
Melchior
JR
,
Roberts
DC
,
Jones
SR
.
Hypocretin 1/orexin A in the ventral tegmental area enhances dopamine responses to cocaine and promotes cocaine self-administration
.
Psychopharmacology (Berl).
2011
;
214
(
2
):
415
-
426
.

161.

Borgland
SL
,
Chang
SJ
,
Bowers
MS
, et al.
Orexin A/hypocretin-1 selectively promotes motivation for positive reinforcers
.
J Neurosci.
2009
;
29
(
36
):
11215
-
11225
.

162.

Hollander
JA
,
Lu
Q
,
Cameron
MD
,
Kamenecka
TM
,
Kenny
PJ
.
Insular hypocretin transmission regulates nicotine reward
.
Proc Natl Acad Sci U S A.
2008
;
105
(
49
):
19480
-
19485
.

163.

LeSage
MG
,
Perry
JL
,
Kotz
CM
,
Shelley
D
,
Corrigall
WA
.
Nicotine self-administration in the rat: effects of hypocretin antagonists and changes in hypocretin mRNA
.
Psychopharmacology (Berl).
2010
;
209
(
2
):
203
-
212
.

164.

Guilleminault
C
,
Carskadon
M
,
Dement
WC
.
On the treatment of rapid eye movement narcolepsy
.
Arch Neurol.
1974
;
30
(
1
):
90
-
93
.

165.

Pasumarthi
RK
,
Fadel
J
.
Activation of orexin/hypocretin projections to basal forebrain and paraventricular thalamus by acute nicotine
.
Brain Res Bull.
2008
;
77
(
6
):
367
-
373
.

166.

Quarta
D
,
Valerio
E
,
Hutcheson
DM
,
Hedou
G
,
Heidbreder
C
.
The orexin-1 receptor antagonist SB-334867 reduces amphetamine-evoked dopamine outflow in the shell of the nucleus accumbens and decreases the expression of amphetamine sensitization
.
Neurochem Int.
2010
;
56
(
1
):
11
-
15
.

167.

Ubaldi
M
,
Giordano
A
,
Severi
I
, et al.
Activation of hypocretin-1/orexin-a neurons projecting to the bed nucleus of the stria terminalis and paraventricular nucleus is critical for reinstatement of alcohol seeking by neuropeptide S
.
Biol Psychiatry.
2016
;
79
(
6
):
452
-
462
.

168.

Matzeu
A
,
Kerr
TM
,
Weiss
F
,
Martin-Fardon
R
.
Orexin-A/hypocretin-1 mediates cocaine-seeking behavior in the posterior paraventricular nucleus of the thalamus via orexin/hypocretin receptor-2
.
J Pharmacol Exp Ther.
2016
;
359
(
2
):
273
-
279
.

169.

Schmeichel
BE
,
Herman
MA
,
Roberto
M
,
Koob
GF
.
Hypocretin neurotransmission within the central amygdala mediates escalated cocaine self-administration and stress-induced reinstatement in rats
.
Biol Psychiatry.
2017
;
81
(
7
):
606
-
615
.

170.

Yeoh
JW
,
James
MH
,
Jobling
P
,
Bains
JS
,
Graham
BA
,
Dayas
CV
.
Cocaine potentiates excitatory drive in the perifornical/lateral hypothalamus
.
J Physiol.
2012
;
590
(
16
):
3677
-
3689
.

171.

Flanigan
ME
,
Aleyasin
H
,
Li
L
, et al.
Orexin signaling in GABAergic lateral habenula neurons modulates aggressive behavior in male mice
.
Nat Neurosci
.
2020
;
23
(
5
):
638
-
650
.

172.

Gammie
SC
,
Auger
AP
,
Jessen
HM
,
Vanzo
RJ
,
Awad
TA
,
Stevenson
SA
.
Altered gene expression in mice selected for high maternal aggression
.
Genes Brain Behav.
2007
;
6
(
5
):
432
-
443
.

173.

Grafe
LA
,
Bhatnagar
S
.
The contribution of orexins to sex differences in the stress response
.
Brain Res.
2020
;
1731
:
145893
.

174.

Martins
PJ
,
D’Almeida
V
,
Pedrazzoli
M
,
Lin
L
,
Mignot
E
,
Tufik
S
.
Increased hypocretin-1 (orexin-a) levels in cerebrospinal fluid of rats after short-term forced activity
.
Regul Pept.
2004
;
117
(
3
):
155
-
158
.

175.

Furlong
TM
,
Vianna
DM
,
Liu
L
,
Carrive
P
.
Hypocretin/orexin contributes to the expression of some but not all forms of stress and arousal
.
Eur J Neurosci.
2009
;
30
(
8
):
1603
-
1614
.

176.

Wang
D
,
Li
A
,
Dong
K
, et al.
Lateral hypothalamus orexinergic inputs to lateral habenula modulate maladaptation after social defeat stress
.
Neurobiol Stress.
2021
;
14
:
100298
.

177.

Kovács
KJ
,
Sawchenko
PE
.
Sequence of stress-induced alterations in indices of synaptic and transcriptional activation in parvocellular neurosecretory neurons
.
J Neurosci.
1996
;
16
(
1
):
262
-
273
.

178.

Chen
X
,
Li
S
,
Kirouac
GJ
.
Blocking of corticotrophin releasing factor receptor-1 during footshock attenuates context fear but not the upregulation of prepro-orexin mRNA in rats
.
Pharmacol Biochem Behav.
2014
;
120
:
1
-
6
.

179.

Ito
N
,
Yabe
T
,
Gamo
Y
, et al.
I.c.v. administration of orexin-A induces an antidepressive-like effect through hippocampal cell proliferation
.
Neuroscience.
2008
;
157
(
4
):
720
-
732
.

180.

Li
P
,
Li
SB
,
Wang
X
, et al.
Brain circuit of claustrophobia-like behavior in mice identified by upstream tracing of sighing
.
Cell Rep.
2020
;
31
(
11
):
107779
.

181.

Dong
X
,
Li
Y
,
Kirouac
GJ
.
Blocking of orexin receptors in the paraventricular nucleus of the thalamus has no effect on the expression of conditioned fear in rats
.
Front Behav Neurosci.
2015
;
9
:
161
.

182.

Flores
Á
,
Valls-Comamala
V
,
Costa
G
,
Saravia
R
,
Maldonado
R
,
Berrendero
F
.
The hypocretin/orexin system mediates the extinction of fear memories
.
Neuropsychopharmacology.
2014
;
39
(
12
):
2732
-
2741
.

183.

Heydendael
W
,
Sengupta
A
,
Beck
S
,
Bhatnagar
S
.
Optogenetic examination identifies a context-specific role for orexins/hypocretins in anxiety-related behavior
.
Physiol Behav.
2014
;
130
:
182
-
190
.

184.

Bonnavion
P
,
Jackson
AC
,
Carter
ME
,
de Lecea
L
.
Antagonistic interplay between hypocretin and leptin in the lateral hypothalamus regulates stress responses
.
Nat Commun.
2015
;
6
:
6266
.

185.

Johnson
PL
,
Truitt
W
,
Fitz
SD
, et al.
A key role for orexin in panic anxiety
.
Nat Med
.
2010
;
16
(
1
):
111
-
5
.

186.

Lutter
M
,
Krishnan
V
,
Russo
SJ
,
Jung
S
,
McClung
CA
,
Nestler
EJ
.
Orexin signaling mediates the antidepressant-like effect of calorie restriction
.
J Neurosci.
2008
;
28
(
12
):
3071
-
3075
.

187.

Cohen
S
,
Matar
MA
,
Vainer
E
,
Zohar
J
,
Kaplan
Z
,
Cohen
H
.
Significance of the orexinergic system in modulating stress-related responses in an animal model of post-traumatic stress disorder
.
Transl Psychiatry.
2020
;
10
(
1
):
10
.

188.

Dorsey
A
,
de Lecea
L
,
Jennings
KJ
.
Neurobiological and hormonal mechanisms regulating women’s sleep
.
Front Neurosci.
2020
;
14
:
625397
.

189.

Karteris
E
,
Chen
J
,
Randeva
HS
.
Expression of human prepro-orexin and signaling characteristics of orexin receptors in the male reproductive system
.
J Clin Endocrinol Metab.
2004
;
89
(
4
):
1957
-
1962
.

190.

Di Sebastiano
AR
,
Yong-Yow
S
,
Wagner
L
,
Lehman
MN
,
Coolen
LM
.
Orexin mediates initiation of sexual behavior in sexually naive male rats, but is not critical for sexual performance
.
Horm Behav.
2010
;
58
(
3
):
397
-
404
.

191.

Silveyra
P
,
Catalano
PN
,
Lux-Lantos
V
,
Libertun
C
.
Impact of proestrous milieu on expression of orexin receptors and prepro-orexin in rat hypothalamus and hypophysis: actions of Cetrorelix and Nembutal
.
Am J Physiol Endocrinol Metab.
2007
;
292
(
3
):
E820
-
E828
.

192.

D’Anna
KL
,
Gammie
SC
.
Hypocretin-1 dose-dependently modulates maternal behaviour in mice
.
J Neuroendocrinol.
2006
;
18
(
8
):
553
-
566
.

193.

Grieb
ZA
,
Holschbach
MA
,
Lonstein
JS
.
Interaction between postpartum stage and litter age on maternal caregiving and medial preoptic area orexin
.
Physiol Behav.
2018
;
194
:
430
-
436
.

194.

Kautzky-Willer
A
,
Harreiter
J
,
Pacini
G
.
Sex and gender differences in risk, pathophysiology and complications of type 2 diabetes mellitus
.
Endocr Rev.
2016
;
37
(
3
):
278
-
316
.

195.

Eva
C
,
Oberto
A
,
Longo
A
,
Palanza
P
,
Bertocchi
I
.
Sex differences in behavioral and metabolic effects of gene inactivation: the neuropeptide Y and Y receptors in the brain
.
Neurosci Biobehav Rev.
2020
;
119
:
333
-
347
.

196.

Sahu
A
,
Kalra
SP
,
Crowley
WR
,
Kalra
PS
.
Testosterone raises neuropeptide-Y concentration in selected hypothalamic sites and in vitro release from the medial basal hypothalamus of castrated male rats
.
Endocrinology.
1989
;
124
(
1
):
410
-
414
.

197.

Urban
JH
,
Bauer-Dantoin
AC
,
Levine
JE
.
Neuropeptide Y gene expression in the arcuate nucleus: sexual dimorphism and modulation by testosterone
.
Endocrinology.
1993
;
132
(
1
):
139
-
145
.

198.

Fontana
R
,
Della Torre
S
,
Meda
C
,
Longo
A
,
Eva
C
,
Maggi
AC
.
Estrogen replacement therapy regulation of energy metabolism in female mouse hypothalamus
.
Endocrinology.
2014
;
155
(
6
):
2213
-
2221
.

199.

Olofsson
LE
,
Pierce
AA
,
Xu
AW
.
Functional requirement of AgRP and NPY neurons in ovarian cycle-dependent regulation of food intake
.
Proc Natl Acad Sci U S A.
2009
;
106
(
37
):
15932
-
15937
.

200.

Qiu
J
,
Bosch
MA
,
Zhang
C
,
Rønnekleiv
OK
,
Kelly
MJ
.
Estradiol protects neuropeptide y/agouti-related peptide neurons against insulin resistance in females
.
Neuroendocrinology.
2020
;
110
(
1-2
):
105
-
118
.

201.

Becker
JB
,
Chartoff
E
.
Sex differences in neural mechanisms mediating reward and addiction
.
Neuropsychopharmacology.
2019
;
44
(
1
):
166
-
183
.

202.

Harp
SJ
,
Martini
M
,
Lynch
WJ
,
Rissman
EF
.
Sexual differentiation and substance use: a mini-review
.
Endocrinology
.
2020
;
161
(
9
):
bqaa129
.

203.

DeVito
EE
,
Babuscio
TA
,
Nich
C
,
Ball
SA
,
Carroll
KM
.
Gender differences in clinical outcomes for cocaine dependence: randomized clinical trials of behavioral therapy and disulfiram
.
Drug Alcohol Depend.
2014
;
145
:
156
-
167
.

204.

Carroll
ME
,
Lynch
WJ
.
How to study sex differences in addiction using animal models
.
Addict Biol.
2016
;
21
(
5
):
1007
-
1029
.

205.

Terner
JM
,
de Wit
H
.
Menstrual cycle phase and responses to drugs of abuse in humans
.
Drug Alcohol Depend.
2006
;
84
(
1
):
1
-
13
.

206.

Justice
AJ
,
de Wit
H
.
Acute effects of d-amphetamine during the follicular and luteal phases of the menstrual cycle in women
.
Psychopharmacology (Berl).
1999
;
145
(
1
):
67
-
75
.

207.

Justice
AJ
,
de Wit
H
.
Acute effects of estradiol pretreatment on the response to d-amphetamine in women
.
Neuroendocrinology.
2000
;
71
(
1
):
51
-
59
.

208.

Moran-Santa Maria
MM
,
Flanagan
J
,
Brady
K
.
Ovarian hormones and drug abuse
.
Curr Psychiatry Rep.
2014
;
16
(
11
):
511
.

209.

Lynch
WJ
,
Carroll
ME
.
Sex differences in the acquisition of intravenously self-administered cocaine and heroin in rats
.
Psychopharmacology (Berl).
1999
;
144
(
1
):
77
-
82
.

210.

Roth
ME
,
Carroll
ME
.
Sex differences in the escalation of intravenous cocaine intake following long- or short-access to cocaine self-administration
.
Pharmacol Biochem Behav.
2004
;
78
(
2
):
199
-
207
.

211.

Lynch
WJ
,
Taylor
JR
.
Sex differences in the behavioral effects of 24-h/day access to cocaine under a discrete trial procedure
.
Neuropsychopharmacology.
2004
;
29
(
5
):
943
-
951
.

212.

Jackson
LR
,
Robinson
TE
,
Becker
JB
.
Sex differences and hormonal influences on acquisition of cocaine self-administration in rats
.
Neuropsychopharmacology.
2006
;
31
(
1
):
129
-
138
.

213.

Hu
M
,
Becker
JB
.
Acquisition of cocaine self-administration in ovariectomized female rats: effect of estradiol dose or chronic estradiol administration
.
Drug Alcohol Depend.
2008
;
94
(
1-3
):
56
-
62
.

214.

Meitzen
J
,
Meisel
RL
,
Mermelstein
PG
.
Sex Differences and the effects of estradiol on striatal function
.
Curr Opin Behav Sci.
2018
;
23
:
42
-
48
.

215.

Calipari
ES
,
Juarez
B
,
Morel
C
,
Walker
DM
,
Cahill
ME
,
Ribeiro
E
,
Roman-Ortiz
C
,
Ramakrishnan
C
,
Deisseroth
K
,
Han
MH
,
Nestler
EJ
.
Dopaminergic dynamics underlying sex-specific cocaine reward
.
Nat Commun
.
2017
;
8
:
13877
.

216.

Grafe
LA
,
Cornfeld
A
,
Luz
S
,
Valentino
R
,
Bhatnagar
S
.
Orexins mediate sex differences in the stress response and in cognitive flexibility
.
Biol Psychiatry.
2017
;
81
(
8
):
683
-
692
.

217.

Viveros
MP
,
Llorente
R
,
López-Gallardo
M
, et al.
Sex-dependent alterations in response to maternal deprivation in rats
.
Psychoneuroendocrinology.
2009
;
34
Suppl 1
:
S217
-
S226
.

218.

James
MH
,
Campbell
EJ
,
Walker
FR
, et al.
Exercise reverses the effects of early life stress on orexin cell reactivity in male but not female rats
.
Front Behav Neurosci.
2014
;
8
:
244
.

This article is published and distributed under the terms of the Oxford University Press, Standard Journals Publication Model (https://academic.oup.com/journals/pages/open_access/funder_policies/chorus/standard_publication_model)