Skip to main content

REVIEW article

Front. Neurosci., 22 August 2023
Sec. Translational Neuroscience
This article is part of the Research Topic Recent Advances in Sleep and Circadian Rhythms: The Hypothalamus and its Relationship with Appetite View all 5 articles

Orexin and MCH neurons: regulators of sleep and metabolism

  • 1Pharmacology Institute, Medical Faculty Heidelberg, Heidelberg University, Heidelberg, Germany
  • 2Department of Systems Pharmacology and Translational Therapeutics, Perelman School of Medicine, University of Pennsylvania, Philadelphia, PA, United States

Sleep-wake and fasting-feeding are tightly coupled behavioral states that require coordination between several brain regions. The mammalian lateral hypothalamus (LH) is a functionally and anatomically complex brain region harboring heterogeneous cell populations that regulate sleep, feeding, and energy metabolism. Significant attempts were made to understand the cellular and circuit bases of LH actions. Rapid advancements in genetic and electrophysiological manipulation help to understand the role of discrete LH cell populations. The opposing action of LH orexin/hypocretin and melanin-concentrating hormone (MCH) neurons on metabolic sensing and sleep-wake regulation make them the candidate to explore in detail. This review surveys the molecular, genetic, and neuronal components of orexin and MCH signaling in the regulation of sleep and metabolism.

Introduction

Living beings on Earth maintain internal stability despite enormous environmental challenges. This process of maintenance of physiological stability is called homeostasis. In the mammalian body, homeostasis applies to the processes that regulate critical physiological parameters such as blood pressure, heart rate, plasma glucose, body temperature, feeding, and sleep. Feeding and sleep are mutually exclusive behaviors requiring distinct but interdependent homeostatic needs. To survive, organisms require strong coordination of these behaviors to achieve their respective homeostatic conditions. For example, wakefulness is required for foraging and food consumption. The mammalian hypothalamus crucially regulates these homeostatic functions.

The hypothalamus is one of the most complex and heterogeneous brain structures involved in the regulation of numerous homeostatic functions by integrating peripheral and central signals of circadian rhythms, sleep pressure, and energy metabolism. This diverse region of the brain is subdivided into 11 anatomically distinct nuclei having 34 neuronal and 11 non-neuronal cell types, cumulative actions of these cells regulate the sleep and metabolic processes (Chen et al., 2017; Rossi et al., 2019; Jha et al., 2022; Bear et al., 2023). The complex neuronal network made up of the projections from these nuclei to the entire brain, including intrahypothalamic connections regulates behavior and physiology. The metabolic aberrations in disturbed sleep conditions and the prevalence of sleep abnormalities in metabolic syndrome indicate the involvement of proximal hypothalamic neuronal circuitry regulating sleep and metabolism. The mechanistic understanding of these networks would be essential and bring wide-ranging clinical significance.

Hypothalamic regulation of sleep and metabolism

The interactive action of arousal and sleep-promoting areas of the mammalian brain involves the regulation of the sleep-wake cycle (Saper and Fuller, 2017). Wake-sleep transition is the manifestation of inhibition of sleep or wake-promoting areas in the brainstem and hypothalamus (Saper, 2006). The pathway that stimulates and maintains wakefulness consists of glutamatergic inputs from parabrachial and pedunculopontine tegmental nuclei (PPT) to the basal forebrain (BF), and GABAergic and cholinergic neurons in the BF that innervates the cerebral cortex (Saper and Fuller, 2017). Further, GABAergic neurons in the lateral hypothalamus (LH) promote wakefulness by inhibiting sleep-promoting neurons in the thalamus and preoptic area (Herrera et al., 2016; Venner et al., 2016). Wake-promoting orexinergic neurons in the LH that mainly use glutamate to transmit their signals are spatially intermingled with sleep-promoting melanin-concentrating-hormone (MCH)-expressing cells (Rosin et al., 2003). The MCH-expressing cells are primarily GABAergic and found in LH and Zona Incerta (ZI) (Adamantidis and de Lecea, 2008b; Rolls et al., 2010). The projections of both orexin and MCH-expressing neurons to the cortex, hippocampus, amygdala, nucleus accumbens (NAc), hypothalamus, thalamus, ventral tegmental area (VTA), locus coeruleus (LC), and raphe nucleus indicating their intra- and extrahypothalamic functions (Concetti and Burdakov, 2021). Wake promotion inhibits sleep-promoting circuitries lie mainly in the hypothalamic ventrolateral preoptic (VLPO) and median preoptic (MnPO) areas. Sleep-active GABAergic neurons of preoptic areas and brainstem project to the wake-promoting area and inhibiting them in regulated manners (Saper and Fuller, 2017). This mutual inhibition regulates the sleep-wake transition.

Another critical role of the hypothalamus is to maintain energy homeostasis by regulating food intake. Like sleep, energy metabolism is also regulated by mutual inhibitory circuitries. The arcuate nucleus (ARC) harbors appetite-promoting Neuropeptide Y (NPY) and Agouti-related protein (AgRP) neurons that mutually inhibit the appetite-suppressing pro-opiomelanocortin (POMC) and amphetamine-related transcript (CART) neurons. These sets of neurons act as sensors of satiety-promoting leptin and appetite-stimulating ghrelin hormones. Leptins inhibit NPY/AgRP neurons and activate POMC/CART whereas ghrelin activates NPY/AgRP and inhibit POMC/CART neurons. The ARC integrates these peripheral signals and transmits them to other hypothalamic areas such as the dorsomedial nucleus (DMH), the paraventricular nucleus (PVH), and the LH (Milbank and Lopez, 2019). The orexin neurons in LH sense peripheral hormonal signals and levels of metabolites like glucose and amino acids. This is evident from the anatomical connection of orexin neurons to the other metabolic nuclei of the hypothalamus. Further, electrophysiological studies reveal that wake-promoting orexin neurons functionally regulate the NPY, POMC, and glucose-responsive neurons in the ARC and ventromedial nucleus of the hypothalamus (VMH) (Muroya et al., 2004). Interestingly, sleep-promoting neurons also regulate metabolism as fasting increases the expression of MCH levels, and activation of MCH neurons reduces energy expenditures (Pissios et al., 2006).

The LH is the heterogeneous structure located in the posterior hypothalamus and its diverse cell populations have been implicated in the regulation of an array of fundamental physiological processes that includes sleep, feeding, and energy metabolism (Stuber and Wise, 2016). The LH harbors heterogeneous neuronal subtypes including orexin, MCH, GABA, glutamatergic, galanin, neurotensin-releasing (Nts), leptin-receptor (LepRb) expressing neurons, and substance P-releasing neurons (Mickelsen et al., 2019; Rossi et al., 2019). In this review, we discuss the LH’s orexin and MCH neuronal circuitries that regulate energy metabolism and sleep.

LHMCH neurons

Melanin-concentrating hormone neurons are abundant in the LH, though few cells are located within the ZI. The LHMCH neurons were reported to send extensive projections to different brain areas (Skofitsch et al., 1985; Bittencourt et al., 1992; Risold et al., 1997; Murray et al., 2000; Bittencourt, 2011), including structures involved in regulating sleep-wake cycle, feeding behavior, body weight and energy balance (Qu et al., 1996; Shimada et al., 1998; Verret et al., 2003; Jego et al., 2013; Konadhode et al., 2013; Yoon and Lee, 2013). Besides MCH peptide, LHMCH neurons co-express many other neurotransmitters and neuropeptides including CART (Broberger, 1999; Elias et al., 2001), nesfatin-1 (Fort et al., 2008), neuropeptide-EI and neuropeptide-GE (Nahon et al., 1989; Bittencourt, 2011). To date, the molecular phenotypes of LHMCH neurons is still a matter of debate to decipher whether these neurons co-release GABA, glutamate, or both. Indeed, previous studies have demonstrated that LHMCH neurons express glutamic acid decarboxylase (GAD) 67 and GAD65 a key enzyme in GABA synthesis (Elias et al., 2001; Harthoorn et al., 2005; Shin et al., 2007). In addition, the results from the immunohistochemical study have reported that LHMCH terminals express the vesicular GABA transporter (VGAT) that plays an essential role in carrying GABA from the neuronal cytoplasm into the synaptic cleft (Del Cid-Pellitero and Jones, 2012). In contrast, consecutive work showed that LHMCH neurons do not overlap with VGAT-positive GABAergic neurons and were not able to express VGAT (Chee et al., 2015; Jennings et al., 2015; Mickelsen et al., 2017). Furthermore, Jego et al. (2013) confirmed that LHMCH neurons release the inhibitory neurotransmitter GABA. Collectively, these findings hint that LHMCH neurons may perhaps synthesize and release GABA. Paradoxically, other studies revealed that LHMCH neurons are not exclusively GABAergic, but they also express the vesicular glutamate transporter 2 (VGLUT2) and presumably might produce glutamate (Abrahamson et al., 2001; Chee et al., 2015). More recently, by using molecular profiling including RNAscope combined with the immunohistochemical approach Schneeberger et al. (2018) highlighted that 97% of LHMCH neurons express VGLUT2 but not VGAT suggesting that the vast majority of LHMCH neurons are glutamatergic.

LHMCH neurons and energy metabolism

The role of LHMCH neurons in the regulation of energy balance and metabolism has been studied vastly (Pissios et al., 2006; Al-Massadi et al., 2021; Lord et al., 2021). Earlier studies combing electrical stimulation and lesion approaches distinguished LH as a feeding center (Anand and Brobeck, 1951; Delgado and Anand, 1953; Teitelbaum and Stellar, 1954). More precisely, it was shown that LHMCH neurons were directly involved in the modulation of energy balance and glucose homeostasis by controlling feeding behavior, adipose tissue thermogenesis, and locomotor activity. Under fasting conditions, MCH mRNA expression increased in lean mice as well as in leptin-deficient (ob/ob) obese mice (Qu et al., 1996). Acute intracerebroventricular (ICV) administration of MCH in rodents induced short-term but robust increase in food intake (Qu et al., 1996; Sahu, 1998; Della-Zuana et al., 2002) and chronic infusion enhances food consumption and body weight associated with the substantial increase in energy storage and reduction in energy expenditure (Della-Zuana et al., 2002; Gomori et al., 2003; Ito et al., 2003; Glick et al., 2009). Further insights supporting the critical role of LHMCH neurons in regulation of energy homeostasis have emerged from genetic studies. Transgenic mice that overexpress MCH showed sustained hyperphagia, and mild weight gain associated with impaired glucose tolerance and insulin resistance (Ludwig et al., 2001). In contrast, targeted deletion of the Mch gene and MCH neurons-ablation exhibit leanness and weight loss due to hypophagia and increased energy expenditure in mice (Shimada et al., 1998; Kokkotou et al., 2005; Alon and Friedman, 2006; Izawa et al., 2022). On the same line, Jeon et al. (2006) confirmed the leaned phenotypes in aged mice lacking the Mch gene, but also reported better glucose tolerance and insulin sensitivity in these animals. Moreover, mice lacking MCH receptor 1 (MCH-R1) present normal body weight, and lean phenotype with decreased fat mass because of their hyperactivity. Intriguingly, MCH-R1 deficient mice are hyperphagic when fed on a regular chow diet and substantially resistant to high-fat diet induced obesity (Marsh et al., 2002). Furthermore, the hyperphagic phenotype persisted in ob/ob mice lacking the Mch gene, however, they exhibit a remarkable reduction in body fat due to increased energy expenditure. Regarding glucose homeostasis, the disruption of the Mch gene in ob/ob mice has improved glucose tolerance but hyperinsulinemia remained (Segal-Lieberman et al., 2003). It is noteworthy that insulin increases the activity of LHMCH neurons through phosphatidylinositol 3-kinase signaling. Thus, it has been revealed that specific deletion of insulin receptors in LHMCH neurons does not affect energy balance and glucose homeostasis in mice fed on a regular chow diet whereas it improved peripheral glucose metabolism by enhancing hepatic insulin sensitivity and suppressing the production of hepatic glucose in mice exposed to high-fat diet (Hausen et al., 2016). In aggregate, the abovementioned data (summarized in Table 1) indicate that LHMCH neurons are fundamental in regulating energy expenditure and glucose homeostasis, therefore they might be an attractive target for innovative and efficient treatment for obesity and its comorbidities. The mechanistic signaling of LHMCH neurons implicated in regulating energy expenditure and glucose metabolism is not clearly understood yet. Recently, Izawa et al. (2022) suggested that LHMCH neurons might regulate brown adipose tissue (BAT) activity and energy expenditure by sending projections to the medullary raphe nucleus to inhibit sympathetic inputs in BAT. Besides the traditional synaptic transmission, LHMCH neurons convey its orexigenic effects through a complementary pathway involving the cerebral spinal fluid (CSF). It has been shown that LHMCH neurons modulate CSF flow by regulating the frequency of ciliated ependymal cells in the third ventricle (Conductier et al., 2013). Additionally, chemogenetic activation of LHMCH neurons triggers the release of MCH peptide into the CSF which in turn promotes an increment in food intake whereas the limitation of the bioavailability of MCH present in the CSF significantly reduced feeding (Noble et al., 2018). Importantly, recent outcomes unveiled that LHMCH neurons expressing the vascular endothelial growth factor A (VEGFA) regulate the permeability of the median eminence (ME) microvascular plexus and, thus, modulate leptin action in the arcuate nucleus (ARC) to control food intake through VEGFA-dependent mechanism (Jiang et al., 2020). Together, these data propose the functional interaction between LHMCH neurons and ME barrier components in sensing and processing circulating metabolic signals fundamental to regulating energy homeostasis and metabolism.

TABLE 1
www.frontiersin.org

Table 1. Summary of studies that investigated the role of MCH system in food intake and metabolism.

LHMCH neurons and sleep-wake cycle

There is numerous evidence that establishes the role of LHMCH neurons in sleep-wake regulation. Based upon the earlier neuroanatomical experiments where c-Fos was used as a marker of neuronal activity, Verret et al. (2003) noticed that a large majority of LHMCH neurons were active during rapid eye-movement (REM) sleep rebound that followed 72 h of selective REM sleep deprivation. In addition, they found that ICV infusions of MCH peptide significantly increased the number of REM bouts (up to 200%) without affecting their duration and provoked a modest prolongation in the time spent in non-rapid eye-movement (NREM) sleep (up to 70%) in a dose-dependent manner (Verret et al., 2003). Subsequent histological studies supported the previous findings and reinforced the hypothesis implying LHMCH neurons in promoting sleep (Modirrousta et al., 2005; Hanriot et al., 2007; Kitka et al., 2011). Similarly, other studies explore the effects of microinjections of MCH into different brain areas involved in sleep-wake regulation. Of note, targeted injections of MCH into wake-promoting nuclei including the dorsal raphe nucleus (DRN) in the rat and cat (Lagos et al., 2009; Devera et al., 2015), median raphe nucleus (MnR) (Pascovich et al., 2020, 2021), LC (Monti et al., 2015), BF (Lagos et al., 2012), and nucleus pontis oralis (NPO) of the cat (Torterolo et al., 2009) produced a dose-dependent increase in REM sleep. In contrast, the local infusions of MCH into the sublaterodorsal tegmental nucleus (SLD), recognized as the key structure involved in REM sleep generation, significantly impeded REM sleep in rats because of decreasing the time spent in REM sleep and the number of REM bouts (Monti et al., 2016). Moreover, microinjections of MCH peptide directly into the VLPO, one of the major NREM-promoting regions, increased the time spent in NREM sleep without affecting REM sleep (Benedetto et al., 2013). Curiously, subcutaneous administration of MCH-R1 antagonists decreased the time spent in sleep stages and prolonged the onset latency of both NREM and REM sleep (Ahnaou et al., 2008), however, the oral supply of MCH-R1 antagonist has no effects on sleep-wake pattern (Able et al., 2009). In agreement with the pharmacological studies, MCH-R1 knockout mice exhibit a significant decrease in NREM sleep through the light-dark cycle, along with this an enhancement in wakefulness and reduction of REM sleep was detected in these transgenic mice when they were exposed to a restraint stress procedure, followed by a homeostatic rebound sleep (Ahnaou et al., 2011). Moreover, targeted deletion of the Mch gene in mice increased wakefulness and reduced time spent in NREM and REM sleep compared to wild-type animals. Under fasting conditions, these transgenic mice displayed a massive reduction in REM sleep and remarkable hyperactivity correlated with their lean phenotype (Willie et al., 2008). These behavioral responses in mice lacking the Mch gene drew the attention of researchers to investigate MCH-dependent mechanisms underlying sleep-wake regulation in response to changes in energy homeostasis (Willie et al., 2001; Arrigoni et al., 2019). Inconsistent with the previous reports which depicted the integral role of the MCH system in REM sleep regulation, Adamantidis et al. (2008) showed that MCH-R1 knockout mice present an unexpected increase in the REM sleep during the natural sleep-wake cycle and after total sleep deprivation. This differing outcome might be related to compensatory mechanisms established to counterbalance the MCH-R1 disruption caused by the targeted gene deletion approach. Consistent with these crucial findings, in vivo electrophysiology recordings combined with juxtacellular labeling of neurons in head-fixed rats revealed that LHMCH neurons were quiet during wakefulness, occasionally firing during NREM sleep and discharging at their maximum rate during REM sleep (Hassani et al., 2009). Subsequently, both deep-brain calcium imaging and fiber photometry studies performed in freely behaving mice have also confirmed that LHMCH neurons displayed a robust activity during REM sleep as well as during the transition from NREM to REM sleep, whereas there were less active in wakefulness and during the transition from REM sleep to wakefulness (Blanco-Centurion et al., 2019; Izawa et al., 2019). Despite this outstanding experimental evidence, no consensus has been reached yet to decipher the specific role of LHMCH neurons in the regulation of REM and NREM sleep. To clarify this point, a collection of optogenetic or chemogenetic experiments were deployed to scrutinize the defined role of LHMCH neurons in the neurobiological mechanisms of sleep-wake behavior. For instance, acute optogenetic activation of LHMCH neurons at 20 Hz during NREM sleep enhanced transitions from NREM to REM sleep while the duration of REM sleep episodes was significantly extended when the optogenetic stimulation of LHMCH neurons occurred at the onset of REM sleep (Jego et al., 2013). Another group of researchers showed that optogenetic stimulation of LHMCH neurons at 10 Hz for 3 h facilitated the transition from NREM to REM sleep, resulting in a significant increase in the time spent in REM sleep and a concomitant decrease in NREM sleep time (Tsunematsu et al., 2014). Surprisingly, chronic optogenetic activation of MCH neurons (ZI and LH) induced a robust increase in the total time in NREM and REM sleep during the night period and notably increased electroencephalogram (EEG) delta power (0.5–4 Hz), an electrophysiological indicator of sleep intensity. For note, Konadhode et al. (2013) and Blanco-Centurion et al. (2016) found that only REM sleep time extended upon optogenetic stimulation during the daytime in nocturnal rodents. Presumably, these divergent outcomes reported in the abovementioned optogenetic studies might be due to differences in genetic strategies implemented to selectively target MCH neurons as well as to differences in the light pulse stimulation paradigms applied to manipulate the activity of these neurons. In addition, chemogenetic activation of LHMCH neurons increased the number of REM bouts during the light period, which was doubled during the dark period, whereas the duration of bouts did not change (Vetrivelan et al., 2016). In this study, the authors demonstrated that LHMCH neurons play a crucial role in promoting REM sleep and facilitating transitions from NREM to REM sleep. Nevertheless, the role of LHMCH neurons in the spontaneous REM sleep generation is still perplexing since several experiments have yielded inconsistent outcomes. For instance, selective ablation of MCH neurons using a genetically targeted diphtheria toxin approach significantly increased the number of REM bouts and shortened the mean bout duration of wake during the light period (Vetrivelan et al., 2016). In another study, Tet-Off system mice were used to specifically disrupt MCH neurons in a reversible and controlled manner. Paradoxically, the results obtained from this study showed that MCH neurons ablation did not affect the total time in REM sleep and the mean episode duration of REM sleep (Tsunematsu et al., 2014). Moreover, transgenic mice with Ataxin 3-mediated ablation of MCH neurons unexpectedly displayed an increase in the REM sleep amounts during the light period (Varin et al., 2016). Conversely, acute optogenetic silencing of MCH neurons, while mice were in REM sleep, did not change REM sleep episode duration (Jego et al., 2013). In addition, chemogenetic inhibition of MCH neurons increased the NREM sleep amounts and notably extended the mean bout duration of NREM sleep without affecting REM sleep duration, suggesting that active MCH neurons hinder the generation of NREM sleep to facilitate the entry into REM sleep (Varin et al., 2018).

Collectively, evidence from pharmacological, electrophysiology, genetic, chemogenetic, and optogenetic studies (summarized in Table 2) revealed that the MCH system plays a critical role in the regulation of REM sleep, whereas further investigations are required to unravel their role in NREM sleep modulation.

TABLE 2
www.frontiersin.org

Table 2. Summary of studies that investigated the role of MCH system in sleep-wake regulation.

LHOrexin neurons

Orexin neurons are exclusively localized in LH and the adjacent perifornical area (PFH) and send widespread projections throughout the central nervous system (CNS) (Peyron et al., 1998; Nambu et al., 1999) implicating in the regulation of various behavioral and physiological processes predominantly associated with feeding behavior, energy homeostasis, sleep-wake cycle, and reward system (Willie et al., 2001; Yamanaka et al., 2002, 2003; Harris et al., 2005). Previous studies demonstrated that orexin neurons produce two excitatory neuropeptides orexin-A and orexin-B (also known as hypocretin 1 and hypocretin 2) (de Lecea et al., 1998; Sakurai et al., 1998) and also co-release glutamate (Torrealba et al., 2003; Henny et al., 2010) as well as the inhibitory neuropeptide dynorphin (Chou et al., 2001) and the inhibitory neurotransmitter GABA (Harthoorn et al., 2005). Additionally, orexin-A and orexin-B depolarize the post-synaptic target membrane resulting in increased neuronal excitability by acting selectively on two G protein-coupled receptors (GPCR) named orexin receptor type 1 (OX1R) and orexin receptor type 2 (OX2R). Interestingly, orexin-A binds to OX1R and OX2R, however, orexin-B binds specifically to OX2R (Sakurai et al., 1998; Ammoun et al., 2003; Scammell and Winrow, 2011). Pieces of evidence from subsequent experiments revealed that OX1R couples exclusively to the Gq/11 subclass of GPCR, whereas OX2R couples to Gi/o and Gq subclass of GPCR (Sakurai et al., 1998; van den Pol et al., 1998), mediating orexinergic signaling through the activation of Na+/Ca2+ exchanger (Eriksson et al., 2001; Yang and Ferguson, 2002; Burdakov et al., 2003), or through the decrease of potassium conductance (Ivanov and Aston-Jones, 2000; Bisetti et al., 2006). Additionally, orexin signaling pathways involved other intracellular mechanisms including the activation of phospholipase D/phosphatidic acid (Johansson et al., 2008), phospholipase A/arachidonic acid (Turunen et al., 2012), and mitogen-activated protein kinase cascade (Ramanjaneya et al., 2009; Wen et al., 2015). It is noteworthy to highlight that the activation of orexin neurons triggers excitatory post-synaptic responses whereas the stimulation of MCH neurons engenders inhibitory post-synaptic effects (Sakurai, 2007; Adamantidis and de Lecea, 2008a).

LHOrexin neurons and energy metabolism

A myriad of investigations in the field of molecular, cellular, and behavioral neuroscience provide evidence suggesting the implication of orexin neurons in the regulation of feeding behavior and energy homeostasis. Several studies established that orexin system dysfunction has been implicated in serious neurological disorders including narcolepsy (Lin et al., 1999; Thannickal et al., 2000), addiction (Georgescu et al., 2003; Boutrel et al., 2005; Espana et al., 2011), depression (Taheri et al., 2001; Allard et al., 2004; Brundin et al., 2007; Rotter et al., 2011), anxiety (Suzuki et al., 2005; Li et al., 2010; Avolio et al., 2011; Lungwitz et al., 2012), post-traumatic disorder (Strawn et al., 2010), schizophrenia (Nishino et al., 2002; Dalal et al., 2003; Meerabux et al., 2005; Fukunaka et al., 2007), and severe eating behaviors and metabolic impairments such as anorexia nervosa (Bronsky et al., 2011; Janas-Kozik et al., 2011), hyperphagia and eventually obesity in Prader-Willi syndrome (Nevsimalova et al., 2005; Fronczek et al., 2009).

Several studies have reported that prepro-orexin mRNA level and also the activity of orexin neurons was significantly increased during fasting (Sakurai et al., 1998; Lopez et al., 2000; Diano et al., 2003; Horvath and Gao, 2005). Further, ICV microinjection experiments of orexins have confirmed the potential role of orexins in feeding behavior and energy homeostasis. In fact, acute ICV administration of orexin-A in freely fed rats enhanced food consumption in a dose-dependent manner during the light phase (Sakurai et al., 1998; Edwards et al., 1999; Haynes et al., 1999; Dube et al., 2000; Jain et al., 2000; Yamanaka et al., 2000; Lopez et al., 2002). In the same line with these previous outcomes, microinjections of orexin-A directly into different hypothalamic nuclei such as paraventricular nucleus (PVN), dorsomedial nucleus (DMN), LH, and PFH significantly increased food intake, however, no effect was detected after performing microinjections into ARC, ventromedial nucleus (VMN), preoptic area (POA), central nucleus of the amygdala (CeA) and nucleus of the tractus solitaries (NTS). Similar experiments showed that orexin-B failed to stimulate feeding behavior after infusing it into the aforementioned brain areas (Dube et al., 1999; Thorpe et al., 2003). However, Sweet et al. (1999) reported that orexin-B stimulated feeding only after ICV administration. Subsequent pharmacological studies established that the blockade of OX1R by a selective antagonist (SB-334867-A) provokes a robust reduction in food intake in both fed and fasted rats (Haynes et al., 2000; Rodgers et al., 2001; White et al., 2005). Recently, Jin et al. (2020) reported that infusion of orexin-A into the CeA robustly enhanced palatable high-fat diet consumption suggesting a possible role of the orexin system in the regulation of hedonic feeding. Surprisingly, microinjection of orexin-A into the VLPO significantly increased spontaneous physical activity and non-exercise thermogenesis without affecting food consumption resulting in body weight loss, while blockade of OX1R and OX2R abolished the aforesaid effects of orexin-A and presumably lead to body weight gain (Mavanji et al., 2015; Coborn et al., 2017). These pharmaceutical and behavioral studies endeavoring to elucidate the role of the orexin system in feeding behavior and energy homeostasis are also supported by implementing genetic approaches. Mice lacking orexin neurons (orexin/ataxin-3 transgenic mice) or orexin gene (prepro-orexin knockout mice) exhibit a significant reduction in food intake, water intake, locomotor activity, energy expenditure, and unexpectedly late-onset obesity despite their hypophagic phenotype (Hara et al., 2001, 2005; Fujiki et al., 2006; Zhang et al., 2007). Likewise, selective ablation of orexin neurons using diphtheria toxin fragment A reduced food intake and water intake in orexin-Cre mice even though their body weight was significantly higher compared with control mice (Inutsuka et al., 2014). In agreement with these findings, chemogenetic activation of orexin neurons leads to a robust increase in food intake, water intake, spontaneous physical activity, and the respiratory exchange ratio (Inutsuka et al., 2014). Conversely, another study showed that pharmacogenetic activation of orexin neurons produced a strong enhancement in spontaneous physical activity concomitant with an increase in energy expenditure and unpredictably without inducing any changes in food intake and water intake (Zink et al., 2018). Collectively, these findings (summarized in Table 3) suggest that orexins and their receptors might be considered a promising therapeutic target for the treatment of eating disorders and energy metabolism disturbances including obesity, diabetes, and cardiovascular diseases.

TABLE 3
www.frontiersin.org

Table 3. Summary of studies that investigated the role of orexin system in food intake and metabolism.

LHOrexin neurons and sleep-wake cycle

The orexin neurons are wake-active neurons that fire during the wake period and the extracellular level of orexin peak during wakefulness (Kiyashchenko et al., 2002; Lee et al., 2005) and remain silent during NREM and REM sleep with the exception of burst discharge in phasic REM (Mileykovskiy et al., 2005). It has been shown that ICV injection of orexin induces long periods of wakefulness and suppresses the NREM period (Mieda et al., 2011). Both chemogenetic and optogenetic stimulation of orexin neurons produce wakefulness and strongly suppress REM sleep (Adamantidis et al., 2007; Sasaki et al., 2011). It is argued that the most essential role of orexin is to maintain wakefulness (summarized in Table 4). For example, selective loss of orexin neurons in humans causes narcolepsy (Blouin et al., 2005). The deletion of OX2R produces a phenotype like narcolepsy and restoration of OX2R in double knock-out mice rescues normal sleep-wake phenotype in the mice (Willie et al., 2003; Mochizuki et al., 2011). These experiments suggest that OX2R signaling is crucial for controlling sleep-wake.

TABLE 4
www.frontiersin.org

Table 4. Summary of studies that investigated the role of orexin system in sleep-wake regulation.

There are conflicting results reported regarding histamine as a signaling element in orexin actions. Carter et al. (2009) have shown that optogenetic stimulation of orexin neurons promotes arousal in mice lacking central histamine. Contrary to this, central administration of orexin induces wakefulness in wild-type animals but not in histamine receptor 1 knock-out mice. Orexin neurons also influence sleep as it reduces the NREM and REM episodes. The central orexin signaling results in reduced REM sleep duration (Williams et al., 2008; Mieda et al., 2011; Sasaki et al., 2011). This effect is possibly mediated by the activation of both OX1R and OX2R (Mieda et al., 2011). Narcoleptic patients show short latency of REM sleep and random nap often include bouts of REM sleep (Dantz et al., 1994; Andlauer et al., 2013). Like narcoleptic individuals, mice lacking orexin signaling are also unable to suppress REM sleep bouts during the active period, indicating the role of orexin signaling during the active period to suppress REM sleep and meet temporal needs (Arrigoni et al., 2019).

MCH and orexin neuronal circuitries regulating sleep and metabolism

The MCH neurons in LH and ZI project to the nuclei that involve in promoting sleep and arousal (Monti et al., 2013). These projections positively modulate sleep, especially REM sleep (Torterolo et al., 2011). The LC, DRN, and regions of the ventrolateral periaqueductal gray matter and lateral pontine tegmentum (vlPAG/LPT) that are implicated in REM sleep regulation receive dense MCH projections (Torterolo et al., 2009; Costa et al., 2018). The activation of the MCH terminal in vlPAG/LPT tends to increase the duration of REM sleep (Kroeger et al., 2019). Overall, MCH neurons promote sleep by inhibiting wake-promoting areas like the medial septum (MS) and TMN (Jego et al., 2013; Figure 1). The SLD within the dorsolateral pons and NPO in the subcoeruleus (anatomical equivalent of the SLD in cats) are characterized as REM promoting area (Torterolo et al., 2009; Luppi et al., 2013). It is conceivable that MCH neurons directly activate SLD or NPO, likely through the release of glutamate (Torterolo et al., 2009, 2013; Monti et al., 2016). Further, it is considered that MCH may interact with REM-promoting cholinergic neurons within LDT and PPT based on the identification of MCH axons in these areas (Costa et al., 2018), however, there are no functional data supporting this circuit (Figure 1).

FIGURE 1
www.frontiersin.org

Figure 1. Schematic representation of MCH system. MCH neurons in the lateral hypothalamus and zona incerta project to metabolic relevant and sleep-wake controlling nuclei (Burdakov et al., 2005; Torterolo et al., 2009; Lagos et al., 2012; Benedetto et al., 2013; Jego et al., 2013; Monti et al., 2016; Costa et al., 2018; Kroeger et al., 2019; Jiang et al., 2020). BF, basal Forebrain; DRN, dorsal raphe nucleus; Hipp, hippocampus; ME, median eminence; PPT, pedunculopontine tegmentum; LDT, laterodorsal tegmentum; LC, locus coeruleus; TMN, tuberomammillary nucleus; vLPAG/LPT, ventrolateral periaqueductal gray matter and lateral pontine tegmentum; VLPO, ventrolateral preoptic area; MS, medial septum; SLD/NPO, sublaterodorsal tegmental nucleus and nucleus pontis oralis; MCH, melanin-concentrating hormone neurons. Sleep and metabolic-relevant nuclei are color-coded and excitatory and inhibitory inputs are arrow represented.

Orexin neurons are found only in LH and PFH and similarly project in the CNS like MCH neurons, however, having the opposite effect on the modulation of sleep-wake and metabolism (Figure 2). Orexin neurons project to wake-associated neurons in BF, LC, TMN, VTA, and DRN, in vitro electrophysiology studies have shown that orexin activates neurons in all these regions (Peyron et al., 1998; Horvath et al., 1999b; Brown et al., 2001; Eggermann et al., 2001; Eriksson et al., 2001; Baimel et al., 2017). So far there is no explicit explanation of the action of orexin neurons on these wake-associated areas, however, a general apprehension is that orexin neurons co-release orexin, dynorphin, and glutamate to likely activate target neurons (Arrigoni et al., 2019). The neurons from sleep-promoting areas like MPO, POA, and VLPO project to orexin neurons in LH, these areas harbor GABAergic neurons that are active during the NREM and/or REM sleep episodes and promote NREM sleep (Yoshida et al., 2006; Benedetto et al., 2012; Saito et al., 2013, 2018a; Alam et al., 2014; Chung et al., 2017). It has been shown that GABA release in LH is higher during the sleep period and blocking GABAergic signaling during the sleeping period activates orexin neurons (Nitz and Siegel, 1996; Alam et al., 2005). Moreover, GABAergic input to LH also reached from VTA as activation of GABAergic neuronal terminals in the LH promoted NREM sleep by inhibiting orexin neurons (Chowdhury et al., 2019). A recent study shows that orexin neurons indirectly target and inhibit sleep-promoting VLPO neurons to promote arousal (De Luca et al., 2022). These findings indicate that sleep-active GABAergic input from preoptic areas inhibits orexin neurons. This GABAergic inhibitions and orexin-mediated activation of wake-associated neurons and inhibition of sleep-associated neurons could be a possible mechanism by which orexin neurons regulate sleep and arousal.

FIGURE 2
www.frontiersin.org

Figure 2. Schematic representation of the orexin system. Orexin neurons in the lateral hypothalamus project and receive projection from metabolic-relevant and sleep-wake-controlling nuclei (de Lecea et al., 1998; Peyron et al., 1998; Date et al., 1999; Horvath et al., 1999b; Brown et al., 2001; Eggermann et al., 2001; Eriksson et al., 2001; Funahashi et al., 2003; Yamanaka et al., 2003; Burdakov et al., 2006; Yoshida et al., 2006; Benedetto et al., 2012; Saito et al., 2013, 2018b; Alam et al., 2014; Baimel et al., 2017; Chung et al., 2017; Chowdhury et al., 2019). BF, basal forebrain; DRN, dorsal raphe nucleus; Hipp, hippocampus; Orx, orexin; POA/MPO, preoptic area/medial preoptic area; LDT, laterodorsal tegmentum; PPT, pedunculopontine tegmentum; PVH, Paraventricular nucleus of the hypothalamus; LC, locus coeruleus; TMN, tuberomammillary nucleus; VLPO, ventrolateral preoptic area; VTA, ventral tegmental area; ARC, arcuate nucleus; NPY, Neuropeptide Y; AgRP, agouti-related protein; POMC, pro-opiomelanocortin; CART, amphetamine-related transcript. Sleep and metabolic-relevant nuclei are color-coded and excitatory and inhibitory inputs are arrow represented.

The LH neurons modulate the metabolism by regulating the feeding. The connectivity of LH to ARC may adjust the food intake depending on the energy needs of the animals. The orexin neurons project to ARC that harbor NPY and POMC neurons expressing orexin and leptin receptors (de Lecea et al., 1998; Date et al., 1999; Funahashi et al., 2003). Orexin induces feeding by activating NPY and inhibiting POMC neurons (Dube et al., 2000; Jain et al., 2000; Guan et al., 2001; Ma et al., 2007; Figure 2). The feeding circuity of orexin may extend to PVH as orexin neurons project to PVH and ICV injection of orexin activates ARC (Date et al., 1999; Edwards et al., 1999). However, it is not explicitly known how orexin acts on PVH and whether orexin regulates feeding through PVH. Orexin neurons also act as metabolic sensors as they respond to peripheral metabolic cues. Extracellular glucose inhibits orexin neurons (Burdakov et al., 2006). In addition to that direct sensing of extracellular glucose levels, orexin neurons sense the other peripheral indicators of energy status such as the satiety hormone leptin and hunger hormone ghrelin. Leptin inhibits orexin neurons whereas ghrelin activates the same (Yamanaka et al., 2003). Thus, negative energy balance activates orexin neurons and hence hunger keeps animals awake. Contrary to this, MCH neurons promote positive energy balance. The MCH neurons are activated by high glucose levels and physiological shifts in glucose have the opposite effects on the electrical activity of orexin neurons (Burdakov et al., 2005, 2006). This differential glucose-sensing ability of orexin and MCH neurons suggests that hyperglycemia may reduce feeding by hyperpolarization of excitatory (orexin neurons) and depolarization of inhibitory (MCH neurons) input to ARC neurons. The direct projection of MCH neurons to ARC is not known, however, a recent study suggests that MCH neurons project to the median eminence (ME), and its activation enhances leptin action in the ARC (Jiang et al., 2020).

Orexin and MCH neurons act as sensors of metabolic changes and arousal

In mammals, maintaining the balance between energy intake and energy expenditure is crucial for survival. However, energy homeostasis imbalance underlies serious metabolic disturbances and diseases such as obesity, diabetes, hyperlipidemia, hypertension, cardiovascular diseases, and cancers (Crowley et al., 2002; Calle et al., 2003; Kim et al., 2016; Tanaka and Itoh, 2019).

A panoply of experimental evidence revealed that energy homeostasis is regulated via a complex and widespread neuronal circuit located mainly in the brainstem and hypothalamus (Schwartz et al., 2000; Myers and Olson, 2012; Morton et al., 2014; Roh et al., 2016). Distinct neuronal populations within particular nuclei of the brainstem and the hypothalamus sense variations in the energy status of the body by integrating and responding to multiple peripheral (glucose, insulin, leptin, ghrelin, glucagon-like peptide 1) and central [GABA, NPY, AgRP, α-melanocyte-stimulating hormone (α-MSH), serotonin] metabolic signals to maintain energy homeostasis by coordinating energy intake with energy expenditure over time (Burdakov et al., 2005; Schwartz and Porte, 2005; Williams et al., 2008; Morton et al., 2014; Gautron et al., 2015; Roh et al., 2016; Timper and Bruning, 2017; Mavanji et al., 2022). In this context, it is of interest to highlight the crucial role of orexin and MCH systems in regulating energy balance in response to fasting. In fact, the activation of orexin neurons promotes food foraging and increases energy expenditure, whereas the activation of MCH neurons enhances food intake and decreases energy expenditure leading to an increase in energy storage (Figure 3). Both orexin and MCH neurons are activated by fasting (Qu et al., 1996; Sakurai et al., 1998; Lopez et al., 2000; Yamamoto et al., 2000; Diano et al., 2003; Horvath and Gao, 2005; Mogi et al., 2005; Buczek et al., 2020). Interestingly, it was recently revealed that MCH neurons are activated during the early phase of fasting (12 h of fasting), however, orexin neurons exhibit a delayed activation during food deprivation (24 h of fasting). This alternate activation of MCH and orexin neurons play a potential role in coordinating foraging behaviors and energy storage to adjust energy homeostasis during prolonged fasting (Linehan and Hirasawa, 2022). Taken together, these findings insinuate that orexin and MCH neurons are capable of sensing and integrating circulating metabolic signals that convey precise information regarding the status of energy stores, leading to dynamic coordination between energy intake and energy expenditure to restore energy balance.

FIGURE 3
www.frontiersin.org

Figure 3. Simplified schematic representation of the orexin and MCH systems as a sensor of metabolic changes and arousal. In negative energy balance, low extracellular glucose concentration and high circulating level of ghrelin activate the orexin system but inhibit MCH neurons leading to an increase in orexin release and a decrease in MCH release to promote wakefulness, activity, foraging, and food intake. By contrast, in positive energy balance, high extracellular glucose concentration activates MCH neurons but suppresses orexin neurons which are also inhibited by the circulating level of leptin promoting sleep and decreasing energy expenditure. Orx, orexin neurons; MCH, melanin-concentrating hormone neurons.

Pioneering studies reported that orexin-producing neurons are involved in sensing glucose, ghrelin, and leptin levels and eventually promoting arousal (Figure 2). Indeed, electrophysiological evidence revealed that increasing glucose levels induced a striking hyperpolarization and cessation of both spontaneous and evoked action potentials in isolated orexin neurons (Yamanaka et al., 2003; Burdakov et al., 2005; Sheng et al., 2014). Furthermore, the blockade of glycolytic metabolism of glucose by selective inhibitors of glucokinase failed to change the effects of glucose on the action potentials of orexin neurons. These results indicate that orexin neurons are capable to sense trends in glucose levels independently of glucose metabolism (Gonzalez et al., 2008, 2009). Actually, glucose inhibits orexin neurons by acting at the extracellular tandem-pore K + (K2P) channels to induce membrane hyperpolarization and decrease the firing rate of orexin neurons (Burdakov et al., 2006). Here, it is worthwhile to highlight that glucose inhibited orexin neurons only when their intracellular energy levels are low, but paradoxically glucose failed to block orexin neurons when the intracellular levels of lactate, pyruvate, and ATP are high. These results reveal an unexpected glucose-sensing mechanism in orexin neurons that is tightly modulated by the cellular energy status (Venner et al., 2011). Strikingly, recent experimental findings showed for the first time an unexpected complex relationship between orexin neuron activity and blood glucose changes in living organisms. In fact, orexin neurons activity vs. blood glucose variability exhibited a non-canonical temporal profile instead of the expected linear pattern. Basically, orexin neurons track blood glucose concentration at the temporal resolution of minutes and promptly convey its changes into targeted brain regions to trigger adaptive behavior strategies in order to optimize energy balance (Viskaitis et al., 2022).

In addition to sensing peripheral glucose changes, orexin neurons are also involved in detecting and processing signals from other circulating factors such as leptin. Leptin, a product of ob gene, is an anorexigenic hormone predominantly released by adipose tissues (Zhang et al., 1994) and plays a critical role in regulating satiety, blood glucose levels, and energy homeostasis by acting on defined target neurons of the CNS (Schwartz et al., 1996). Previous findings demonstrated that ICV administration of leptin prevents an increase of prepro-orexin mRNA and orexin receptor 1 mRNA in fasted rats, suggesting that leptin has inhibitory feedback on the regulation of orexin gene expression (Lopez et al., 2000). Moreover, Zhu et al. (2002) confirmed orexin neurons induce feeding behavior through both leptin-sensitive and leptin-insensitive pathways. In this sense, we can speculate that leptin might regulate the activity of orexin neurons via complex circuit mechanisms. Indeed, earlier reports yielded conflicting results concerning the expression of leptin receptors (LepRb) on orexin neurons. Findings from immunohistochemistry studies performed in rodent and monkey brains demonstrated that orexin neurons in the LH possess LepRb and thus supporting the hypothesis that leptin might act directly upon these neurons to reduce food seeking and regulate energy balance (Hakansson et al., 1999; Horvath et al., 1999a; Iqbal et al., 2001). Subsequent investigation revealed that bath application of leptin onto isolated orexin neurons provoked hyperpolarization of the membrane potential and suppressed the action potential firing in these cells, resulting in inhibition of orexin neurons (Yamanaka et al., 2003). However, using transgenic LepRbEGFP mice where enhanced green fluorescence protein (EGFP) expression is under the control of the LepRb promotor to scrutinize the possible colocalization of EGFP with orexin neurons, displayed that LepRb-expressing neurons represent a distinct population from orexin neurons in the LH (Leinninger et al., 2009; Louis et al., 2010; Laque et al., 2013). In general support of these results, further experimental works were performed using electrophysiology recordings in brain slices, knock-in mice lines and single-cell expression profiling approaches to elucidate that orexin neurons do not express LepRb and are only indirectly regulated by leptin (Leinninger et al., 2011; Goforth et al., 2014; Sheng et al., 2014; Mickelsen et al., 2017). Several studies have shown that LepRb-expressing neurons lie in synaptic contact with orexin neurons within the LH and the majority of these LepRb neurons contain neurotensin (LepRbNts) (Louis et al., 2010; Leinninger et al., 2011). In addition, pharmacogenetic activation of LepRbNts in hypothalamic slices hyperpolarized membrane potential and reduced action potential firing in orexin neurons. Likewise, the selective genetic deletion of LepRb from LH LepRbNts neurons abolishes leptin-induced inhibition of orexin neurons (Leinninger et al., 2011; Goforth et al., 2014). Together these data suggest that leptin inhibits indirectly the activity of orexin neurons by acting on LepRbNts cells within the LH. Here it is worthwhile to emphasize that LepRbNts also co-release the inhibitory neuropeptide galanin (Laque et al., 2013) which plays an important role in the regulation of orexin neurons by leptin whereas Nts has a tendency to stimulate these cells indicating that this peptide is not implicated in leptin-induced inhibition of orexin neurons (Goforth et al., 2014). It was also reported that leptin failed to significantly enhance GABAA-mediated inhibitory synaptic transmission in orexin neurons and the blockade of GABA receptors could not prevent leptin inhibition of orexin neurons (Goforth et al., 2014). In aggregate, leptin indirectly inhibits orexin neurons by activating LepRbNts neurons through the release of galanin and via GABA-independent mechanisms including the presynaptic inhibition of glutamate inputs onto orexin neurons and the post-synaptic opening of ATP-sensitive potassium KATP channels.

In addition to glucose and leptin, the orexin system is also involved in sensing other circulating factors and hormones such as ghrelin to coordinate behaviors with metabolic needs. Ghrelin is a gastrointestinal hormone released predominantly from the stomach during periods of energy deficit to enhance appetite and food intake (Ariyasu et al., 2001; Nakazato et al., 2001; Lawrence et al., 2002; Olszewski et al., 2003). Importantly, ghrelin is also produced in the brain by a distinct hypothalamic neuronal population adjacent to the third ventricle between the DMH, the VMH, and the ARC. These neurons send wide projections into several hypothalamic nuclei including the ARC and LH to synapse, respectively, with NPY and orexin neurons (Cowley et al., 2003; Toshinai et al., 2003). For note, ghrelin mediates its effects by binding to growth hormone secretagogue receptors, a subtype of the GPCR family highly expressed in the brain as well as in peripheral tissues including stomach, intestine, pancreas, liver, heart, and skeletal muscles (Kojima et al., 1999; Papotti et al., 2000; Gnanapavan et al., 2002; Sun et al., 2004; Yin et al., 2014). Hence, ghrelin can participate in regulating multiple biological processes comprising glucose metabolism (Broglio et al., 2001; Saad et al., 2002; Dezaki et al., 2004; Verhulst and Depoortere, 2012), energy homeostasis (Ravussin et al., 2001; Druce et al., 2005; Malik et al., 2008; Lin et al., 2011), cardiovascular functions (Mao et al., 2012, 2013; Cao et al., 2013; Khazaei and Tahergorabi, 2013), reproduction (Comninos et al., 2014), cell proliferation (Costa et al., 2011; Delhanty et al., 2014; Miao et al., 2019), inflammation and immune system (Lee et al., 2010; Baatar et al., 2011; Wei et al., 2015; Azizzadeh et al., 2017; Santos et al., 2017), learning and memory performance (Carlini et al., 2002; Diano et al., 2006; Kanoski et al., 2013), sleep-wake cycle and (Tolle et al., 2002; Weikel et al., 2003; Szentirmai et al., 2007a,b), and other circadian rhythms (Yannielli et al., 2007; Wang et al., 2018; Qian et al., 2019). Here it is worthwhile to report that ghrelin regulates feeding behaviors and energy homeostasis by interacting with distinct neuronal populations within the CNS including orexin neurons (Olszewski et al., 2003; Toshinai et al., 2003, 2006). In contrast to leptin and glucose, it has been reported that ghrelin stimulates orexin neurons (Yamanaka et al., 2003; Goforth et al., 2014; Sheng et al., 2014). Previous studies have shown that peripheral or central administration of ghrelin robustly increased food intake and induced Fos expression in orexin-immunoreactive neurons but not in MCH-containing neurons (Nakazato et al., 2001; Lawrence et al., 2002; Tolle et al., 2002; Toshinai et al., 2003). Moreover, electrophysiological evidence showed that ghrelin directly activates isolated orexin neurons by inducing membrane depolarization and increasing the action potential firing in these cells (Yamanaka et al., 2003; Sheng et al., 2014). During periods of starvation, elevated circulating levels of ghrelin enhanced the sensitivity of orexin neurons to glucose changes, and thus contribute to maintaining energy homeostasis (Sheng et al., 2014).

In contrast to orexin neurons which are inhibited by glucose, MCH neurons are activated by glucose. A rise in the extracellular glucose levels directly enhanced the excitability of MCH neurons by inducing membrane depolarization of MCH neurons accompanied by an increase in its resistance (Burdakov et al., 2005). Subsequent investigations revealed that glucose sensing by MCH neurons implicates KATP channels and is modulated by a mitochondrial protein UCP2 that decreases ATP production (Krauss et al., 2003; Kong et al., 2010). Additionally, the action of leptin and ghrelin on MCH neurons is yet to be precisely delineated (Figure 3).

Perspective

Based on the currently available data, it emerges as orexin and MCH system in LH mediating its opposing action on sleep-wake and energy metabolism by utilizing multiple neuronal circuits and peripheral cues. The behavioral strategy required to regulate arousal with respect to hunger and satiety with respect to sleep is under the control of orexin and MCH neurons and their extended circuitries. The MCH neurons promote sleep whereas orexin promotes wakefulness, whereas both these neurons promote feeding by interacting with ARC neurons. However, the feeding preference is different as orexin neurons motivate palatable food consumption, whereas MCH neurons motivate caloric food consumption. Interestingly, hunger and hypoglycemia activate orexin neurons that induce arousal required for foraging and food consumption. On the other hand, MCH neurons sense the rise in glucose levels and promote inactivity and sleep. This indicates interconnectivity of LH to ARC is crucial in maintaining sleep and energy homeostasis and effective to deal with challenges such as starvation and sleep disruption. Sleep disruption influences metabolic processes (Donga et al., 2010; Jha et al., 2016). Sleep deprivation increases ghrelin and decreases leptin levels (Schmid et al., 2008; Mosavat et al., 2021), which activates the orexin system. Thus, these arousing cues promote consummatory behavior, inhibition of sleep, and energy conservation. The inhibition of this signaling in recovery sleep may stabilize it by maintaining the sleep-wake cycle. How the MCH system senses these metabolic cues are not clear yet, however, the MCH system may respond to it by stabilizing sleep and decreasing the energy expenditure by interacting with the orexin neuron activity and brain circuits involved in sleep and metabolism. Moreover, metabolic abnormalities also disrupt sleep (Ogilvie and Patel, 2017). Disruption of the sleep-wake cycle in obesity and other metabolic conditions is not studied at the mechanistic level. There are possibilities that the peripheral metabolic cues may directly or indirectly interact with the orexin/MCH system to alter the sleep phenotype in metabolic disorders. Metabolic disruption may influence LH’s neuronal systems as it has been shown that obesity shifts the activity and transcriptional profile of LHA glutamatergic neurons (Rossi et al., 2019). By knowing these co-localized and interacting neural systems that govern the distinct and interdependent behavioral programs—sleep and feeding, it would be enticing to dissect the neuronal bases of the interaction of both these behaviors.

Conclusion

Sleep and the metabolic system are bidirectionally linked to maintaining homeostasis in challenging environments. In this review, we summarized how molecular and cellular components of MCH and orexin signaling maintain this bidirectionality by integration of sleep-wake and energy metabolism. Both these classes of neurons sense the metabolic signals and regulate the sleep-wake states. A substantial chunk of work has been done to understand how orexin and MCH neurons in LH coordinate the metabolism and behavioral states. Future works on how sleep and metabolism influence each other, and the mechanistic explanation of their interaction would be helpful to assign the target for therapeutic intervention for metabolic and arousal-related disorders.

Author contributions

Both authors listed have made a substantial, direct, and intellectual contribution to the work, and approved it for publication.

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

Able, S. L., Ivarsson, M., Fish, R. L., Clarke, T. L., McCourt, C., Duckworth, J. M., et al. (2009). Localisation of melanin-concentrating hormone receptor 1 in rat brain and evidence that sleep parameters are not altered despite high central receptor occupancy. Eur. J. Pharmacol. 616, 101–106. doi: 10.1016/j.ejphar.2009.06.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Abrahamson, E. E., Leak, R. K., and Moore, R. Y. (2001). The suprachiasmatic nucleus projects to posterior hypothalamic arousal systems. Neuroreport 12, 435–440.

Google Scholar

Adamantidis, A. R., Zhang, F., Aravanis, A. M., Deisseroth, K., and de Lecea, L. (2007). Neural substrates of awakening probed with optogenetic control of hypocretin neurons. Nature 450, 420–424. doi: 10.1038/nature06310

PubMed Abstract | CrossRef Full Text | Google Scholar

Adamantidis, A., and de Lecea, L. (2008a). Physiological arousal: A role for hypothalamic systems. Cell Mol. Life Sci. 65, 1475–1488.

Google Scholar

Adamantidis, A., and de Lecea, L. (2008b). Sleep and metabolism: Shared circuits, new connections. Trends Endocrinol. Metab. 19, 362–370. doi: 10.1016/j.tem.2008.08.007

PubMed Abstract | CrossRef Full Text | Google Scholar

Adamantidis, A., Salvert, D., Goutagny, R., Lakaye, B., Gervasoni, D., Grisar, T., et al. (2008). Sleep architecture of the melanin-concentrating hormone receptor 1-knockout mice. Eur. J. Neurosci. 27, 1793–1800. doi: 10.1111/j.1460-9568.2008.06129.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Ahnaou, A., Dautzenberg, F. M., Huysmans, H., Steckler, T., and Drinkenburg, W. H. (2011). Contribution of melanin-concentrating hormone (MCH1) receptor to thermoregulation and sleep stabilization: Evidence from MCH1 (-/-) mice. Behav. Brain Res. 218, 42–50. doi: 10.1016/j.bbr.2010.11.019

PubMed Abstract | CrossRef Full Text | Google Scholar

Ahnaou, A., Drinkenburg, W. H., Bouwknecht, J. A., Alcazar, J., Steckler, T., and Dautzenberg, F. M. (2008). Blocking melanin-concentrating hormone MCH1 receptor affects rat sleep-wake architecture. Eur. J. Pharmacol. 579, 177–188. doi: 10.1016/j.ejphar.2007.10.017

PubMed Abstract | CrossRef Full Text | Google Scholar

Alam, M. A., Kumar, S., McGinty, D., Alam, M. N., and Szymusiak, R. (2014). Neuronal activity in the preoptic hypothalamus during sleep deprivation and recovery sleep. J. Neurophysiol. 111, 287–299.

Google Scholar

Alam, M. N., Kumar, S., Bashir, T., Suntsova, N., Methippara, M. M., Szymusiak, R., et al. (2005). GABA-mediated control of hypocretin- but not melanin-concentrating hormone-immunoreactive neurones during sleep in rats. J. Physiol. 563, 569–582. doi: 10.1113/jphysiol.2004.076927

PubMed Abstract | CrossRef Full Text | Google Scholar

Allard, J. S., Tizabi, Y., Shaffery, J. P., Trouth, C. O., and Manaye, K. (2004). Stereological analysis of the hypothalamic hypocretin/orexin neurons in an animal model of depression. Neuropeptides 38, 311–315. doi: 10.1016/j.npep.2004.06.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Al-Massadi, O., Dieguez, C., Schneeberger, M., Lopez, M., Schwaninger, M., Prevot, V., et al. (2021). Multifaceted actions of melanin-concentrating hormone on mammalian energy homeostasis. Nat. Rev. Endocrinol. 17, 745–755. doi: 10.1038/s41574-021-00559-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Alon, T., and Friedman, J. M. (2006). Late-onset leanness in mice with targeted ablation of melanin concentrating hormone neurons. J. Neurosci. 26, 389–397. doi: 10.1523/JNEUROSCI.1203-05.2006

PubMed Abstract | CrossRef Full Text | Google Scholar

Ammoun, S., Holmqvist, T., Shariatmadari, R., Oonk, H. B., Detheux, M., Parmentier, M., et al. (2003). Distinct recognition of OX1 and OX2 receptors by orexin peptides. J. Pharmacol. Exp. Ther. 305, 507–514.

Google Scholar

Anand, B. K., and Brobeck, J. R. (1951). Localization of a “feeding center” in the hypothalamus of the rat. Proc. Soc. Exp. Biol. Med. 77, 323–324.

Google Scholar

Andlauer, O., Moore, H., Jouhier, L., Drake, C., Peppard, P. E., Han, F., et al. (2013). Nocturnal rapid eye movement sleep latency for identifying patients with narcolepsy/hypocretin deficiency. JAMA Neurol. 70, 891–902.

Google Scholar

Ariyasu, H., Takaya, K., Tagami, T., Ogawa, Y., Hosoda, K., Akamizu, T., et al. (2001). Stomach is a major source of circulating ghrelin, and feeding state determines plasma ghrelin-like immunoreactivity levels in humans. J. Clin. Endocrinol. Metab. 86, 4753–4758. doi: 10.1210/jcem.86.10.7885

PubMed Abstract | CrossRef Full Text | Google Scholar

Arrigoni, E., Chee, M. J. S., and Fuller, P. M. (2019). To eat or to sleep: That is a lateral hypothalamic question. Neuropharmacology 154, 34–49.

Google Scholar

Avolio, E., Alo, R., Carelli, A., and Canonaco, M. (2011). Amygdalar orexinergic-GABAergic interactions regulate anxiety behaviors of the Syrian golden hamster. Behav. Brain Res. 218, 288–295. doi: 10.1016/j.bbr.2010.11.014

PubMed Abstract | CrossRef Full Text | Google Scholar

Azizzadeh, F., Mahmoodi, J., Sadigh-Eteghad, S., Farajdokht, F., and Mohaddes, G. (2017). Ghrelin exerts analgesic effects through modulation of IL-10 and TGF-beta levels in a rat model of inflammatory pain. Iran. Biomed. J. 21, 114–119. doi: 10.18869/acadpub.ibj.21.2.114

PubMed Abstract | CrossRef Full Text | Google Scholar

Baatar, D., Patel, K., and Taub, D. D. (2011). The effects of ghrelin on inflammation and the immune system. Mol. Cell Endocrinol. 340, 44–58.

Google Scholar

Baimel, C., Lau, B. K., Qiao, M., and Borgland, S. L. (2017). Projection-target-defined effects of orexin and dynorphin on VTA dopamine neurons. Cell Rep. 18, 1346–1355. doi: 10.1016/j.celrep.2017.01.030

PubMed Abstract | CrossRef Full Text | Google Scholar

Bear, M. H., Reddy, V., and Bollu, P. C. (2023). Neuroanatomy, hypothalamus. Treasure Island, FL: StatPearls.

Google Scholar

Benedetto, L., Chase, M. H., and Torterolo, P. (2012). GABAergic processes within the median preoptic nucleus promote NREM sleep. Behav. Brain Res. 232, 60–65.

Google Scholar

Benedetto, L., Rodriguez-Servetti, Z., Lagos, P., D’Almeida, V., Monti, J. M., and Torterolo, P. (2013). Microinjection of melanin concentrating hormone into the lateral preoptic area promotes non-REM sleep in the rat. Peptides 39, 11–15. doi: 10.1016/j.peptides.2012.10.005

PubMed Abstract | CrossRef Full Text | Google Scholar

Bisetti, A., Cvetkovic, V., Serafin, M., Bayer, L., Machard, D., Jones, B. E., et al. (2006). Excitatory action of hypocretin/orexin on neurons of the central medial amygdala. Neuroscience 142, 999–1004. doi: 10.1016/j.neuroscience.2006.07.018

PubMed Abstract | CrossRef Full Text | Google Scholar

Bittencourt, J. C. (2011). Anatomical organization of the melanin-concentrating hormone peptide family in the mammalian brain. Gen. Comp. Endocrinol. 172, 185–197. doi: 10.1016/j.ygcen.2011.03.028

PubMed Abstract | CrossRef Full Text | Google Scholar

Bittencourt, J. C., Presse, F., Arias, C., Peto, C., Vaughan, J., Nahon, J. L., et al. (1992). The melanin-concentrating hormone system of the rat brain: An immuno- and hybridization histochemical characterization. J. Comp. Neurol. 319, 218–245. doi: 10.1002/cne.903190204

PubMed Abstract | CrossRef Full Text | Google Scholar

Blanco-Centurion, C., Liu, M., Konadhode, R. P., Zhang, X., Pelluru, D., van den Pol, A. N., et al. (2016). Optogenetic activation of melanin-concentrating hormone neurons increases non-rapid eye movement and rapid eye movement sleep during the night in rats. Eur. J. Neurosci. 44, 2846–2857. doi: 10.1111/ejn.13410

PubMed Abstract | CrossRef Full Text | Google Scholar

Blanco-Centurion, C., Luo, S., Spergel, D. J., Vidal-Ortiz, A., Oprisan, S. A., Van den Pol, A. N., et al. (2019). Dynamic network activation of hypothalamic MCH neurons in REM sleep and exploratory behavior. J. Neurosci. 39, 4986–4998. doi: 10.1523/JNEUROSCI.0305-19.2019

PubMed Abstract | CrossRef Full Text | Google Scholar

Blouin, A. M., Thannickal, T. C., Worley, P. F., Baraban, J. M., Reti, I. M., and Siegel, J. M. (2005). Narp immunostaining of human hypocretin (orexin) neurons: Loss in narcolepsy. Neurology 65, 1189–1192. doi: 10.1212/01.wnl.0000175219.01544.c8

PubMed Abstract | CrossRef Full Text | Google Scholar

Boutrel, B., Kenny, P. J., Specio, S. E., Martin-Fardon, R., Markou, A., Koob, G. F., et al. (2005). Role for hypocretin in mediating stress-induced reinstatement of cocaine-seeking behavior. Proc. Natl. Acad. Sci. U. S. A. 102, 19168–19173.

Google Scholar

Broberger, C. (1999). Hypothalamic cocaine- and amphetamine-regulated transcript (CART) neurons: Histochemical relationship to thyrotropin-releasing hormone, melanin-concentrating hormone, orexin/hypocretin and neuropeptide Y. Brain Res. 848, 101–113. doi: 10.1016/s0006-8993(99)01977-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Broglio, F., Arvat, E., Benso, A., Gottero, C., Muccioli, G., Papotti, M., et al. (2001). Ghrelin, a natural GH secretagogue produced by the stomach, induces hyperglycemia and reduces insulin secretion in humans. J. Clin. Endocrinol. Metab. 86, 5083–5086. doi: 10.1210/jcem.86.10.8098

PubMed Abstract | CrossRef Full Text | Google Scholar

Bronsky, J., Nedvidkova, J., Krasnicanova, H., Vesela, M., Schmidtova, J., Koutek, J., et al. (2011). Changes of orexin A plasma levels in girls with anorexia nervosa during eight weeks of realimentation. Int. J. Eat Disord. 44, 547–552. doi: 10.1002/eat.20857

PubMed Abstract | CrossRef Full Text | Google Scholar

Brown, R. E., Sergeeva, O., Eriksson, K. S., and Haas, H. L. (2001). Orexin A excites serotonergic neurons in the dorsal raphe nucleus of the rat. Neuropharmacology 40, 457–459.

Google Scholar

Brundin, L., Bjorkqvist, M., Petersen, A., and Traskman-Bendz, L. (2007). Reduced orexin levels in the cerebrospinal fluid of suicidal patients with major depressive disorder. Eur. Neuropsychopharmacol. 17, 573–579.

Google Scholar

Buczek, L., Migliaccio, J., and Petrovich, G. D. (2020). Hedonic eating: Sex differences and characterization of orexin activation and signaling. Neuroscience 436, 34–45. doi: 10.1016/j.neuroscience.2020.04.008

PubMed Abstract | CrossRef Full Text | Google Scholar

Burdakov, D., Gerasimenko, O., and Verkhratsky, A. (2005). Physiological changes in glucose differentially modulate the excitability of hypothalamic melanin-concentrating hormone and orexin neurons in situ. J. Neurosci. 25, 2429–2433. doi: 10.1523/JNEUROSCI.4925-04.2005

PubMed Abstract | CrossRef Full Text | Google Scholar

Burdakov, D., Jensen, L. T., Alexopoulos, H., Williams, R. H., Fearon, I. M., O’Kelly, I., et al. (2006). Tandem-pore K+ channels mediate inhibition of orexin neurons by glucose. Neuron 50, 711–722.

Google Scholar

Burdakov, D., Liss, B., and Ashcroft, F. M. (2003). Orexin excites GABAergic neurons of the arcuate nucleus by activating the sodium–calcium exchanger. J. Neurosci. 23, 4951–4957. doi: 10.1523/JNEUROSCI.23-12-04951.2003

PubMed Abstract | CrossRef Full Text | Google Scholar

Calle, E. E., Rodriguez, C., Walker-Thurmond, K., and Thun, M. J. (2003). Overweight, obesity, and mortality from cancer in a prospectively studied cohort of U.S. adults. N. Engl. J. Med. 348, 1625–1638. doi: 10.1056/NEJMoa021423

PubMed Abstract | CrossRef Full Text | Google Scholar

Cao, Y., Tang, J., Yang, T., Ma, H., Yi, D., Gu, C., et al. (2013). Cardioprotective effect of ghrelin in cardiopulmonary bypass involves a reduction in inflammatory response. PLoS One 8:e55021. doi: 10.1371/journal.pone.0055021

PubMed Abstract | CrossRef Full Text | Google Scholar

Carlini, V. P., Monzon, M. E., Varas, M. M., Cragnolini, A. B., Schioth, H. B., Scimonelli, T. N., et al. (2002). Ghrelin increases anxiety-like behavior and memory retention in rats. Biochem. Biophys. Res. Commun. 299, 739–743.

Google Scholar

Carter, M. E., Adamantidis, A., Ohtsu, H., Deisseroth, K., and de Lecea, L. (2009). Sleep homeostasis modulates hypocretin-mediated sleep-to-wake transitions. J. Neurosci. 29, 10939–10949. doi: 10.1523/JNEUROSCI.1205-09.2009

PubMed Abstract | CrossRef Full Text | Google Scholar

Chee, M. J., Arrigoni, E., and Maratos-Flier, E. (2015). Melanin-concentrating hormone neurons release glutamate for feedforward inhibition of the lateral septum. J. Neurosci. 35, 3644–3651. doi: 10.1523/JNEUROSCI.4187-14.2015

PubMed Abstract | CrossRef Full Text | Google Scholar

Chemelli, R. M., Willie, J. T., Sinton, C. M., Elmquist, J. K., Scammell, T., Lee, C., et al. (1999). Narcolepsy in orexin knockout mice: Molecular genetics of sleep regulation. Cell 98, 437–451.

Google Scholar

Chen, R., Wu, X., Jiang, L., and Zhang, Y. (2017). Single-cell RNA-seq reveals hypothalamic cell diversity. Cell Rep. 18, 3227–3241.

Google Scholar

Chou, T. C., Lee, C. E., Lu, J., Elmquist, J. K., Hara, J., Willie, J. T., et al. (2001). Orexin (hypocretin) neurons contain dynorphin. J. Neurosci. 21:RC168.

Google Scholar

Chowdhury, S., Matsubara, T., Miyazaki, T., Ono, D., Fukatsu, N., Abe, M., et al. (2019). GABA neurons in the ventral tegmental area regulate non-rapid eye movement sleep in mice. Elife 8:e44928.

Google Scholar

Chung, S., Weber, F., Zhong, P., Tan, C. L., Nguyen, T. N., Beier, K. T., et al. (2017). Identification of preoptic sleep neurons using retrograde labelling and gene profiling. Nature 545, 477–481. doi: 10.1038/nature22350

PubMed Abstract | CrossRef Full Text | Google Scholar

Coborn, J. E., DePorter, D. P., Mavanji, V., Sinton, C. M., Kotz, C. M., Billington, C. J., et al. (2017). Role of orexin-A in the ventrolateral preoptic area on components of total energy expenditure. Int. J. Obes. 41, 1256–1262. doi: 10.1038/ijo.2017.92

PubMed Abstract | CrossRef Full Text | Google Scholar

Comninos, A. N., Jayasena, C. N., and Dhillo, W. S. (2014). The relationship between gut and adipose hormones, and reproduction. Hum. Reprod. Update 20, 153–174.

Google Scholar

Concetti, C., and Burdakov, D. (2021). Orexin/hypocretin and MCH neurons: Cognitive and motor roles beyond arousal. Front. Neurosci. 15:639313. doi: 10.3389/fnins.2021.639313

PubMed Abstract | CrossRef Full Text | Google Scholar

Conductier, G., Martin, A. O., Risold, P. Y., Jego, S., Lavoie, R., Lafont, C., et al. (2013). Control of ventricular ciliary beating by the melanin concentrating hormone-expressing neurons of the lateral hypothalamus: A functional imaging survey. Front. Endocrinol. 4:182. doi: 10.3389/fendo.2013.00182

PubMed Abstract | CrossRef Full Text | Google Scholar

Costa, A., Castro-Zaballa, S., Lagos, P., Chase, M. H., and Torterolo, P. (2018). Distribution of MCH-containing fibers in the feline brainstem: Relevance for REM sleep regulation. Peptides 104, 50–61. doi: 10.1016/j.peptides.2018.04.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Costa, J. L., Naot, D., Lin, J. M., Watson, M., Callon, K. E., Reid, I. R., et al. (2011). Ghrelin is an osteoblast mitogen and increases osteoclastic bone resorption In Vitro. Int. J. Pept. 2011:605193. doi: 10.1155/2011/605193

PubMed Abstract | CrossRef Full Text | Google Scholar

Cowley, M. A., Smith, R. G., Diano, S., Tschop, M., Pronchuk, N., Grove, K. L., et al. (2003). The distribution and mechanism of action of ghrelin in the CNS demonstrates a novel hypothalamic circuit regulating energy homeostasis. Neuron 37, 649–661. doi: 10.1016/s0896-6273(03)00063-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Crowley, V. E., Yeo, G. S., and O’Rahilly, S. (2002). Obesity therapy: Altering the energy intake-and-expenditure balance sheet. Nat. Rev. Drug Discov. 1, 276–286. doi: 10.1038/nrd770

PubMed Abstract | CrossRef Full Text | Google Scholar

Dalal, M. A., Schuld, A., and Pollmacher, T. (2003). Lower CSF orexin A (hypocretin-1) levels in patients with schizophrenia treated with haloperidol compared to unmedicated subjects. Mol. Psychiatry 8, 836–837. doi: 10.1038/sj.mp.4001363

PubMed Abstract | CrossRef Full Text | Google Scholar

Dantz, B., Edgar, D. M., and Dement, W. C. (1994). Circadian rhythms in narcolepsy: Studies on a 90 minute day. Electroencephalogr. Clin. Neurophysiol. 90, 24–35. doi: 10.1016/0013-4694(94)90110-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Date, Y., Ueta, Y., Yamashita, H., Yamaguchi, H., Matsukura, S., Kangawa, K., et al. (1999). Orexins, orexigenic hypothalamic peptides, interact with autonomic, neuroendocrine and neuroregulatory systems. Proc. Natl. Acad. Sci. U. S. A. 96, 748–753. doi: 10.1073/pnas.96.2.748

PubMed Abstract | CrossRef Full Text | Google Scholar

de Lecea, L., Kilduff, T. S., Peyron, C., Gao, X., Foye, P. E., Danielson, P. E., et al. (1998). The hypocretins: Hypothalamus-specific peptides with neuroexcitatory activity. Proc. Natl. Acad. Sci. U. S. A. 95, 322–327. doi: 10.1073/pnas.95.1.322

PubMed Abstract | CrossRef Full Text | Google Scholar

De Luca, R., Nardone, S., Grace, K. P., Venner, A., Cristofolini, M., Bandaru, S. S., et al. (2022). Orexin neurons inhibit sleep to promote arousal. Nat. Commun. 13:4163.

Google Scholar

Del Cid-Pellitero, E., and Jones, B. E. (2012). Immunohistochemical evidence for synaptic release of GABA from melanin-concentrating hormone containing varicosities in the locus coeruleus. Neuroscience 223, 269–276. doi: 10.1016/j.neuroscience.2012.07.072

PubMed Abstract | CrossRef Full Text | Google Scholar

Delgado, J. M., and Anand, B. K. (1953). Increase of food intake induced by electrical stimulation of the lateral hypothalamus. Am. J. Physiol. 172, 162–168.

Google Scholar

Delhanty, P. J., van der Eerden, B. C., and van Leeuwen, J. P. (2014). Ghrelin and bone. Biofactors 40, 41–48.

Google Scholar

Della-Zuana, O., Presse, F., Ortola, C., Duhault, J., Nahon, J. L., and Levens, N. (2002). Acute and chronic administration of melanin-concentrating hormone enhances food intake and body weight in Wistar and Sprague-Dawley rats. Int. J. Obes. Relat. Metab. Disord. 26, 1289–1295. doi: 10.1038/sj.ijo.0802079

PubMed Abstract | CrossRef Full Text | Google Scholar

Devera, A., Pascovich, C., Lagos, P., Falconi, A., Sampogna, S., Chase, M. H., et al. (2015). Melanin-concentrating hormone (MCH) modulates the activity of dorsal raphe neurons. Brain Res. 1598, 114–128.

Google Scholar

Dezaki, K., Hosoda, H., Kakei, M., Hashiguchi, S., Watanabe, M., Kangawa, K., et al. (2004). Endogenous ghrelin in pancreatic islets restricts insulin release by attenuating Ca2+ signaling in beta-cells: Implication in the glycemic control in rodents. Diabetes 53, 3142–3151. doi: 10.2337/diabetes.53.12.3142

PubMed Abstract | CrossRef Full Text | Google Scholar

Diano, S., Farr, S. A., Benoit, S. C., McNay, E. C., da Silva, I., Horvath, B., et al. (2006). Ghrelin controls hippocampal spine synapse density and memory performance. Nat. Neurosci. 9, 381–388. doi: 10.1038/nn1656

PubMed Abstract | CrossRef Full Text | Google Scholar

Diano, S., Horvath, B., Urbanski, H. F., Sotonyi, P., and Horvath, T. L. (2003). Fasting activates the nonhuman primate hypocretin (orexin) system and its postsynaptic targets. Endocrinology 144, 3774–3778. doi: 10.1210/en.2003-0274

PubMed Abstract | CrossRef Full Text | Google Scholar

Donga, E., van Dijk, M., van Dijk, J. G., Biermasz, N. R., Lammers, G. J., van Kralingen, K. W., et al. (2010). A single night of partial sleep deprivation induces insulin resistance in multiple metabolic pathways in healthy subjects. J. Clin. Endocrinol. Metab. 95, 2963–2968. doi: 10.1210/jc.2009-2430

PubMed Abstract | CrossRef Full Text | Google Scholar

Druce, M. R., Wren, A. M., Park, A. J., Milton, J. E., Patterson, M., Frost, G., et al. (2005). Ghrelin increases food intake in obese as well as lean subjects. Int. J. Obes. 29, 1130–1136.

Google Scholar

Dube, M. G., Horvath, T. L., Kalra, P. S., and Kalra, S. P. (2000). Evidence of NPY Y5 receptor involvement in food intake elicited by orexin A in sated rats. Peptides 21, 1557–1560. doi: 10.1016/s0196-9781(00)00311-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Dube, M. G., Kalra, S. P., and Kalra, P. S. (1999). Food intake elicited by central administration of orexins/hypocretins: Identification of hypothalamic sites of action. Brain Res. 842, 473–477. doi: 10.1016/s0006-8993(99)01824-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Edwards, C. M., Abusnana, S., Sunter, D., Murphy, K. G., Ghatei, M. A., and Bloom, S. R. (1999). The effect of the orexins on food intake: Comparison with neuropeptide Y, melanin-concentrating hormone and galanin. J. Endocrinol. 160, R7–R12. doi: 10.1677/joe.0.160r007

PubMed Abstract | CrossRef Full Text | Google Scholar

Eggermann, E., Serafin, M., Bayer, L., Machard, D., Saint-Mleux, B., Jones, B. E., et al. (2001). Orexins/hypocretins excite basal forebrain cholinergic neurones. Neuroscience 108, 177–181. doi: 10.1016/s0306-4522(01)00512-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Elias, C. F., Lee, C. E., Kelly, J. F., Ahima, R. S., Kuhar, M., Saper, C. B., et al. (2001). Characterization of CART neurons in the rat and human hypothalamus. J. Comp. Neurol. 432, 1–19.

Google Scholar

Eriksson, K. S., Sergeeva, O., Brown, R. E., and Haas, H. L. (2001). Orexin/hypocretin excites the histaminergic neurons of the tuberomammillary nucleus. J. Neurosci. 21, 9273–9279.

Google Scholar

Espana, R. A., Melchior, J. R., Roberts, D. C., and Jones, S. R. (2011). Hypocretin 1/orexin A in the ventral tegmental area enhances dopamine responses to cocaine and promotes cocaine self-administration. Psychopharmacology 214, 415–426. doi: 10.1007/s00213-010-2048-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Espana, R. A., Plahn, S., and Berridge, C. W. (2002). Circadian-dependent and circadian-independent behavioral actions of hypocretin/orexin. Brain Res. 943, 224–236. doi: 10.1016/s0006-8993(02)02653-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Fort, P., Salvert, D., Hanriot, L., Jego, S., Shimizu, H., Hashimoto, K., et al. (2008). The satiety molecule nesfatin-1 is co-expressed with melanin concentrating hormone in tuberal hypothalamic neurons of the rat. Neuroscience 155, 174–181.

Google Scholar

Fronczek, R., Baumann, C. R., Lammers, G. J., Bassetti, C. L., and Overeem, S. (2009). Hypocretin/orexin disturbances in neurological disorders. Sleep Med. Rev. 13, 9–22.

Google Scholar

Fujiki, N., Yoshida, Y., Zhang, S., Sakurai, T., Yanagisawa, M., and Nishino, S. (2006). Sex difference in body weight gain and leptin signaling in hypocretin/orexin deficient mouse models. Peptides 27, 2326–2331. doi: 10.1016/j.peptides.2006.03.011

PubMed Abstract | CrossRef Full Text | Google Scholar

Fukunaka, Y., Shinkai, T., Hwang, R., Hori, H., Utsunomiya, K., Sakata, S., et al. (2007). The orexin 1 receptor (HCRTR1) gene as a susceptibility gene contributing to polydipsia-hyponatremia in schizophrenia. Neuromol. Med. 9, 292–297. doi: 10.1007/s12017-007-8001-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Funahashi, H., Yamada, S., Kageyama, H., Takenoya, F., Guan, J. L., and Shioda, S. (2003). Co-existence of leptin- and orexin-receptors in feeding-regulating neurons in the hypothalamic arcuate nucleus-a triple labeling study. Peptides 24, 687–694. doi: 10.1016/s0196-9781(03)00130-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Gautron, L., Elmquist, J. K., and Williams, K. W. (2015). Neural control of energy balance: Translating circuits to therapies. Cell 161, 133–145. doi: 10.1016/j.cell.2015.02.023

PubMed Abstract | CrossRef Full Text | Google Scholar

Georgescu, D., Zachariou, V., Barrot, M., Mieda, M., Willie, J. T., Eisch, A. J., et al. (2003). Involvement of the lateral hypothalamic peptide orexin in morphine dependence and withdrawal. J. Neurosci. 23, 3106–3111.

Google Scholar

Glick, M., Segal-Lieberman, G., Cohen, R., and Kronfeld-Schor, N. (2009). Chronic MCH infusion causes a decrease in energy expenditure and body temperature, and an increase in serum IGF-1 levels in mice. Endocrine 36, 479–485. doi: 10.1007/s12020-009-9252-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Gnanapavan, S., Kola, B., Bustin, S. A., Morris, D. G., McGee, P., Fairclough, P., et al. (2002). The tissue distribution of the mRNA of ghrelin and subtypes of its receptor, GHS-R, in humans. J. Clin. Endocrinol. Metab. 87:2988. doi: 10.1210/jcem.87.6.8739

PubMed Abstract | CrossRef Full Text | Google Scholar

Goforth, P. B., Leinninger, G. M., Patterson, C. M., Satin, L. S., and Myers, M. G. Jr. (2014). Leptin acts via lateral hypothalamic area neurotensin neurons to inhibit orexin neurons by multiple GABA-independent mechanisms. J. Neurosci. 34, 11405–11415. doi: 10.1523/JNEUROSCI.5167-13.2014

PubMed Abstract | CrossRef Full Text | Google Scholar

Gomori, A., Ishihara, A., Ito, M., Mashiko, S., Matsushita, H., Yumoto, M., et al. (2003). Chronic intracerebroventricular infusion of MCH causes obesity in mice. Melanin-concentrating hormone. Am. J. Physiol. Endocrinol. Metab. 284, E583–E588.

Google Scholar

Gonzalez, J. A., Jensen, L. T., Fugger, L., and Burdakov, D. (2008). Metabolism-independent sugar sensing in central orexin neurons. Diabetes 57, 2569–2576. doi: 10.2337/db08-0548

PubMed Abstract | CrossRef Full Text | Google Scholar

Gonzalez, J. A., Reimann, F., and Burdakov, D. (2009). Dissociation between sensing and metabolism of glucose in sugar sensing neurones. J. Physiol. 587, 41–48.

Google Scholar

Guan, J. L., Saotome, T., Wang, Q. P., Funahashi, H., Hori, T., Tanaka, S., et al. (2001). Orexinergic innervation of POMC-containing neurons in the rat arcuate nucleus. Neuroreport 12, 547–551. doi: 10.1097/00001756-200103050-00023

PubMed Abstract | CrossRef Full Text | Google Scholar

Hagan, J. J., Leslie, R. A., Patel, S., Evans, M. L., Wattam, T. A., Holmes, S., et al. (1999). Orexin A activates locus coeruleus cell firing and increases arousal in the rat. Proc. Natl. Acad. Sci. U. S. A. 96, 10911–10916.

Google Scholar

Hakansson, M., de Lecea, L., Sutcliffe, J. G., Yanagisawa, M., and Meister, B. (1999). Leptin receptor- and STAT3-immunoreactivities in hypocretin/orexin neurones of the lateral hypothalamus. J. Neuroendocrinol. 11, 653–663. doi: 10.1046/j.1365-2826.1999.00378.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Hanriot, L., Camargo, N., Courau, A. C., Leger, L., Luppi, P. H., and Peyron, C. (2007). Characterization of the melanin-concentrating hormone neurons activated during paradoxical sleep hypersomnia in rats. J. Comp. Neurol. 505, 147–157.

Google Scholar

Hara, J., Beuckmann, C. T., Nambu, T., Willie, J. T., Chemelli, R. M., Sinton, C. M., et al. (2001). Genetic ablation of orexin neurons in mice results in narcolepsy, hypophagia, and obesity. Neuron 30, 345–354.

Google Scholar

Hara, J., Yanagisawa, M., and Sakurai, T. (2005). Difference in obesity phenotype between orexin-knockout mice and orexin neuron-deficient mice with same genetic background and environmental conditions. Neurosci. Lett. 380, 239–242. doi: 10.1016/j.neulet.2005.01.046

PubMed Abstract | CrossRef Full Text | Google Scholar

Harris, G. C., Wimmer, M., and Aston-Jones, G. (2005). A role for lateral hypothalamic orexin neurons in reward seeking. Nature 437, 556–559.

Google Scholar

Harthoorn, L. F., Sane, A., Nethe, M., and Van Heerikhuize, J. J. (2005). Multi-transcriptional profiling of melanin-concentrating hormone and orexin-containing neurons. Cell Mol. Neurobiol. 25, 1209–1223. doi: 10.1007/s10571-005-8184-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Hassani, O. K., Lee, M. G., and Jones, B. E. (2009). Melanin-concentrating hormone neurons discharge in a reciprocal manner to orexin neurons across the sleep-wake cycle. Proc. Natl. Acad. Sci. U. S. A. 106, 2418–2422. doi: 10.1073/pnas.0811400106

PubMed Abstract | CrossRef Full Text | Google Scholar

Hausen, A. C., Ruud, J., Jiang, H., Hess, S., Varbanov, H., Kloppenburg, P., et al. (2016). Insulin-dependent activation of MCH neurons impairs locomotor activity and insulin sensitivity in obesity. Cell Rep. 17, 2512–2521. doi: 10.1016/j.celrep.2016.11.030

PubMed Abstract | CrossRef Full Text | Google Scholar

Haynes, A. C., Jackson, B., Chapman, H., Tadayyon, M., Johns, A., Porter, R. A., et al. (2000). A selective orexin-1 receptor antagonist reduces food consumption in male and female rats. Regul. Pept. 96, 45–51.

Google Scholar

Haynes, A. C., Jackson, B., Overend, P., Buckingham, R. E., Wilson, S., Tadayyon, M., et al. (1999). Effects of single and chronic intracerebroventricular administration of the orexins on feeding in the rat. Peptides 20, 1099–1105. doi: 10.1016/s0196-9781(99)00105-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Henny, P., Brischoux, F., Mainville, L., Stroh, T., and Jones, B. E. (2010). Immunohistochemical evidence for synaptic release of glutamate from orexin terminals in the locus coeruleus. Neuroscience 169, 1150–1157. doi: 10.1016/j.neuroscience.2010.06.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Herrera, C. G., Cadavieco, M. C., Jego, S., Ponomarenko, A., Korotkova, T., and Adamantidis, A. (2016). Hypothalamic feedforward inhibition of thalamocortical network controls arousal and consciousness. Nat. Neurosci. 19, 290–298. doi: 10.1038/nn.4209

PubMed Abstract | CrossRef Full Text | Google Scholar

Horvath, T. L., and Gao, X. B. (2005). Input organization and plasticity of hypocretin neurons: Possible clues to obesity’s association with insomnia. Cell Metab. 1, 279–286. doi: 10.1016/j.cmet.2005.03.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Horvath, T. L., Diano, S., and van den Pol, A. N. (1999a). Synaptic interaction between hypocretin (orexin) and neuropeptide Y cells in the rodent and primate hypothalamus: A novel circuit implicated in metabolic and endocrine regulations. J. Neurosci. 19, 1072–1087. doi: 10.1523/JNEUROSCI.19-03-01072.1999

PubMed Abstract | CrossRef Full Text | Google Scholar

Horvath, T. L., Peyron, C., Diano, S., Ivanov, A., Aston-Jones, G., Kilduff, T. S., et al. (1999b). Hypocretin (orexin) activation and synaptic innervation of the locus coeruleus noradrenergic system. J. Comp. Neurol. 415, 145–159.

Google Scholar

Inutsuka, A., Inui, A., Tabuchi, S., Tsunematsu, T., Lazarus, M., and Yamanaka, A. (2014). Concurrent and robust regulation of feeding behaviors and metabolism by orexin neurons. Neuropharmacology 85, 451–460. doi: 10.1016/j.neuropharm.2014.06.015

PubMed Abstract | CrossRef Full Text | Google Scholar

Iqbal, J., Pompolo, S., Murakami, T., Grouzmann, E., Sakurai, T., Meister, B., et al. (2001). Immunohistochemical characterization of localization of long-form leptin receptor (OB-Rb) in neurochemically defined cells in the ovine hypothalamus. Brain Res. 920, 55–64. doi: 10.1016/s0006-8993(01)02932-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Ito, M., Gomori, A., Ishihara, A., Oda, Z., Mashiko, S., Matsushita, H., et al. (2003). Characterization of MCH-mediated obesity in mice. Am. J. Physiol. Endocrinol. Metab. 284, E940–E945.

Google Scholar

Ivanov, A., and Aston-Jones, G. (2000). Hypocretin/orexin depolarizes and decreases potassium conductance in locus coeruleus neurons. Neuroreport 11, 1755–1758. doi: 10.1097/00001756-200006050-00031

PubMed Abstract | CrossRef Full Text | Google Scholar

Izawa, S., Chowdhury, S., Miyazaki, T., Mukai, Y., Ono, D., Inoue, R., et al. (2019). REM sleep-active MCH neurons are involved in forgetting hippocampus-dependent memories. Science 365, 1308–1313. doi: 10.1126/science.aax9238

PubMed Abstract | CrossRef Full Text | Google Scholar

Izawa, S., Yoneshiro, T., Kondoh, K., Nakagiri, S., Okamatsu-Ogura, Y., Terao, A., et al. (2022). Melanin-concentrating hormone-producing neurons in the hypothalamus regulate brown adipose tissue and thus contribute to energy expenditure. J. Physiol. 600, 815–827. doi: 10.1113/JP281241

PubMed Abstract | CrossRef Full Text | Google Scholar

Jain, M. R., Horvath, T. L., Kalra, P. S., and Kalra, S. P. (2000). Evidence that NPY Y1 receptors are involved in stimulation of feeding by orexins (hypocretins) in sated rats. Regul. Pept. 87, 19–24.

Google Scholar

Janas-Kozik, M., Stachowicz, M., Krupka-Matuszczyk, I., Szymszal, J., Krysta, K., Janas, A., et al. (2011). Plasma levels of leptin and orexin A in the restrictive type of anorexia nervosa. Regul. Pept. 168, 5–9. doi: 10.1016/j.regpep.2011.02.005

PubMed Abstract | CrossRef Full Text | Google Scholar

Jego, S., Glasgow, S. D., Herrera, C. G., Ekstrand, M., Reed, S. J., Boyce, R., et al. (2013). Optogenetic identification of a rapid eye movement sleep modulatory circuit in the hypothalamus. Nat. Neurosci. 16, 1637–1643. doi: 10.1038/nn.3522

PubMed Abstract | CrossRef Full Text | Google Scholar

Jennings, J. H., Ung, R. L., Resendez, S. L., Stamatakis, A. M., Taylor, J. G., Huang, J., et al. (2015). Visualizing hypothalamic network dynamics for appetitive and consummatory behaviors. Cell 160, 516–527. doi: 10.1016/j.cell.2014.12.026

PubMed Abstract | CrossRef Full Text | Google Scholar

Jeon, J. Y., Bradley, R. L., Kokkotou, E. G., Marino, F. E., Wang, X., Pissios, P., et al. (2006). MCH-/- mice are resistant to aging-associated increases in body weight and insulin resistance. Diabetes 55, 428–434.

Google Scholar

Jha, P. K., Foppen, E., Kalsbeek, A., and Challet, E. (2016). Sleep restriction acutely impairs glucose tolerance in rats. Physiol. Rep. 4:e12839.

Google Scholar

Jha, P. K., Valekunja, U. K., Ray, S., Nollet, M., and Reddy, A. B. (2022). Single-cell transcriptomics and cell-specific proteomics reveals molecular signatures of sleep. Commun. Biol. 5:846. doi: 10.1038/s42003-022-03800-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Jiang, H., Gallet, S., Klemm, P., Scholl, P., Folz-Donahue, K., Altmuller, J., et al. (2020). MCH neurons regulate permeability of the median eminence barrier. Neuron 107, 306–319.e309. doi: 10.1016/j.neuron.2020.04.020

PubMed Abstract | CrossRef Full Text | Google Scholar

Jin, T., Jiang, Z., Luan, X., Qu, Z., Guo, F., Gao, S., et al. (2020). Exogenous orexin-A microinjected into central nucleus of the amygdala modulates feeding and gastric motility in rats. Front. Neurosci. 14:274. doi: 10.3389/fnins.2020.00274

PubMed Abstract | CrossRef Full Text | Google Scholar

Johansson, L., Ekholm, M. E., and Kukkonen, J. P. (2008). Multiple phospholipase activation by OX(1) orexin/hypocretin receptors. Cell Mol. Life Sci. 65, 1948–1956. doi: 10.1007/s00018-008-8206-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Kanoski, S. E., Fortin, S. M., Ricks, K. M., and Grill, H. J. (2013). Ghrelin signaling in the ventral hippocampus stimulates learned and motivational aspects of feeding via PI3K-Akt signaling. Biol. Psychiatry 73, 915–923. doi: 10.1016/j.biopsych.2012.07.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Khazaei, M., and Tahergorabi, Z. (2013). Systemic ghrelin administration alters serum biomarkers of angiogenesis in diet-induced obese mice. Int. J. Pept. 2013:249565.

Google Scholar

Kim, S. H., Despres, J. P., and Koh, K. K. (2016). Obesity and cardiovascular disease: Friend or foe? Eur. Heart J. 37, 3560–3568.

Google Scholar

Kitka, T., Adori, C., Katai, Z., Vas, S., Molnar, E., Papp, R. S., et al. (2011). Association between the activation of MCH and orexin immunorective neurons and REM sleep architecture during REM rebound after a three day long REM deprivation. Neurochem. Int. 59, 686–694. doi: 10.1016/j.neuint.2011.06.015

PubMed Abstract | CrossRef Full Text | Google Scholar

Kiyashchenko, L. I., Mileykovskiy, B. Y., Maidment, N., Lam, H. A., Wu, M. F., John, J., et al. (2002). Release of hypocretin (orexin) during waking and sleep states. J. Neurosci. 22, 5282–5286.

Google Scholar

Kojima, M., Hosoda, H., Date, Y., Nakazato, M., Matsuo, H., and Kangawa, K. (1999). Ghrelin is a growth-hormone-releasing acylated peptide from stomach. Nature 402, 656–660.

Google Scholar

Kokkotou, E., Jeon, J. Y., Wang, X., Marino, F. E., Carlson, M., Trombly, D. J., et al. (2005). Mice with MCH ablation resist diet-induced obesity through strain-specific mechanisms. Am. J. Physiol. Regul. Integr. Comp. Physiol. 289, R117–R124. doi: 10.1152/ajpregu.00861.2004

PubMed Abstract | CrossRef Full Text | Google Scholar

Konadhode, R. R., Pelluru, D., Blanco-Centurion, C., Zayachkivsky, A., Liu, M., Uhde, T., et al. (2013). Optogenetic stimulation of MCH neurons increases sleep. J. Neurosci. 33, 10257–10263.

Google Scholar

Kong, D., Vong, L., Parton, L. E., Ye, C., Tong, Q., Hu, X., et al. (2010). Glucose stimulation of hypothalamic MCH neurons involves K(ATP) channels, is modulated by UCP2, and regulates peripheral glucose homeostasis. Cell Metab. 12, 545–552. doi: 10.1016/j.cmet.2010.09.013

PubMed Abstract | CrossRef Full Text | Google Scholar

Krauss, S., Zhang, C. Y., Scorrano, L., Dalgaard, L. T., St-Pierre, J., Grey, S. T., et al. (2003). Superoxide-mediated activation of uncoupling protein 2 causes pancreatic beta cell dysfunction. J. Clin. Invest. 112, 1831–1842. doi: 10.1172/JCI19774

PubMed Abstract | CrossRef Full Text | Google Scholar

Kroeger, D., Bandaru, S. S., Madara, J. C., and Vetrivelan, R. (2019). Ventrolateral periaqueductal gray mediates rapid eye movement sleep regulation by melanin-concentrating hormone neurons. Neuroscience 406, 314–324. doi: 10.1016/j.neuroscience.2019.03.020

PubMed Abstract | CrossRef Full Text | Google Scholar

Lagos, P., Monti, J. M., Jantos, H., and Torterolo, P. (2012). Microinjection of the melanin-concentrating hormone into the lateral basal forebrain increases REM sleep and reduces wakefulness in the rat. Life Sci. 90, 895–899. doi: 10.1016/j.lfs.2012.04.019

PubMed Abstract | CrossRef Full Text | Google Scholar

Lagos, P., Torterolo, P., Jantos, H., Chase, M. H., and Monti, J. M. (2009). Effects on sleep of melanin-concentrating hormone (MCH) microinjections into the dorsal raphe nucleus. Brain Res. 1265, 103–110.

Google Scholar

Laque, A., Zhang, Y., Gettys, S., Nguyen, T. A., Bui, K., Morrison, C. D., et al. (2013). Leptin receptor neurons in the mouse hypothalamus are colocalized with the neuropeptide galanin and mediate anorexigenic leptin action. Am. J. Physiol. Endocrinol. Metab. 304, E999–E1011. doi: 10.1152/ajpendo.00643.2012

PubMed Abstract | CrossRef Full Text | Google Scholar

Lawrence, C. B., Snape, A. C., Baudoin, F. M., and Luckman, S. M. (2002). Acute central ghrelin and GH secretagogues induce feeding and activate brain appetite centers. Endocrinology 143, 155–162. doi: 10.1210/endo.143.1.8561

PubMed Abstract | CrossRef Full Text | Google Scholar

Lee, J., Lim, E., Kim, Y., Li, E., and Park, S. (2010). Ghrelin attenuates kainic acid-induced neuronal cell death in the mouse hippocampus. J. Endocrinol. 205, 263–270. doi: 10.1677/JOE-10-0040

PubMed Abstract | CrossRef Full Text | Google Scholar

Lee, M. G., Hassani, O. K., and Jones, B. E. (2005). Discharge of identified orexin/hypocretin neurons across the sleep-waking cycle. J. Neurosci. 25, 6716–6720. doi: 10.1523/JNEUROSCI.1887-05.2005

PubMed Abstract | CrossRef Full Text | Google Scholar

Leinninger, G. M., Jo, Y. H., Leshan, R. L., Louis, G. W., Yang, H., Barrera, J. G., et al. (2009). Leptin acts via leptin receptor-expressing lateral hypothalamic neurons to modulate the mesolimbic dopamine system and suppress feeding. Cell Metab. 10, 89–98. doi: 10.1016/j.cmet.2009.06.011

PubMed Abstract | CrossRef Full Text | Google Scholar

Leinninger, G. M., Opland, D. M., Jo, Y. H., Faouzi, M., Christensen, L., Cappellucci, L. A., et al. (2011). Leptin action via neurotensin neurons controls orexin, the mesolimbic dopamine system and energy balance. Cell Metab. 14, 313–323.

Google Scholar

Li, Y., Li, S., Wei, C., Wang, H., Sui, N., and Kirouac, G. J. (2010). Orexins in the paraventricular nucleus of the thalamus mediate anxiety-like responses in rats. Psychopharmacology 212, 251–265. doi: 10.1007/s00213-010-1948-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Lin, L., Faraco, J., Li, R., Kadotani, H., Rogers, W., Lin, X., et al. (1999). The sleep disorder canine narcolepsy is caused by a mutation in the hypocretin (orexin) receptor 2 gene. Cell 98, 365–376.

Google Scholar

Lin, L., Saha, P. K., Ma, X., Henshaw, I. O., Shao, L., Chang, B. H., et al. (2011). Ablation of ghrelin receptor reduces adiposity and improves insulin sensitivity during aging by regulating fat metabolism in white and brown adipose tissues. Aging Cell 10, 996–1010. doi: 10.1111/j.1474-9726.2011.00740.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Linehan, V., and Hirasawa, M. (2022). Short-term fasting induces alternate activation of orexin and melanin-concentrating hormone neurons in rats. Neuroscience 491, 156–165. doi: 10.1016/j.neuroscience.2022.04.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Lopez, M., Seoane, L. M., Garcia Mdel, C., Dieguez, C., and Senaris, R. (2002). Neuropeptide Y, but not agouti-related peptide or melanin-concentrating hormone, is a target peptide for orexin-A feeding actions in the rat hypothalamus. Neuroendocrinology 75, 34–44.

Google Scholar

Lopez, M., Seoane, L., Garcia, M. C., Lago, F., Casanueva, F. F., Senaris, R., et al. (2000). Leptin regulation of prepro-orexin and orexin receptor mRNA levels in the hypothalamus. Biochem. Biophys. Res. Commun. 269, 41–45.

Google Scholar

Lord, M. N., Subramanian, K., Kanoski, S. E., and Noble, E. E. (2021). Melanin-concentrating hormone and food intake control: Sites of action, peptide interactions, and appetition. Peptides 137:170476. doi: 10.1016/j.peptides.2020.170476

PubMed Abstract | CrossRef Full Text | Google Scholar

Louis, G. W., Leinninger, G. M., Rhodes, C. J., and Myers, M. G. Jr. (2010). Direct innervation and modulation of orexin neurons by lateral hypothalamic LepRb neurons. J. Neurosci. 30, 11278–11287.

Google Scholar

Ludwig, D. S., Tritos, N. A., Mastaitis, J. W., Kulkarni, R., Kokkotou, E., Elmquist, J., et al. (2001). Melanin-concentrating hormone overexpression in transgenic mice leads to obesity and insulin resistance. J. Clin. Invest. 107, 379–386. doi: 10.1172/JCI10660

PubMed Abstract | CrossRef Full Text | Google Scholar

Lungwitz, E. A., Molosh, A., Johnson, P. L., Harvey, B. P., Dirks, R. C., Dietrich, A., et al. (2012). Orexin-A induces anxiety-like behavior through interactions with glutamatergic receptors in the bed nucleus of the stria terminalis of rats. Physiol. Behav. 107, 726–732. doi: 10.1016/j.physbeh.2012.05.019

PubMed Abstract | CrossRef Full Text | Google Scholar

Luppi, P. H., Clement, O., and Fort, P. (2013). Paradoxical (REM) sleep genesis by the brainstem is under hypothalamic control. Curr. Opin. Neurobiol. 23, 786–792. doi: 10.1016/j.conb.2013.02.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Ma, X., Zubcevic, L., Bruning, J. C., Ashcroft, F. M., and Burdakov, D. (2007). Electrical inhibition of identified anorexigenic POMC neurons by orexin/hypocretin. J. Neurosci. 27, 1529–1533. doi: 10.1523/JNEUROSCI.3583-06.2007

PubMed Abstract | CrossRef Full Text | Google Scholar

Malik, S., McGlone, F., Bedrossian, D., and Dagher, A. (2008). Ghrelin modulates brain activity in areas that control appetitive behavior. Cell Metab. 7, 400–409. doi: 10.1016/j.cmet.2008.03.007

PubMed Abstract | CrossRef Full Text | Google Scholar

Mao, Y., Tokudome, T., Otani, K., Kishimoto, I., Miyazato, M., and Kangawa, K. (2013). Excessive sympathoactivation and deteriorated heart function after myocardial infarction in male ghrelin knockout mice. Endocrinology 154, 1854–1863. doi: 10.1210/en.2012-2132

PubMed Abstract | CrossRef Full Text | Google Scholar

Mao, Y., Tokudome, T., Otani, K., Kishimoto, I., Nakanishi, M., Hosoda, H., et al. (2012). Ghrelin prevents incidence of malignant arrhythmia after acute myocardial infarction through vagal afferent nerves. Endocrinology 153, 3426–3434. doi: 10.1210/en.2012-1065

PubMed Abstract | CrossRef Full Text | Google Scholar

Marsh, D. J., Weingarth, D. T., Novi, D. E., Chen, H. Y., Trumbauer, M. E., Chen, A. S., et al. (2002). Melanin-concentrating hormone 1 receptor-deficient mice are lean, hyperactive, and hyperphagic and have altered metabolism. Proc. Natl. Acad. Sci. U. S. A. 99, 3240–3245. doi: 10.1073/pnas.052706899

PubMed Abstract | CrossRef Full Text | Google Scholar

Mavanji, V., Perez-Leighton, C. E., Kotz, C. M., Billington, C. J., Parthasarathy, S., Sinton, C. M., et al. (2015). Promotion of wakefulness and energy expenditure by orexin-A in the ventrolateral preoptic area. Sleep 38, 1361–1370.

Google Scholar

Mavanji, V., Pomonis, B., and Kotz, C. M. (2022). Orexin, serotonin, and energy balance. Wires Mech. Dis. 14:e1536.

Google Scholar

Meerabux, J., Iwayama, Y., Sakurai, T., Ohba, H., Toyota, T., Yamada, K., et al. (2005). Association of an orexin 1 receptor 408Val variant with polydipsia-hyponatremia in schizophrenic subjects. Biol. Psychiatry 58, 401–407. doi: 10.1016/j.biopsych.2005.04.015

PubMed Abstract | CrossRef Full Text | Google Scholar

Miao, H., Pan, H., Wang, L., Yang, H., Zhu, H., and Gong, F. (2019). Ghrelin promotes proliferation and inhibits differentiation of 3T3-L1 and human primary preadipocytes. Front. Physiol. 10:1296. doi: 10.3389/fphys.2019.01296

PubMed Abstract | CrossRef Full Text | Google Scholar

Mickelsen, L. E., Bolisetty, M., Chimileski, B. R., Fujita, A., Beltrami, E. J., Costanzo, J. T., et al. (2019). Single-cell transcriptomic analysis of the lateral hypothalamic area reveals molecularly distinct populations of inhibitory and excitatory neurons. Nat. Neurosci. 22, 642–656. doi: 10.1038/s41593-019-0349-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Mickelsen, L. E., Kolling, F. W. T., Chimileski, B. R., Fujita, A., Norris, C., Chen, K., et al. (2017). Neurochemical heterogeneity among lateral hypothalamic hypocretin/orexin and melanin-concentrating hormone neurons identified through single-cell gene expression analysis. eNeuro 4, ENEURO.0013–ENEURO.17. doi: 10.1523/ENEURO.0013-17.2017

PubMed Abstract | CrossRef Full Text | Google Scholar

Mieda, M., Hasegawa, E., Kisanuki, Y. Y., Sinton, C. M., Yanagisawa, M., and Sakurai, T. (2011). Differential roles of orexin receptor-1 and -2 in the regulation of non-REM and REM sleep. J. Neurosci. 31, 6518–6526. doi: 10.1523/JNEUROSCI.6506-10.2011

PubMed Abstract | CrossRef Full Text | Google Scholar

Milbank, E., and Lopez, M. (2019). Orexins/hypocretins: Key regulators of energy homeostasis. Front. Endocrinol. 10:830. doi: 10.3389/fendo.2019.00830

PubMed Abstract | CrossRef Full Text | Google Scholar

Mileykovskiy, B. Y., Kiyashchenko, L. I., and Siegel, J. M. (2005). Behavioral correlates of activity in identified hypocretin/orexin neurons. Neuron 46, 787–798.

Google Scholar

Mochizuki, T., Arrigoni, E., Marcus, J. N., Clark, E. L., Yamamoto, M., Honer, M., et al. (2011). Orexin receptor 2 expression in the posterior hypothalamus rescues sleepiness in narcoleptic mice. Proc. Natl. Acad. Sci. U. S. A. 108, 4471–4476. doi: 10.1073/pnas.1012456108

PubMed Abstract | CrossRef Full Text | Google Scholar

Mochizuki, T., Crocker, A., McCormack, S., Yanagisawa, M., Sakurai, T., and Scammell, T. E. (2004). Behavioral state instability in orexin knock-out mice. J. Neurosci. 24, 6291–6300.

Google Scholar

Modirrousta, M., Mainville, L., and Jones, B. E. (2005). Orexin and MCH neurons express c-Fos differently after sleep deprivation vs. recovery and bear different adrenergic receptors. Eur. J. Neurosci. 21, 2807–2816. doi: 10.1111/j.1460-9568.2005.04104.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Mogi, K., Funabashi, T., Mitsushima, D., Hagiwara, H., and Kimura, F. (2005). Sex difference in the response of melanin-concentrating hormone neurons in the lateral hypothalamic area to glucose, as revealed by the expression of phosphorylated cyclic adenosine 3’,5’-monophosphate response element-binding protein. Endocrinology 146, 3325–3333. doi: 10.1210/en.2005-0078

PubMed Abstract | CrossRef Full Text | Google Scholar

Monti, J. M., Lagos, P., Jantos, H., and Torterolo, P. (2015). Increased REM sleep after intra-locus coeruleus nucleus microinjection of melanin-concentrating hormone (MCH) in the rat. Prog. Neuropsychopharmacol. Biol. Psychiatry 56, 185–188. doi: 10.1016/j.pnpbp.2014.09.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Monti, J. M., Torterolo, P., and Lagos, P. (2013). Melanin-concentrating hormone control of sleep-wake behavior. Sleep Med. Rev. 17, 293–298.

Google Scholar

Monti, J. M., Torterolo, P., Jantos, H., and Lagos, P. (2016). Microinjection of the melanin-concentrating hormone into the sublaterodorsal tegmental nucleus inhibits REM sleep in the rat. Neurosci. Lett. 630, 66–69. doi: 10.1016/j.neulet.2016.07.035

PubMed Abstract | CrossRef Full Text | Google Scholar

Morton, G. J., Meek, T. H., and Schwartz, M. W. (2014). Neurobiology of food intake in health and disease. Nat. Rev. Neurosci. 15, 367–378.

Google Scholar

Mosavat, M., Mirsanjari, M., Arabiat, D., Smyth, A., and Whitehead, L. (2021). The role of sleep curtailment on leptin levels in obesity and diabetes mellitus. Obes. Facts 14, 214–221.

Google Scholar

Muroya, S., Funahashi, H., Yamanaka, A., Kohno, D., Uramura, K., Nambu, T., et al. (2004). Orexins (hypocretins) directly interact with neuropeptide Y, POMC and glucose-responsive neurons to regulate Ca 2+ signaling in a reciprocal manner to leptin: Orexigenic neuronal pathways in the mediobasal hypothalamus. Eur. J. Neurosci. 19, 1524–1534. doi: 10.1111/j.1460-9568.2004.03255.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Murray, J. F., Mercer, J. G., Adan, R. A., Datta, J. J., Aldairy, C., Moar, K. M., et al. (2000). The effect of leptin on luteinizing hormone release is exerted in the zona incerta and mediated by melanin-concentrating hormone. J. Neuroendocrinol. 12, 1133–1139. doi: 10.1046/j.1365-2826.2000.00577.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Myers, M. G. Jr., and Olson, D. P. (2012). Central nervous system control of metabolism. Nature 491, 357–363.

Google Scholar

Nahon, J. L., Presse, F., Bittencourt, J. C., Sawchenko, P. E., and Vale, W. (1989). The rat melanin-concentrating hormone messenger ribonucleic acid encodes multiple putative neuropeptides coexpressed in the dorsolateral hypothalamus. Endocrinology 125, 2056–2065. doi: 10.1210/endo-125-4-2056

PubMed Abstract | CrossRef Full Text | Google Scholar

Nakazato, M., Murakami, N., Date, Y., Kojima, M., Matsuo, H., Kangawa, K., et al. (2001). A role for ghrelin in the central regulation of feeding. Nature 409, 194–198.

Google Scholar

Nambu, T., Sakurai, T., Mizukami, K., Hosoya, Y., Yanagisawa, M., and Goto, K. (1999). Distribution of orexin neurons in the adult rat brain. Brain Res. 827, 243–260.

Google Scholar

Nevsimalova, S., Vankova, J., Stepanova, I., Seemanova, E., Mignot, E., and Nishino, S. (2005). Hypocretin deficiency in Prader-Willi syndrome. Eur. J. Neurol. 12, 70–72.

Google Scholar

Nishino, S., Ripley, B., Mignot, E., Benson, K. L., and Zarcone, V. P. (2002). CSF hypocretin-1 levels in schizophrenics and controls: Relationship to sleep architecture. Psychiatry Res. 110, 1–7. doi: 10.1016/s0165-1781(02)00032-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Nitz, D., and Siegel, J. M. (1996). GABA release in posterior hypothalamus across sleep-wake cycle. Am. J. Physiol. 271, R1707–R1712. doi: 10.1152/ajpregu.1996.271.6.R1707

PubMed Abstract | CrossRef Full Text | Google Scholar

Noble, E. E., Hahn, J. D., Konanur, V. R., Hsu, T. M., Page, S. J., Cortella, A. M., et al. (2018). Control of feeding behavior by cerebral ventricular volume transmission of melanin-concentrating hormone. Cell Metab. 28:e57. doi: 10.1016/j.cmet.2018.05.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Ogilvie, R. P., and Patel, S. R. (2017). The epidemiology of sleep and obesity. Sleep Health 3, 383–388.

Google Scholar

Olszewski, P. K., Li, D., Grace, M. K., Billington, C. J., Kotz, C. M., and Levine, A. S. (2003). Neural basis of orexigenic effects of ghrelin acting within lateral hypothalamus. Peptides 24, 597–602. doi: 10.1016/s0196-9781(03)00105-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Papotti, M., Ghe, C., Cassoni, P., Catapano, F., Deghenghi, R., Ghigo, E., et al. (2000). Growth hormone secretagogue binding sites in peripheral human tissues. J. Clin. Endocrinol. Metab. 85, 3803–3807.

Google Scholar

Pascovich, C., Lagos, P., Urbanavicius, J., Devera, A., Rivas, M., Costa, A., et al. (2020). Melanin-concentrating hormone (MCH) in the median raphe nucleus: Fibers, receptors and cellular effects. Peptides 126:170249.

Google Scholar

Pascovich, C., Nino, S., Mondino, A., Lopez-Hill, X., Urbanavicius, J., Monti, J., et al. (2021). Microinjection of melanin-concentrating hormone (MCH) into the median raphe nucleus promotes REM sleep in rats. Sleep Sci. 14, 229–235.

Google Scholar

Peyron, C., Tighe, D. K., van den Pol, A. N., de Lecea, L., Heller, H. C., Sutcliffe, J. G., et al. (1998). Neurons containing hypocretin (orexin) project to multiple neuronal systems. J. Neurosci. 18, 9996–10015.

Google Scholar

Piper, D. C., Upton, N., Smith, M. I., and Hunter, A. J. (2000). The novel brain neuropeptide, orexin-A, modulates the sleep-wake cycle of rats. Eur. J. Neurosci. 12, 726–730.

Google Scholar

Pissios, P., Bradley, R. L., and Maratos-Flier, E. (2006). Expanding the scales: The multiple roles of MCH in regulating energy balance and other biological functions. Endocr. Rev. 27, 606–620. doi: 10.1210/er.2006-0021

PubMed Abstract | CrossRef Full Text | Google Scholar

Qian, J., Morris, C. J., Caputo, R., Garaulet, M., and Scheer, F. (2019). Ghrelin is impacted by the endogenous circadian system and by circadian misalignment in humans. Int. J. Obes. 43, 1644–1649. doi: 10.1038/s41366-018-0208-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Qu, D., Ludwig, D. S., Gammeltoft, S., Piper, M., Pelleymounter, M. A., Cullen, M. J., et al. (1996). A role for melanin-concentrating hormone in the central regulation of feeding behaviour. Nature 380, 243–247.

Google Scholar

Ramanjaneya, M., Conner, A. C., Chen, J., Kumar, P., Brown, J. E., Johren, O., et al. (2009). Orexin-stimulated MAP kinase cascades are activated through multiple G-protein signalling pathways in human H295R adrenocortical cells: Diverse roles for orexins A and B. J. Endocrinol. 202, 249–261. doi: 10.1677/JOE-08-0536

PubMed Abstract | CrossRef Full Text | Google Scholar

Ravussin, E., Tschop, M., Morales, S., Bouchard, C., and Heiman, M. L. (2001). Plasma ghrelin concentration and energy balance: Overfeeding and negative energy balance studies in twins. J. Clin. Endocrinol. Metab. 86, 4547–4551. doi: 10.1210/jcem.86.9.8003

PubMed Abstract | CrossRef Full Text | Google Scholar

Risold, P. Y., Thompson, R. H., and Swanson, L. W. (1997). The structural organization of connections between hypothalamus and cerebral cortex. Brain Res. Brain Res. Rev. 24, 197–254.

Google Scholar

Rodgers, R. J., Halford, J. C., Nunes de Souza, R. L., Canto, de Souza, A. L., Piper, D. C., et al. (2001). SB-334867, a selective orexin-1 receptor antagonist, enhances behavioural satiety and blocks the hyperphagic effect of orexin-A in rats. Eur. J. Neurosci. 13, 1444–1452. doi: 10.1046/j.0953-816x.2001.01518.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Roh, E., Song, D. K., and Kim, M. S. (2016). Emerging role of the brain in the homeostatic regulation of energy and glucose metabolism. Exp. Mol. Med. 48:e216.

Google Scholar

Rolls, A., Schaich Borg, J., and de Lecea, L. (2010). Sleep and metabolism: Role of hypothalamic neuronal circuitry. Best Pract. Res. Clin. Endocrinol. Metab. 24, 817–828.

Google Scholar

Rosin, D. L., Weston, M. C., Sevigny, C. P., Stornetta, R. L., and Guyenet, P. G. (2003). Hypothalamic orexin (hypocretin) neurons express vesicular glutamate transporters VGLUT1 or VGLUT2. J. Comp. Neurol. 465, 593–603.

Google Scholar

Rossi, M. A., Basiri, M. L., McHenry, J. A., Kosyk, O., Otis, J. M., van den Munkhof, H. E., et al. (2019). Obesity remodels activity and transcriptional state of a lateral hypothalamic brake on feeding. Science 364, 1271–1274. doi: 10.1126/science.aax1184

PubMed Abstract | CrossRef Full Text | Google Scholar

Rotter, A., Asemann, R., Decker, A., Kornhuber, J., and Biermann, T. (2011). Orexin expression and promoter-methylation in peripheral blood of patients suffering from major depressive disorder. J. Affect. Disord. 131, 186–192. doi: 10.1016/j.jad.2010.12.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Saad, M. F., Bernaba, B., Hwu, C. M., Jinagouda, S., Fahmi, S., Kogosov, E., et al. (2002). Insulin regulates plasma ghrelin concentration. J. Clin. Endocrinol. Metab. 87, 3997–4000.

Google Scholar

Sahu, A. (1998). Leptin decreases food intake induced by melanin-concentrating hormone (MCH), galanin (GAL) and neuropeptide Y (NPY) in the rat. Endocrinology 139, 4739–4742. doi: 10.1210/endo.139.11.6432

PubMed Abstract | CrossRef Full Text | Google Scholar

Saito, Y. C., Maejima, T., Nishitani, M., Hasegawa, E., Yanagawa, Y., Mieda, M., et al. (2018a). Monoamines inhibit GABAergic neurons in ventrolateral preoptic area that make direct synaptic connections to hypothalamic arousal neurons. J. Neurosci. 38, 6366–6378. doi: 10.1523/JNEUROSCI.2835-17.2018

PubMed Abstract | CrossRef Full Text | Google Scholar

Saito, Y. C., Tsujino, N., Abe, M., Yamazaki, M., Sakimura, K., and Sakurai, T. (2018b). Serotonergic input to orexin neurons plays a role in maintaining wakefulness and REM sleep architecture. Front. Neurosci. 12:892. doi: 10.3389/fnins.2018.00892

PubMed Abstract | CrossRef Full Text | Google Scholar

Saito, Y. C., Tsujino, N., Hasegawa, E., Akashi, K., Abe, M., Mieda, M., et al. (2013). GABAergic neurons in the preoptic area send direct inhibitory projections to orexin neurons. Front. Neural Circuits 7:192.

Google Scholar

Sakurai, T. (2007). The neural circuit of orexin (hypocretin): Maintaining sleep and wakefulness. Nat. Rev. Neurosci. 8, 171–181.

Google Scholar

Sakurai, T., Amemiya, A., Ishii, M., Matsuzaki, I., Chemelli, R. M., Tanaka, H., et al. (1998). Orexins and orexin receptors: A family of hypothalamic neuropeptides and G protein-coupled receptors that regulate feeding behavior. Cell 92:696.

Google Scholar

Santos, V. V., Stark, R., Rial, D., Silva, H. B., Bayliss, J. A., Lemus, M. B., et al. (2017). Acyl ghrelin improves cognition, synaptic plasticity deficits and neuroinflammation following amyloid beta (Abeta1-40) administration in mice. J. Neuroendocrinol. 29, 1–11. doi: 10.1111/jne.12476

PubMed Abstract | CrossRef Full Text | Google Scholar

Saper, C. B. (2006). Staying awake for dinner: Hypothalamic integration of sleep, feeding, and circadian rhythms. Prog. Brain Res. 153, 243–252. doi: 10.1016/S0079-6123(06)53014-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Saper, C. B., and Fuller, P. M. (2017). Wake-sleep circuitry: An overview. Curr. Opin. Neurobiol. 44, 186–192. doi: 10.1016/j.conb.2017.03.021

PubMed Abstract | CrossRef Full Text | Google Scholar

Sasaki, K., Suzuki, M., Mieda, M., Tsujino, N., Roth, B., and Sakurai, T. (2011). Pharmacogenetic modulation of orexin neurons alters sleep/wakefulness states in mice. PLoS One 6:e20360. doi: 10.1371/journal.pone.0020360

PubMed Abstract | CrossRef Full Text | Google Scholar

Scammell, T. E., and Winrow, C. J. (2011). Orexin receptors: Pharmacology and therapeutic opportunities. Annu. Rev. Pharmacol. Toxicol. 51, 243–266.

Google Scholar

Schmid, S. M., Hallschmid, M., Jauch-Chara, K., Born, J., and Schultes, B. (2008). A single night of sleep deprivation increases ghrelin levels and feelings of hunger in normal-weight healthy men. J. Sleep Res. 17, 331–334. doi: 10.1111/j.1365-2869.2008.00662.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Schneeberger, M., Tan, K., Nectow, A. R., Parolari, L., Caglar, C., Azevedo, E., et al. (2018). Functional analysis reveals differential effects of glutamate and MCH neuropeptide in MCH neurons. Mol. Metab. 13, 83–89. doi: 10.1016/j.molmet.2018.05.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Schwartz, M. W., and Porte, D. Jr. (2005). Diabetes, obesity, and the brain. Science 307, 375–379.

Google Scholar

Schwartz, M. W., Baskin, D. G., Bukowski, T. R., Kuijper, J. L., Foster, D., Lasser, G., et al. (1996). Specificity of leptin action on elevated blood glucose levels and hypothalamic neuropeptide Y gene expression in ob/ob mice. Diabetes 45, 531–535. doi: 10.2337/diab.45.4.531

PubMed Abstract | CrossRef Full Text | Google Scholar

Schwartz, M. W., Woods, S. C., Porte, D. Jr., Seeley, R. J., and Baskin, D. G. (2000). Central nervous system control of food intake. Nature 404, 661–671.

Google Scholar

Segal-Lieberman, G., Bradley, R. L., Kokkotou, E., Carlson, M., Trombly, D. J., Wang, X., et al. (2003). Melanin-concentrating hormone is a critical mediator of the leptin-deficient phenotype. Proc. Natl. Acad. Sci. U. S. A. 100, 10085–10090. doi: 10.1073/pnas.1633636100

PubMed Abstract | CrossRef Full Text | Google Scholar

Sheng, Z., Santiago, A. M., Thomas, M. P., and Routh, V. H. (2014). Metabolic regulation of lateral hypothalamic glucose-inhibited orexin neurons may influence midbrain reward neurocircuitry. Mol. Cell Neurosci. 62, 30–41. doi: 10.1016/j.mcn.2014.08.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Shimada, M., Tritos, N. A., Lowell, B. B., Flier, J. S., and Maratos-Flier, E. (1998). Mice lacking melanin-concentrating hormone are hypophagic and lean. Nature 396, 670–674.

Google Scholar

Shin, S. Y., Yang, J. H., Lee, H., Erdelyi, F., Szabo, G., Lee, S. Y., et al. (2007). Identification of the adrenoceptor subtypes expressed on GABAergic neurons in the anterior hypothalamic area and rostral zona incerta of GAD65-eGFP transgenic mice. Neurosci. Lett. 422, 153–157. doi: 10.1016/j.neulet.2007.05.060

PubMed Abstract | CrossRef Full Text | Google Scholar

Skofitsch, G., Jacobowitz, D. M., and Zamir, N. (1985). Immunohistochemical localization of a melanin concentrating hormone-like peptide in the rat brain. Brain Res. Bull. 15, 635–649. doi: 10.1016/0361-9230(85)90213-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Strawn, J. R., Pyne-Geithman, G. J., Ekhator, N. N., Horn, P. S., Uhde, T. W., Shutter, L. A., et al. (2010). Low cerebrospinal fluid and plasma orexin-A (hypocretin-1) concentrations in combat-related posttraumatic stress disorder. Psychoneuroendocrinology 35, 1001–1007. doi: 10.1016/j.psyneuen.2010.01.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Stuber, G. D., and Wise, R. A. (2016). Lateral hypothalamic circuits for feeding and reward. Nat. Neurosci. 19, 198–205.

Google Scholar

Sun, Y., Wang, P., Zheng, H., and Smith, R. G. (2004). Ghrelin stimulation of growth hormone release and appetite is mediated through the growth hormone secretagogue receptor. Proc. Natl. Acad. Sci. U. S. A. 101, 4679–4684.

Google Scholar

Suzuki, M., Beuckmann, C. T., Shikata, K., Ogura, H., and Sawai, T. (2005). Orexin-A (hypocretin-1) is possibly involved in generation of anxiety-like behavior. Brain Res. 1044, 116–121. doi: 10.1016/j.brainres.2005.03.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Sweet, D. C., Levine, A. S., Billington, C. J., and Kotz, C. M. (1999). Feeding response to central orexins. Brain Res. 821, 535–538.

Google Scholar

Szentirmai, E., Kapas, L., and Krueger, J. M. (2007a). Ghrelin microinjection into forebrain sites induces wakefulness and feeding in rats. Am. J. Physiol. Regul. Integr. Comp. Physiol. 292, R575–R585. doi: 10.1152/ajpregu.00448.2006

PubMed Abstract | CrossRef Full Text | Google Scholar

Szentirmai, E., Kapas, L., Sun, Y., Smith, R. G., and Krueger, J. M. (2007b). Spontaneous sleep and homeostatic sleep regulation in ghrelin knockout mice. Am. J. Physiol. Regul. Integr. Comp. Physiol. 293, R510–R517. doi: 10.1152/ajpregu.00155.2007

PubMed Abstract | CrossRef Full Text | Google Scholar

Taheri, S., Gardiner, J., Hafizi, S., Murphy, K., Dakin, C., Seal, L., et al. (2001). Orexin A immunoreactivity and preproorexin mRNA in the brain of Zucker and WKY rats. Neuroreport 12, 459–464. doi: 10.1097/00001756-200103050-00008

PubMed Abstract | CrossRef Full Text | Google Scholar

Tanaka, M., and Itoh, H. (2019). Hypertension as a metabolic disorder and the novel role of the gut. Curr. Hypertens. Rep. 21:63.

Google Scholar

Teitelbaum, P., and Stellar, E. (1954). Recovery from the failure to eat produced by hypothalamic lesions. Science 120, 894–895. doi: 10.1126/science.120.3126.894

PubMed Abstract | CrossRef Full Text | Google Scholar

Thannickal, T. C., Moore, R. Y., Nienhuis, R., Ramanathan, L., Gulyani, S., Aldrich, M., et al. (2000). Reduced number of hypocretin neurons in human narcolepsy. Neuron 27, 469–474.

Google Scholar

Thorpe, A. J., Mullett, M. A., Wang, C., and Kotz, C. M. (2003). Peptides that regulate food intake: Regional, metabolic, and circadian specificity of lateral hypothalamic orexin A feeding stimulation. Am. J. Physiol. Regul. Integr. Comp. Physiol. 284, R1409–R1417. doi: 10.1152/ajpregu.00344.2002

PubMed Abstract | CrossRef Full Text | Google Scholar

Timper, K., and Bruning, J. C. (2017). Hypothalamic circuits regulating appetite and energy homeostasis: Pathways to obesity. Dis. Model Mech. 10, 679–689.

Google Scholar

Tolle, V., Bassant, M. H., Zizzari, P., Poindessous-Jazat, F., Tomasetto, C., Epelbaum, J., et al. (2002). Ultradian rhythmicity of ghrelin secretion in relation with GH, feeding behavior, and sleep-wake patterns in rats. Endocrinology 143, 1353–1361. doi: 10.1210/endo.143.4.8712

PubMed Abstract | CrossRef Full Text | Google Scholar

Torrealba, F., Yanagisawa, M., and Saper, C. B. (2003). Colocalization of orexin a and glutamate immunoreactivity in axon terminals in the tuberomammillary nucleus in rats. Neuroscience 119, 1033–1044. doi: 10.1016/s0306-4522(03)00238-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Torterolo, P., Lagos, P., and Monti, J. M. (2011). Melanin-concentrating hormone: A new sleep factor? Front. Neurol. 2:14. doi: 10.3389/fneur.2011.00014

PubMed Abstract | CrossRef Full Text | Google Scholar

Torterolo, P., Sampogna, S., and Chase, M. H. (2009). MCHergic projections to the nucleus pontis oralis participate in the control of active (REM) sleep. Brain Res. 1268, 76–87. doi: 10.1016/j.brainres.2009.02.055

PubMed Abstract | CrossRef Full Text | Google Scholar

Torterolo, P., Sampogna, S., and Chase, M. H. (2013). Hypocretinergic and non-hypocretinergic projections from the hypothalamus to the REM sleep executive area of the pons. Brain Res. 1491, 68–77. doi: 10.1016/j.brainres.2012.10.050

PubMed Abstract | CrossRef Full Text | Google Scholar

Toshinai, K., Date, Y., Murakami, N., Shimada, M., Mondal, M. S., Shimbara, T., et al. (2003). Ghrelin-induced food intake is mediated via the orexin pathway. Endocrinology 144, 1506–1512. doi: 10.1210/en.2002-220788

PubMed Abstract | CrossRef Full Text | Google Scholar

Toshinai, K., Yamaguchi, H., Sun, Y., Smith, R. G., Yamanaka, A., Sakurai, T., et al. (2006). Des-acyl ghrelin induces food intake by a mechanism independent of the growth hormone secretagogue receptor. Endocrinology 147, 2306–2314.

Google Scholar

Tsunematsu, T., Ueno, T., Tabuchi, S., Inutsuka, A., Tanaka, K. F., Hasuwa, H., et al. (2014). Optogenetic manipulation of activity and temporally controlled cell-specific ablation reveal a role for MCH neurons in sleep/wake regulation. J. Neurosci. 34, 6896–6909. doi: 10.1523/JNEUROSCI.5344-13.2014

PubMed Abstract | CrossRef Full Text | Google Scholar

Turunen, P. M., Jantti, M. H., and Kukkonen, J. P. (2012). OX1 orexin/hypocretin receptor signaling through arachidonic acid and endocannabinoid release. Mol. Pharmacol. 82, 156–167. doi: 10.1124/mol.112.078063

PubMed Abstract | CrossRef Full Text | Google Scholar

van den Pol, A. N., Gao, X. B., Obrietan, K., Kilduff, T. S., and Belousov, A. B. (1998). Presynaptic and postsynaptic actions and modulation of neuroendocrine neurons by a new hypothalamic peptide, hypocretin/orexin. J. Neurosci. 18, 7962–7971.

Google Scholar

Varin, C., Arthaud, S., Salvert, D., Gay, N., Libourel, P. A., Luppi, P. H., et al. (2016). Sleep architecture and homeostasis in mice with partial ablation of melanin-concentrating hormone neurons. Behav. Brain Res. 298, 100–110. doi: 10.1016/j.bbr.2015.10.051

PubMed Abstract | CrossRef Full Text | Google Scholar

Varin, C., Luppi, P. H., and Fort, P. (2018). Melanin-concentrating hormone-expressing neurons adjust slow-wave sleep dynamics to catalyze paradoxical (REM) sleep. Sleep 41:zsy068. doi: 10.1093/sleep/zsy068

PubMed Abstract | CrossRef Full Text | Google Scholar

Venner, A., Anaclet, C., Broadhurst, R. Y., Saper, C. B., and Fuller, P. M. (2016). A novel population of wake-promoting GABAergic neurons in the ventral lateral hypothalamus. Curr. Biol. 26, 2137–2143. doi: 10.1016/j.cub.2016.05.078

PubMed Abstract | CrossRef Full Text | Google Scholar

Venner, A., Karnani, M. M., Gonzalez, J. A., Jensen, L. T., Fugger, L., and Burdakov, D. (2011). Orexin neurons as conditional glucosensors: Paradoxical regulation of sugar sensing by intracellular fuels. J. Physiol. 589, 5701–5708. doi: 10.1113/jphysiol.2011.217000

PubMed Abstract | CrossRef Full Text | Google Scholar

Verhulst, P. J., and Depoortere, I. (2012). Ghrelin’s second life: From appetite stimulator to glucose regulator. World J. Gastroenterol. 18, 3183–3195.

Google Scholar

Verret, L., Goutagny, R., Fort, P., Cagnon, L., Salvert, D., Leger, L., et al. (2003). A role of melanin-concentrating hormone producing neurons in the central regulation of paradoxical sleep. BMC Neurosci. 4:19. doi: 10.1186/1471-2202-4-19

PubMed Abstract | CrossRef Full Text | Google Scholar

Vetrivelan, R., Kong, D., Ferrari, L. L., Arrigoni, E., Madara, J. C., Bandaru, S. S., et al. (2016). Melanin-concentrating hormone neurons specifically promote rapid eye movement sleep in mice. Neuroscience 336, 102–113.

Google Scholar

Viskaitis, P., Tesmer, A. L., Karnani, M. M., Arnold, M., Donegan, D., Bracey, E. F., et al. (2022). Orexin cells efficiently decode blood glucose dynamics to drive adaptive behavior. bioRxiv [Preprint]. doi: 10.1101/2022.04.14.488310

CrossRef Full Text | Google Scholar

Wang, Q., Yin, Y., and Zhang, W. (2018). Ghrelin restores the disruption of the circadian clock in steatotic liver. Int. J. Mol. Sci. 19:3134. doi: 10.3390/ijms19103134

PubMed Abstract | CrossRef Full Text | Google Scholar

Wei, H., Cao, X., Zeng, Q., Zhang, F., Xue, Q., Luo, Y., et al. (2015). Ghrelin inhibits proinflammatory responses and prevents cognitive impairment in septic rats. Crit. Care Med. 43, e143–e150. doi: 10.1097/CCM.0000000000000930

PubMed Abstract | CrossRef Full Text | Google Scholar

Weikel, J. C., Wichniak, A., Ising, M., Brunner, H., Friess, E., Held, K., et al. (2003). Ghrelin promotes slow-wave sleep in humans. Am. J. Physiol. Endocrinol. Metab. 284, E407–E415.

Google Scholar

Wen, J., Zhao, Y., Shen, Y., and Guo, L. (2015). Effect of orexin A on apoptosis in BGC-823 gastric cancer cells via OX1R through the AKT signaling pathway. Mol. Med. Rep. 11, 3439–3444. doi: 10.3892/mmr.2015.3190

PubMed Abstract | CrossRef Full Text | Google Scholar

White, C. L., Ishii, Y., Mendoza, T., Upton, N., Stasi, L. P., Bray, G. A., et al. (2005). Effect of a selective OX1R antagonist on food intake and body weight in two strains of rats that differ in susceptibility to dietary-induced obesity. Peptides 26, 2331–2338. doi: 10.1016/j.peptides.2005.03.042

PubMed Abstract | CrossRef Full Text | Google Scholar

Williams, R. H., Alexopoulos, H., Jensen, L. T., Fugger, L., and Burdakov, D. (2008). Adaptive sugar sensors in hypothalamic feeding circuits. Proc. Natl. Acad. Sci. U. S. A. 105, 11975–11980.

Google Scholar

Willie, J. T., Chemelli, R. M., Sinton, C. M., and Yanagisawa, M. (2001). To eat or to sleep? Orexin in the regulation of feeding and wakefulness. Annu. Rev. Neurosci. 24, 429–458.

Google Scholar

Willie, J. T., Chemelli, R. M., Sinton, C. M., Tokita, S., Williams, S. C., Kisanuki, Y. Y., et al. (2003). Distinct narcolepsy syndromes in Orexin receptor-2 and Orexin null mice: Molecular genetic dissection of Non-REM and REM sleep regulatory processes. Neuron 38, 715–730. doi: 10.1016/s0896-6273(03)00330-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Willie, J. T., Sinton, C. M., Maratos-Flier, E., and Yanagisawa, M. (2008). Abnormal response of melanin-concentrating hormone deficient mice to fasting: Hyperactivity and rapid eye movement sleep suppression. Neuroscience 156, 819–829. doi: 10.1016/j.neuroscience.2008.08.048

PubMed Abstract | CrossRef Full Text | Google Scholar

Yamamoto, Y., Ueta, Y., Serino, R., Nomura, M., Shibuya, I., and Yamashita, H. (2000). Effects of food restriction on the hypothalamic prepro-orexin gene expression in genetically obese mice. Brain Res. Bull. 51, 515–521. doi: 10.1016/s0361-9230(99)00271-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Yamanaka, A., Beuckmann, C. T., Willie, J. T., Hara, J., Tsujino, N., Mieda, M., et al. (2003). Hypothalamic orexin neurons regulate arousal according to energy balance in mice. Neuron 38, 701–713. doi: 10.1016/s0896-6273(03)00331-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Yamanaka, A., Kunii, K., Nambu, T., Tsujino, N., Sakai, A., Matsuzaki, I., et al. (2000). Orexin-induced food intake involves neuropeptide Y pathway. Brain Res. 859, 404–409.

Google Scholar

Yamanaka, A., Tsujino, N., Funahashi, H., Honda, K., Guan, J. L., Wang, Q. P., et al. (2002). Orexins activate histaminergic neurons via the orexin 2 receptor. Biochem. Biophys. Res. Commun. 290, 1237–1245.

Google Scholar

Yang, B., and Ferguson, A. V. (2002). Orexin-A depolarizes dissociated rat area postrema neurons through activation of a nonselective cationic conductance. J. Neurosci. 22, 6303–6308. doi: 10.1523/JNEUROSCI.22-15-06303.2002

PubMed Abstract | CrossRef Full Text | Google Scholar

Yannielli, P. C., Molyneux, P. C., Harrington, M. E., and Golombek, D. A. (2007). Ghrelin effects on the circadian system of mice. J. Neurosci. 27, 2890–2895.

Google Scholar

Yin, Y., Li, Y., and Zhang, W. (2014). The growth hormone secretagogue receptor: Its intracellular signaling and regulation. Int. J. Mol. Sci. 15, 4837–4855.

Google Scholar

Yoon, Y. S., and Lee, H. S. (2013). Projections from melanin-concentrating hormone (MCH) neurons to the dorsal raphe or the nuclear core of the locus coeruleus in the rat. Brain Res. 1490, 72–82.

Google Scholar

Yoshida, K., McCormack, S., Espana, R. A., Crocker, A., and Scammell, T. E. (2006). Afferents to the orexin neurons of the rat brain. J. Comp. Neurol. 494, 845–861. doi: 10.1002/cne.20859

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, S., Zeitzer, J. M., Sakurai, T., Nishino, S., and Mignot, E. (2007). Sleep/wake fragmentation disrupts metabolism in a mouse model of narcolepsy. J. Physiol. 581, 649–663. doi: 10.1113/jphysiol.2007.129510

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, Y., Proenca, R., Maffei, M., Barone, M., Leopold, L., and Friedman, J. M. (1994). Positional cloning of the mouse obese gene and its human homologue. Nature 372, 425–432.

Google Scholar

Zhu, Y., Yamanaka, A., Kunii, K., Tsujino, N., Goto, K., and Sakurai, T. (2002). Orexin-mediated feeding behavior involves both leptin-sensitive and -insensitive pathways. Physiol. Behav. 77, 251–257. doi: 10.1016/s0031-9384(02)00843-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Zink, A. N., Bunney, P. E., Holm, A. A., Billington, C. J., and Kotz, C. M. (2018). Neuromodulation of orexin neurons reduces diet-induced adiposity. Int. J. Obes. 42, 737–745. doi: 10.1038/ijo.2017.276

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: sleep, metabolism, orexin, MCH, feeding, wake, lateral hypothalamus

Citation: Bouâouda H and Jha PK (2023) Orexin and MCH neurons: regulators of sleep and metabolism. Front. Neurosci. 17:1230428. doi: 10.3389/fnins.2023.1230428

Received: 28 May 2023; Accepted: 07 August 2023;
Published: 22 August 2023.

Edited by:

Henrik Oster, University of Lübeck, Germany

Reviewed by:

Jin Bao, Chinese Academy of Sciences (CAS), China
Chak Foon Tso, Independent Researcher, Sunnyvale, CA, United States

Copyright © 2023 Bouâouda and Jha. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Hanan Bouâouda, hanan.bouaouda@gmail.com; Pawan Kumar Jha, kj.pawan@gmail.com

These authors have contributed equally to this work

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.