Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2016 May 10.
Published in final edited form as: Cell Metab. 2012 Aug 16;16(3):296–309. doi: 10.1016/j.cmet.2012.06.015

Hindbrain Neurons as an Essential Hub in the Neuroanatomically Distributed Control of Energy Balance

Harvey J Grill 1,*, Matthew R Hayes 1,+
PMCID: PMC4862653  NIHMSID: NIHMS398948  PMID: 22902836

Summary

This review highlights the processing and integration performed by hindbrain nuclei, focusing on the inputs received by nucleus tractus solitarius (NTS) neurons. These inputs include vagally-mediated gastrointestinal satiation signals, blood-borne energy-related hormonal and nutrient signals, and descending neural signals from the forebrain. We propose that NTS (and hindbrain neurons more broadly) integrate these multiple energy status signals and issue output commands controlling the behavioral, autonomic, and endocrine responses that collectively govern energy balance. These hindbrain-mediated controls are neuroanatomically distributed; they involve endemic hindbrain neurons and circuits, hindbrain projections to peripheral circuits, and projections to and from midbrain and forebrain nuclei.

Introduction

The rapid and dramatic increase in obesity prevalence over the last three decades cannot be attributed to “metabolic need” alone. As a result, so-called “need-independent” controls of human eating behavior, such as the feeding triggered by the attraction of food (a.k.a. food reward) or by conditioned, social, and temporal cues to eat are generating renewed attention as an explanation of the chronic hyperphagia that is the hallmark of the obesity “epidemic.” At the same time, recent experiments show that the central nervous system (CNS) action of energy status signals such as leptin, ghrelin, glucagon-like peptide-1 (GLP-1) and peptide YY (PYY) affect food preference, the incentive value of food, and other cognitive controls of food intake (Batterham et al., 2007; Kanoski et al., 2011b; Kenny, 2011; Malik et al., 2008; Rosenbaum et al., 2008). These and other data support the view that the brain processes environmental and hedonic information that interacts with the neural processing of energy status signals to determine when we initiate feeding and how much we eat.

This review addresses the location of the neurons and the neural circuits that comprise the energy balance control system. The review also considers the anatomical path from input (e.g., energy status signal sensing) to output (response production) and discusses neuroanatomical models of energy balance control. A central theme is that hindbrain neurons make several key contributions to a neuroanatomically distributed energy balance control system – they: [1] detect and integrate energy status signals, [2] receive input from neurons in the periphery and from other brain regions that also detect and integrate energy information, [3] project to other brain regions providing information that is integrated by those neurons to control energy balance, and [4] control responses through their local hindbrain, their rostral, and their peripheral projections.

The sections that follow discuss contributions of hindbrain neurons to: [1] input and integration of energy status signals, [2] output and response control, [3] the integration of descending information arising from forebrain neurons, and [4] the transmission of hindbrain-processed signals to midbrain and forebrain neurons.

I. Input: Hindbrain Neurons Receive and Integrate Energy Status Signals

Among the array of CNS nuclei that contribute to energy balance control, it is arguable that the nucleus of the solitary tract (NTS) receives and processes the greatest amount of neuronally mediated and circulating energy status signals (Figure 1a). This section discusses the NTS processing of energy status signals conveyed via: [1] vagal afferent transmissions from the gastrointestinal (GI) tract; [2] blood-borne endocrine signals secreted from peripheral organs (i.e. adipose tissue, pancreas, GI tract); and [3] circulating nutrients (e.g. glucose). We take the perspective that the CNS control of energy balance involves coordinated, multi-synaptic processing by anatomically distributed neurons. In this section we highlight accumulating evidence showing that integrated processing of this array of complimentary and antagonizing signals begins within the NTS.

Figure 1.

Figure 1

a: Afferent input to NTS neurons.

The figure highlights input to medial NTS neurons from blood borne energy status signals such as leptin, ghrelin, and glucose, from the gastrointestinal tract (via the vagus nerve) that processing of sensory signals of gastrointestinal origin as well as responding to leptin and other energy status signals, and from hypothalamic neurons (PVH, LH and ARH) that themselves are responsive to energy status signals. For clarity inputs to medial NTS neurons from other parts of the brain and inputs to neurons in other regions of the NTS that receive cranial nerve input of oral (gustatory) and thoracic origin are not included.

Figure 1b: Output from NTS neurons.

NTS neurons from caudal regions (cNTS) project: to vagal efferent neurons in DMV to control parasympathetic gastrointestinal responses including insulin secretion and gastric emptying; to the intermediolateral cell column of the spinal cord along with projections from neurons in other regions of the hindbrain (RPa, VLM, PBN) and hypothalamus (POA, DMH, ARH) to control sympathetic efferent responses of relevance to energy expenditure and gastrointestinal responses; and to PVH neurons to influence neuroendocrine responses. Neurons from rostral regions of NTS (rNTS) and PBN neurons project to the parvocellular reticular formation (PCRt) to control ingestive consummatory (licking, chewing, swallowing, rejection) response. For clarity input to these response pathways from other brain regions, not highlighted in text, are not included

Figure 1c: Rostral projections of NTS neurons.

The labeled structures receive monosynaptic projections from neurons located in different regions of the NTS. For clarity other projections of NTS neurons to sites within the hindbrain and to the periphery are not shown.

Vagal mediation of gastrointestinal satiation signals

During meal ingestion, the GI tract senses the chemical, nutritive, osmotic, and volumetric properties of the ingested food via a complex communication system that involves interaction between enteroendocrine cells and vagal and spinal afferents. The proximal location of the stomach within the GI tract and its rich vagal afferent innervatation (Wang and Powley, 2000) provide an early monitoring system for the status of meal ingestion. These gastric intake inhibitory signals involve volumetric-mechanical distension of the organ and not the chemical / nutritive properties of the ingestate (Mathis et al., 1998; Phillips and Powley, 1996; Powley and Phillips, 2004; Seeley et al., 1995). The vagal sensory endings responsive to stretch and/or tension appear to include both intraganglionic laminar endings (IGLEs) and intramuscular arrays (IMAs) [see (Ritter, 2004) for review]. Tension and stretch detection of the gastric wall by IGLEs and IMAs results in glutamatergic neuronal transmission from vagal axon projections to NTS neurons. However, serotonin (5-HT) secreted from gastric enterochromaffin (EC) cells provides the principal intake inhibitory secretion signal following volumetric distension. This satiating response following gastric distension-derived serotonergic excitation of the vagus is principally mediated by 5-HT type-3 receptors (5-HT3R) expressed on peripheral dendritic terminals of vagal afferents innervating the stomach (Glatzle et al., 2002; Hayes et al., 2006; Hayes and Covasa, 2006; Hayes et al., 2004; Mazda et al., 2004). Thus, in addition to vagal-glutamatergic transmission to the NTS, gastric distension results in vagal axon serotonergic signaling to caudal NTS neurons (Berthoud et al., 2004; Ritter, 2004).

The chemical and nutritive properties of the ingested food stimulate the release of a number of gut peptides and neurotransmitters from the small intestine. A non-comprehensive list of the prominent intestinally-derived satiation signals (i.e. intake inhibitory signal arising during the meal and influencing its course) includes: cholecystokinin (CCK), 5-HT, PYY, glutamate, enterostatin, and GLP-1 (Chaudhri et al., 2006). While receptors for many of these satiation signals are expressed in the brain, under normal physiological conditions, the available circulating levels of these chemical signals appear to not be elevated in sufficient quantity to have direct action within the brain. Instead, the majority of GI-derived satiation signaling is communicated to the brain via a paracrine-like activation of specific receptors expressed on the dendritic terminals of vagal afferent neurons whose cell bodies are found in the nodose ganglion (Hayes et al., 2010a; Moran, 2006; Ritter, 2004; Smith, 1996). Support for vagal afferent mediation of many GI-derived satiation signals comes from studies that chemically or surgically lesion the vagus, revealing that in the absence of vagal afferent transmission the intake inhibitory response to peripherally administered satiation signals is blocked or reduced (e.g. CCK, GLP-1, gastric distension, PYY) (Kanoski et al., 2011a; Mazda et al., 2004; Neary et al., 2005; Smith et al., 1985).

Vagal Processing and Integration

Given that many gut peptides signal for a reduction in feeding, a fundamentally simple question with a complex answer is: How does the brain distinguish its response to one vagally-mediated gut peptide (e.g. GLP-1) from another (e.g. CCK)? While a multitude of possible explanations exist by which the vagus and brain partition and simultaneously integrate an array of anorectic signals, the solution likely involves a combination of the following hypotheses: 1) selective branches of vagal afferents are more or less responsive to different neuropeptides as a result of varying receptor expression patterns along the GI tract and supporting organs; 2) the type of vagal afferents activated [i.e myelinated (type-A) and/or unmyelinated (type-C)] produces a unique transmission pattern to the NTS (Ritter, 2004); 3) the specific vagal neurotransmitter(s) (i.e. 5-HT and/or glutamate) engaged by specific gut peptides may differentially encode signals to the NTS; and 4) the intensity, frequency, and duration of neuronal excitation by each gut peptide on the vagus is also likely different for each hormone.

An important example of vagal integration of satiation signals is that of CCK and gastric distension. Schwartz and Moran (Schwartz et al., 1993; Schwartz and Moran, 1996) show that single vagal afferent fibers are responsive to both CCK and gastric distension, and when these two GI satiation signals are applied in combination the firing rate and total spike number increase in a dose- and volume-dependent fashion. Indeed, this vagal integration is postulated to contribute to the enhanced behavioral suppression of food intake when CCK and gastric distension are combined (Moran et al., 2001; Ritter, 2004; Schwartz and Moran, 1996). Under normal physiological conditions, activation of CCK-1 receptors suppresses gastric emptying, thereby temporally enhancing gastric distension (Bozkurt et al., 1999; Moran and McHugh, 1982; Schwartz et al., 1991). Importantly, however, CCK-1 receptors do not mediate gastric-distension-induced neuronal signaling (van de Wall et al., 2005; Yoshida-Yoneda et al., 1996). Therefore, the vagal interactions between CCK and gastric distension likely involve participation of other GI-derived satiating signals, such as 5-HT, which is both released in response to gastric distension (Mazda et al., 2004) and interacts with CCK to reduce food intake (Hayes and Covasa, 2005). Indeed, blockade of 5-HT3R attenuates suppression of food intake by CCK (Daughters et al., 2001; Hayes et al., 2004), intraduodenal nutrients (Burton-Freeman et al., 1999; Savastano et al., 2005), and gastric distension (Hayes et al., 2006; Hayes and Covasa, 2006). The interaction and processing of the intake inhibitory signaling of CCK and gastric distension by the vagus nerve is but one example of many interactions that occur between various GI-derived satiation signals.

Integrations performed by vagal afferent neurons are not limited to those involving GI-derived satiation signals but also include the processing of circulating signals of energy availability (i.e. leptin). Given that gastric chief cells and some mucosal cells secrete leptin in response to nutrient ingestion (Bado et al., 1998), and that leptin receptors (LepRb) are expressed on cell bodies of vagal afferent neurons in the nodose ganglion (Burdyga et al., 2002; Buyse et al., 2001), experiments by Peters and colleagues examined the hypotheses that gastric leptin acts as a satiation signal by activating LepRb-expressing terminals of vagal afferent neurons in a paracrine-like fashion (Peters et al., 2005a) and that the vagus integrates leptin's anorectic signal with other vagally mediated satiation signals (e.g. CCK) (Peters et al., 2005b). When combined, IP CCK and intra-celiac infusion of low doses of leptin (resulting in GI specific elevations in leptin) interact to suppress food intake and augment neural activity in nodose neurons. Therefore, it is worth examining whether this type of enhanced nodose electrophysiological response can result in enhanced NTS signaling and subsequently augmented food intake suppression.

Blood-borne Energy Status Signals

Leptin receptor

LepRb are expressed in three caudal brainstem nuclei, the parabrachial nucleus, the area postrema (AP) and the NTS (Patterson et al., 2011; Scott et al., 2009). The functional effects of leptin signaling in the parabrachial nucleus are not well investigated. Direct injection of leptin to the AP parenchyma and targeted AP LepRb knockdown (LepRb KD) via adeno-associated virus short hairpin RNA-interference in rats failed to reveal functional effects of AP leptin signaling (Kanoski et al., 2012) see also (Patterson et al., 2011). By contrast, a variety of evidence shows that NTS leptin signaling is an important contributor to energy balance control. Parenchymal NTS leptin injection reduces food intake and body weight (Grill et al., 2002). NTS LepRb positive neurons are among a variety of neurons labeled by the transynaptic effects of pseudorabies viral injection of BAT indicating their potential role in thermogenic control (Zhang et al., 2011). In fact, low doses of leptin injected into the caudal 4th ventricle, whose effect is confined to the dorsal hindbrain (Hayes et al., 2009b), elevates core temperature and heart rate (Skibicka and Grill, 2009a). When medial NTS LepRb expression is reduced by 40% through virally mediated RNAi LepRb knockdown (KD), rats are hyperphagic (on chow, as well as palatable diets), obese, and do not respond to the intake inhibitory effect of CCK (Hayes et al., 2010b); see also complementary data in (Huo et al., 2007). Collectively, these findings support the hypothesis that endogenous NTS leptin signaling is necessary for body weight control through its effects on food intake.

Hindbrain glucose sensors

Neurophysiologic, behavioral, and neuroendocrine data support the idea first proposed by Mayer (Mayer, 1953) that variations in plasma glucose (or correlates of CNS energy availability) alter the neural excitability of specialized neurons that participate in the neural control of energy balance. Glucose-sensitive neurons are found in hindbrain and hypothalamus and, as discussed below, these neurons, through connections to effector circuits, play a critical role in regulating glucose and energy homeostasis by controlling endocrine pancreas secretions, hepatic glucose production, feeding behavior, and energy expenditure. The mechanisms mediating the detection of glycemic changes (and/or the consequences of cytoglucopenia) that are coupled to alterations in the excitability of these glucose-sensitive neurons are still under investigation (Thorens, 2010). It is already clear however, that this glucose sensing process in CNS neurons avails of some of the same proteins that contribute to glucose sensing in pancreatic β-cells.

A variety of studies have evaluated the role of the low-affinity glucose transporter type 2 (GLUT2), glucokinase, and the KATP channel subunit SUR1, SUR2, and Kir6.2 in central glucose sensing [see reviews (Jordan et al., 2010; Levin, 2006; Marty et al., 2007)]. GLUT2 KO mice are hyperphagic and fail to increase or decrease feeding in response to intracerebroventricular (icv) 2-deoxyglucose or glucose, respectively (Bady et al., 2006). Examination of fluorescent reporter genes that label GLUT2 expressing cells reveals a high concentration of hindbrain neurons with this genotype in the NTS, DMV, and VLM (Arluison et al., 2004; Mounien et al., 2010). Astrocytes in the DVC also express GLUT2 (Marty et al., 2005) and glial activation of neurons may play a role in this system as glial-to-neuron communication/activation has been documented for other functions involving NTS neurons (Hermann et al., 2009; McDougal et al., 2011). These GLUT2-expressing hindbrain neurons are coextensive with the location of hindbrain neurons that are: [1] neurophysiologically responsive to changes in plasma glucose or to local glucose levels (Dallaporta et al., 1999; Mizuno and Oomura, 1984; Yettefti et al., 1997), and [2] part of a circuit that drives counterregulatory responses to control glucose homeostasis (Ritter et al., 2000; Ritter et al., 2011). Hindbrain neuron involvement in counterregulation response production is discussed below.

Ghrelin

Ghrelin, a gastric secreted peptide whose endogenous levels are inhibited by ingested food (Overduin et al., 2005), stimulates feeding by its action on the growth hormone secretogue receptor (GHSR). GHSR is broadly distributed in the CNS including expression in the NTS and other hindbrain neurons (Zigman et al., 2006). Direct NTS injection of low doses of ghrelin increases food intake (Faulconbridge et al., 2003). One mechanism that may contribute to NTS ghrelin-stimulated feeding involves attenuating the neural activity in NTS neurons driven by vagal afferent GI satiation signals. Indeed, the tyrosine hydroxylase (TH) expressing A2/C2 neurons in the mNTS that are activated by electrical stimulation of the solitary tract (visceral afferent stimulation) (Appleyard et al., 2007; Sumal et al., 1983) are inhibited by bath application of ghrelin (Cui et al., 2011).

Hindbrain integration and processing of vagal- and blood-borne energy status signals

Complementing the processing and integration performed by the vagus nerve, the encoding, differentiation, and integration of the array of vagally- or blood-borne energy status signals also involves processing by NTS neurons. The following hypothesized mechanisms may explain distinct neuronal, behavioral, and physiological responses to gut peptides and/or circulating hormones and nutrients. These NTS-mediated processing-mechanisms would include, but are not limited to the: 1) neuroanatomical pattern and percent of NTS neurons activated [see (Peters et al., 2011) for example], 2) intensity, frequency, and duration of neuronal activation, 3) phenotype(s) of NTS neurons activated, and 4) vagally-mediated neurotransmitter(s) that activates NTS neurons. Each of these mechanistic roles in NTS processing of energy balance signals has been reviewed elsewhere [see (Berthoud, 2004; Grill, 2010; Moran and Ladenheim, 2011; Powley, 2000; Rinaman, 2010; Ritter, 2004; Schwartz, 2010) for review]. Here, we review select cellular signaling mechanisms that occur within individual NTS neurons in response to vagal and/or blood-borne energy status signals that may account for integration of multiple energy status signals. Below, we also review the role that descending projections from forebrain structures have on NTS-mediated processes for energy balance control.

Intracellular signaling pathways as a point of convergence for NTS-mediated anorectic signals

Recent evidence suggests that the intracellular signaling pathways within NTS neurons provide points of convergence and antagonism for various vagally-mediated satiation signals and for blood-borne signals of energy availability. One mechanism hypothesized to account for interactions between different energy status signals is that activation of separate receptors expressed on a given neuron could lead to an enhanced response via converging intracellular signaling pathways (Berthoud et al., 2006; Daniels et al., 2005; Hayes et al., 2011). It is also possible that activation of a single receptor expressed on a neuron may produce long-term changes in gene transcription and protein synthesis that could potentiate the intake inhibitory effects of other anorectic systems that also act upon the same neurons. Indeed, NTS mediation of the intake inhibitory responses of a number of anorectic signals [e.g. CCK, melanocortin, leptin, amylin, GLP-1] involves some of the same intracellular signaling pathways (e.g. PKA, MAPK, AMPK) (Boyle et al., 2011; Hayes et al., 2011; Hayes et al., 2009b; Hayes et al., 2010b; Sutton et al., 2005; Sutton et al., 2004).

NTS integration of LepRb signaling and GI-derived satiation signals provides an instructive example. Neuropharmacological studies show hindbrain-delivered leptin amplifies the food intake-suppressive effect of gastric distention (Huo et al., 2007). Complementing these findings are results showing that mNTS/AP targeted RNAi mediated knockdown of LepRb (LepRb KD) abolished the potent food intake-suppressive effects of CCK (Hayes et al., 2010b). Recent findings from our laboratory also show that the hyperphagia induced by mNTS/AP LepRb KD results primarily from an increase in meal size (Kanoski et al., 2012). This result is consistent with the explanation of the hyperphagia associated with global LepRb deficiency in rats [Koletsky (fak / fak)] (Castonguay et al., 1982; Morton et al., 2005) and leptin deficiency in mice (ob/ob) (Ho and Chin, 1988). In both cases the obese phenotype is driven by hyperphagia associated with increased meal size and not by increased meal frequency. Thus, the LepRb processing in the mNTS appears to be a critical mediator of this meal size-regulatory effect.

In addition to the required endogenous role of LepRb signaling in the mNTS to energy balance regulation (Hayes et al., 2010b), GLP-1Rs expressed on NTS neurons are also critically involved in food intake regulation (Hayes et al., 2009a; Hayes et al., 2011). Recent data, using various dose combinations of leptin and the GLP-1R agonist exendin-4, show that hindbrain GLP-1R and LepRb signaling interact in control of food intake and body weight in at least an additive fashion (Zhao et al., 2012). Moreover, blockade of hindbrain GLP-1R attenuated the intake inhibitory effects of hindbrain leptin. Together with previous reports (Hayes et al., 2009a; Hayes et al., 2010b), these findings suggest that the interaction between hindbrain LepRb and GLP-1R signaling may be physiologically relevant for the normal control of food intake.

Common intracellular signaling pathways contribute to the intake suppression produced by CNS LepRb or GLP-1R activation, including the mitogen-activated protein kinase (MAPK) (Bjorbaek et al., 2001; Hayes et al., 2011) and AMPK (Gao et al., 2007; Hayes et al., 2009b; Hayes et al., 2010b; Minokoshi et al., 2004). The role of reduced AMPK signaling in mediating the intake inhibitory response to mNTS LepRb signaling is established as a required pathway using conventional pharmacological activators (i.e. AICAR) and inhibitors (i.e. compound C) (Hayes et al., 2009b; Hayes et al., 2010b), and also through selective mNTS LepR knockdown (Hayes et al., 2010b). Similarly, mNTS GLP-1R signaling potently suppress food intake and does so in part through an AMPK-dependent pathway. That is, mNTS GLP-1R activation suppresses food intake via a coordinated PKA-dependent activation of the p44/42MAPK pathway and simultaneous inhibition of the AMPK pathway. It is hypothesized that the combined activation of common intracellular signaling pathways (e.g. AMPK), putatively within the same NTS neuron contributes to the aforementioned additive suppression of food intake by hindbrain leptin and GLP-1R signaling. Such an effect may also increase the sensitivity of NTS neurons to other satiation signals (e.g. amylin, CCK, and gastric distension) whose intake suppressive effect may also involve these same intracellular signaling pathways [see (Berthoud et al., 2006; Grill, 2010; Hayes et al., 2010a; Potes and Lutz, 2010) for review].

II. Output: Hindbrain Neurons are Critical Contributors to Response Control

The contribution of hindbrain neurons to many aspects of energy balance response control is discussed in the following subsection. The source of the input engaging hindbrain output-related neurons differs for the various types of responses (See Figure 1b). For some responses, like gastric emptying control, incretin effects or taste-driven oral motor feeding responses, the response driving input arises from cranial afferents such as vagal afferents or cranial nerve innervation of the taste receptor cells. For other responses, like brown fat thermogenesis, the inputs to hindbrain premotor neurons arise from many sources including but not limited to forebrain neurons and spinal thermoreceptors.

Feeding Behavior

Craig (Craig, 1918) distinguishes between a contact or consummatory phase for feeding behavior that is preceded by a search or appetitive phase for feeding and other motivated behaviors. Hindbrain neurons are central to the control of the ingestive consummatory behaviors that are driven by oral contact with food. Both the primary motor neurons that control movements of the tongue, jaw, pharynx, and facial muscles and the premotor neurons that constitute the central pattern generators for patterned rhythmic oral motor behaviors such as licking, chewing, swallowing, and food rejection are located in the hindbrain [see (Blessing, 1997) for details and illustrations). Work of Travers and colleagues identify neurons in the intermediate and parvocellular regions of the lateral medullary reticular formation (RF), subjacent to the rostral nucleus of the solitary tract (rNTS), as the site of the premotor neurons that govern the production of patterned ingestion and rejection responses and the transition between ingestion and rejection (Chen and Travers, 2003; Chen et al., 2001; Venugopal et al., 2010). This hindbrain substrate for premotor control of food ingestion and rejection response is the target of descending projections from hypothalamic and other forebrain nuclei whose neurons respond to energy status signals (Gutierrez et al., 2006; Sawchenko, 1998).

Input to the ingestive premotor neurons also arises from caudal brainstem (midbrain and hindbrain) neurons. The importance of caudal brainstem input to hindbrain ingestive premotor neurons is highlighted by procedures that eliminate all forebrain-caudal brainstem bidirectional communication. Such studies reveal that the taste-elicited, oral motor responses of chronically maintained decerebrate rats are comparable to those of intact brain control rats (Grill and Norgren, 1978; Kaplan and Grill, 1989). Additional support for the conclusion that caudal brainstem neural circuits are sufficient for the food contact driven ingestive consummatory behavior comes from clinical data showing that taste-elicited ingestion and rejection responses of anencephalic human neonates (e.g., brains not developed beyond the midbrain) are identical to those of normal neonates (Steiner, 1973). While data from these special cases (decerebrate rats, anencephalic neonates) provide conclusive support for the sufficiency of caudal brainstem circuits in ingestive consummatory response control, they do not speak to a potentially powerful contribution of descending projections from forebrain neurons in the control of feeding behavior in normal animals (see section III, below). Defining which energy balance relevant forebrain neurons connect directly or indirectly to lateral medullary RF neurons or to other hindbrain neurons contributing to response control will be useful in this regard. A later section of the review highlights functional links between hypothalamic [e.g., lateral hypothalamic (LH) and paraventricular hypothalamic (PVH)] axons and NTS neurons.

Nutrient transit, absorption, partitioning, storage, and mobilization

Once ingested, food is transported from the mouth to the stomach and from the stomach to the intestine, the principal site of nutrient absorption. The premotor neurons that contribute to the control of nutrient partitioning, storage, and mobilization by modulating the rate of gastric emptying and influencing nutrient absorption are found in the dorsal motor nucleus of the vagus nerve (DMV) containing parasympathetic vagal efferent neurons, and also in aspects of the sympathetic nervous system [e.g. (Ahren, 2000)]. DMV neurons located in the dorsomedial medulla just ventral to the NTS provide vagal efferent output that controls gastric emptying rates, pancreatic and intestinal enzymatic secretions, and intestinal motility rates. Together, manipulation of these physiological responses influences the exposure of the chemical and nutritive properties of the ingested food to the enteroendocrine cells of the intestine affecting the pattern and magnitude of gastrointestinal and pancreatic hormone secretion (Berthoud, 2008; Reimann et al., 2008). Using ultrastructural analysis, Rinaman et al. (Rinaman et al., 1989) described a monosynaptic gastric vago-vagal circuit that provides an anatomical substrate for the relay of gastric afferent information directly to gastric vagal motor neurons, Such vago-vagal reflexes consist of three components: sensory vagal afferents, second-order integrative neurons of the NTS, and efferent vagal neurons of the DMV. Travagli et al. (Travagli et al., 2003) describe CNS modulation of vago-vagal reflexes via hindbrain thyrotropin-releasing hormone input to the NTS. Abdominal GLP-1R driven incretin and gastric emptying effects are mediated by the paracrine action of released GLP-1 on vagal afferent neurons (Holst, 2007). NTS neurons process this information and relay it to a variety of CNS sites including hindbrain neurons (Hayes et al., 2008). Vagal efferent stimulation of pancreatic islet cells results in insulin secretion and thereby provides another source of nutrient storage control (Ahren, 2000). Many anatomically distributed CNS sites can directly modulate DMV-mediated parasympathetic output controlling for gastric emptying rates and pancreatic projecting signals modulating counterregulatory hormone secretions, e.g., (Jiang et al., 2003; Rinaman et al., 1989; Rogers et al., 1980; Zheng et al., 2005).

Taste and other cephalic sensory stimuli contribute to normal glucose tolerance via a neurally mediated, cephalic phase of insulin secretion that involves the vagal efferent input to the β-cell (Ahren, 2000; Siegel et al., 1980). Caudal brainstem neuronal projections to DMV neurons are sufficient to mediate taste-driven cephalic insulin secretion when descending hypothalamic projections are surgically eliminated (Flynn et al., 1986). A cephalic phase regulation of glucagon secretion has also been found in a variety of species (Teff and Engelman, 1996) and can be triggered by glucopenia (Havel and Taborsky, 1989) directed to the dorsomedial medulla in intact rats (Andrew et al., 2007). Identifying additional sources of forebrain input to the DMV neurons that provide CNS control of nutrient absorption, storage, and mobilization is necessary for a more complete picture of neural control of energy balance.

Glucose homeostasis

Glucose is the primary energy substrate for CNS neurons. Claude Bernard's (Bernard, 1855) report of hyperglycemia induced by probing the medulla suggests that the CNS contributes to the control of blood glucose via stimulation of the sympathetic outflow (Feldberg et al., 1985), an idea that is still widely accepted today (Jordan et al., 2010; Marty et al., 2007). This section reviews findings that link hindbrain neurons to the control of sympathetic and behavioral responses that mediate hepatic glucose production and counterregulation. The related topic, the sensing of blood glucose itself or of correlates of cellular energy availability, was discussed in the section on hindbrain energy status signal processing above.

Counterregulation

Support for the conclusion that hindbrain neurons play an important role in engaging the sympathoadrenal, glucagon, and corticosterone elements of counterregulation comes from a variety of laboratories. Using the induction of cellular glucopenia via direct parenchymal application of glucose analogues (e.g., 5-thioglucose [5TG]), S. Ritter and colleagues show that it is the catecholaminergic (CA) neurons in the medulla that are key trigger zones for sympathoadrenal hyperglycemia, glucagon, and corticosterone secretion, and for food intake responses driven by glucopenia or hypoglycemia. Given the historical focus on hypothalamic energy sensing, e.g., (Mayer, 1953) it is interesting to note that careful mapping of a broad range of hypothalamic nuclei with targeted glucopenia fails to trigger any counterregulatory responses other than corticosterone (Berthoud and Mogenson, 1977; Ritter et al., 2000) cf, (Borg et al., 1995). Using immunohistochemical detection of systemically injected 2-deoxyglucose (2DG)-induced c-Fos activation colocalized with the enzymes TH and phenethanolamine-N-transferase as markers for adrenergic and noradrenergic neurons respectively, the same group concludes that it is primarily adrenergic neurons located in C2 (NTS) and C3 cell groups of the dorsal medulla, as well as the A1/C1 neurons of the (VLM) that are the targets of activation by systemic glucopenia (Ritter et al., 1998). These neurons are thereby key elements in the CNS circuit controlling glucose homeostasis

To identify the neurons receiving the axonal projections of these hindbrain CA neurons these investigators utilized a toxin-antibody complex: saporin conjugated to dopamine beta hydroxylase (DSAP). Injection of DSAP into the PVH would presumably lesion those hindbrain CA neurons with PVH projecting axons, whereas DSAP injected into the interomediolateral cell column in the spinal cord would lesion hindbrain CA that, based on their spinal projection, could be considered sympathetic premotor neurons (Ritter et al., 2001; Ritter et al., 2006; Ritter et al., 2003). Two studies employing PVH DSAP injections conclude that PVH projecting hindbrain CA neurons are required for the glucopenia triggered feeding response (Hudson and Ritter, 2004; Ritter et al., 2001). Importantly however, other data are inconsistent with this conclusion. Two studies show that cytoglucopenia (systemic 2DG) or hypoglycemia (systemic insulin) increases the intake of an orally infused diet, relative to a control treatment, in decerebrate rats whose forebrain projecting axons were surgically severed; intact control rats performed similarly (Darling and Ritter, 2009; Flynn and Grill, 1983). These results and other data suggest that: [1] axons of individual hindbrain CA neurons implicated in the control of counterregulation project to multiple targets (Banihashemi and Rinaman, 2006) and [2] despite severing of forebrain projecting axons of such multi-target CA neurons (i.e., decerebration), their axon projections within the caudal brainstem sustain function [see (Rinaman, 2003) for discussion]. These findings specify the intra-caudal brainstem projection targets of hindbrain CA neurons as critical elements for feeding response controls. One potentially relevant intra-hindbrain target of these glucopenia responsive hindbrain CA neuron are the medullary RF ingestive consummatory premotor neurons (Day et al., 1997). The contribution of hindbrain CA neurons to the neuroendocrine response aspects of counterregulation (e.g., corticosterone secretion) is discussed next.

Excitation of the HPA axis arises from distinct neuron populations including several in the hindbrain and amygdala (Herman and Cullinan, 1997). Notable among these are hindbrain CA neurons in A2/C2 (NTS) and A1 (VLM) that project directly to corticotrophin releasing factor (CRF)-containing PVH neurons (Cunningham and Sawchenko, 1988). Hindbrain CA projections are functionally relevant for HPA hormonal response to various stimuli including hemorrhage/hypotension, respiratory distress, immune challenge, and counterregulation triggered by hypoglycemia/glucopenia. CA deafferentation markedly reduces corticosterone release triggered by insulin or 2-DG treatment (but not the release by circadian control or by the forced-swim stressor) and prevents 2-DG from elevating PVH CRH expression (Ritter et al., 2003). Hypothalamically projecting hindbrain CA neurons express a variety of chemical signals [epinephrine, norepinephrine (NE), neuropeptide Y, or galanin]; direct PVH injection of NE mimics the PVH neuroendocrine responses to glycemic challenge (Khan et al., 2007). Interestingly, PVH neurons send reciprocal connections to A1/C1 VLM neurons and to mNTS neurons including A2 neurons (Geerling et al., 2010).

Hepatic Glucose Production

There is accumulating evidence that the vagal efferent innervation of the liver, with premotor neurons located in hindbrain DMV, contributes to the control of glucose homeostasis and of energy balance more broadly. A variety of recent reports indicate that stimulation of CNS nutrient sensors by various correlates of energy repletion, such as long chain coenzyme A, glucose, insulin, and leptin reduce hepatic glucose production through a neural pathway involving vagal efferent transmission. For example, hepatic glucose production is reduced following ventricular infusion of oleic acid (Obici et al., 2002) and involves a mechanism that includes changes in fatty acid β-oxidation, the activation of CNS KATP channels, and hepatic vagal efferent signaling (Lam et al., 2005; Obici et al., 2003; Pocai et al., 2005). The functional inputs to hindbrain DMV neurons that engage hepatic vagal efferent traffic likely include input from so-called nutrient sensors located in the hypothalamus e.g., (Spanswick et al., 1997) and elsewhere, including the hindbrain itself (Ritter et al., 2011).

Energy Expenditure

Brown adipose tissue (BAT) thermogenesis, a major thermogenic effector for energy balance control and for thermoregulation, is controlled by CNS sympathetic premotor neurons and adrenergic receptors on brown adipocytes (Bachman et al., 2002; Morrison, 2011). Exciting new data demonstrate the presence of BAT and its energetic function in adult humans [e.g., (Ouellet et al., 2011)]. In animal models, pseudo-rabies virus transynaptic tracer injections in BAT labels a broad network of CNS sites participating in the sympathetic efferent control of BAT thermogenesis (Bamshad et al., 1999). Various studies prominently highlight the critical role for the hindbrain sympathetic premotor neurons located in the raphe pallidus (RPa) and their modulation by axonal projections arising from several neuronal groups in the hypothalamus (medial preoptic nucleus, dorsomedial nucleus) and the medulla (NTS, VLM) (Cao et al., 2010; Morrison, 2011). Data from decerebrate rats lacking forebrain- caudal brainstem communication show elevated BAT temperature in response to RPa injection of melanocortin receptor agonist (Skibicka and Grill, 2008); these rats maintain core temperature in response to most cold environments (Nautiyal et al., 2008). These findings support the hypothesis that caudal brainstem integrators and endemic effector circuits mediate some degree of energy expenditure function.

III. Hindbrain Neurons Receive and Integrate Descending Information Arising from Rostral Neural Sites that Contribute to Response Control

In addition to the integrations performed by hindbrain neurons involving direct action of energy status signals (vagal and blood-borne), these neurons receive and process input (axonal projections) from neurons in other parts of the brain that also receive and process energy status signals (see Figure 1b and 1a). Some examples are considered here in which descending forebrain/hypothalamic inputs to NTS neurons contribute to the control of feeding behavior.

LH Orexinergic Projections to NTS

Orexin/hypocretin-containing neurons, uniquely expressed in the LH, project throughout the neuraxis and participate in the control of a variety of functions including the sleep/wake cycle, autonomic control, and energy balance. As many as one quarter of these neurons project to the hindbrain nuclei (NTS, caudal raphe, and VLM) (Ciriello et al., 2003; Zheng et al., 2005). Orexin-immunoreactive (ir) fibers are in close apposition to TH positive neurons in the A2/C2 region (Zheng et al., 2005) and electron microscopy confirms synaptic junctions between orexin fibers and CA-expressing NTS neurons (Puskas et al., 2010). Zheng et al. (Zheng et al., 2005) also show that varicose orexin-ir fibers are in close apposition to mNTS neurons that express c-Fos following gastric nutrient infusion. Given the established literature showing that electrical or chemical stimulation of LH neurons elicits feeding behavior in satiated rats (Berthoud and Munzberg, 2011; Coons et al., 1965), this anatomical arrangement suggests a possible mediating mechanism for the LH-stimulated feeding – orexin acting on NTS neurons attenuates processing of vagally mediated GI satiation signals. The concept is supported, in principal, by results showing that LH electrical stimulation (not unique to orexin neurons) inhibits mNTS neurons and that the excitatory response of intestinal or gastric distension on mNTS neurons is reversed by concurrent LH stimulation (electrical or glutamate) stimulation (Jiang et al., 2003).

ARH Melanocortinergic Projections to NTS

The central melanocortin (MC) system is critical to energy balance control as humans and mice with MC4-R mutation are hyperphagic and obese (Hinney et al., 1999; Huszar et al., 1997). NTS and DMV are among the neurons with the densest MC4-R expression (Kishi et al., 2003). These hindbrain nuclei receive MC agonist ligands (e.g. αMSH) from ARH and from endemic NTS proopiomelanocrotin (POMC) neurons (Zheng et al., 2010), as well as antagonist from other ARH neurons (Broberger et al., 1998). NTS delivery of MC4-R agonist reduces food intake and elevates core temperature and heart rate, whereas antagonist delivery has opposite effects (Skibicka and Grill, 2009b; Williams et al., 2000). Interestingly, the effect of hindbrain MC4-R ligand delivery on meal patterning mirrors effects of GI satiation signals like CCK. Specifically, meal size is reduced by hindbrain delivery of a MC3/4-R agonist whereas hindbrain delivery of a MC3/4-R antagonist increases meal size (Sutton et al., 2005), suggesting to some that hindbrain MC4-R are downstream mediators of CCK-induced intake suppression (Fan et al., 2004). However, others have reported that diminished sensitivity to GI-derived satiation signals (e.g. CCK, bombesin, nutrient preload) is not responsible for mediating the chronic hyperphagia of MC4-R knockout mice (Vaughan et al., 2006).

ARH-NTS melanocortinergic circuits contribute to the food intake suppressive effect of ARH leptin signaling

NTS projecting ARH MC axons appear to be one contributor to the neural circuit mediating the intake suppressive effect of hypothalamic leptin signaling, as MC4-R antagonist injected into NTS attenuates the intake suppressive effect of leptin delivery to ARH (Zheng et al., 2010). Similarly, hindbrain MC4-R signaling is downstream of the effects of hindbrain leptin signaling; a dose of hindbrain MC4-R antagonist without effects of its own reverses the food intake and the thermogenic effects of hindbrain leptin delivery (Skibicka and Grill, 2009a). ARH POMC neurons appear to make another contribution to the intake inhibitory effect of ARH leptin signaling that involves projections to PVH neurons than in turn project to the NTS. ARH melanocortinergic neurons project to MC4-R bearing PVH neurons that express a variety of hormones including oxytocin (OT) (Liu et al., 2003). A large percentage of the OT-expressing neurons send axons to the hindbrain (Rinaman, 1998).

PVH Oxytocinergic projections to NTS

The axons of OT-expressing parvocellular PVH (pPVH) hypothalamic neurons project throughout the neuraxis (Rinaman, 1998; Sawchenko and Swanson, 1982). OT axons are in close apposition to mNTS neurons that express Fos-ir in response to CCK injection (Blevins et al., 2003). A functional relationship between OT and GI satiation signaling is suggested by the attenuation of the intake suppressive action of systemic CCK following hindbrain microinjection of OT receptor antagonist (Blevins et al., 2003; Olson et al., 1991). The intake inhibitory effect of oleoylethanolamide (OEA), another vagally-mediated gut signal, is also attenuated by hindbrain OT antagonist (Gaetani et al., 2010). Recent data (Peters et al., 2008) suggest a cellular mechanism for the influence of NTS OT signaling on the intake inhibition of NTS processed GI vagal signals. NTS neurons were neurophysiologically identified by glutamate driven excitatory postsynaptic currents (EPSCs) induced by solitary tract (axons of vagal afferents) stimulation. OT application to the slice increased the frequency of miniature EPSCs in approximately half of the NTS neurons examined. Analysis of the solitary tract-evoked EPSCs indicates that OT increased the release of glutamate from the vagal afferents terminating on the NTS neurons. The findings suggest that OT release acts on a subset of mNTS neurons to enhance visceral afferent transmission via pre-synaptic and post-synaptic mechanisms.

OT projections to NTS contribute to hypothalamic energy status signaling effects on feeding

Blevins et al. (Blevins et al., 2004) investigate the hypothesis that the intake suppressive effect of hypothalamic leptin signaling involves an ARH POMC-to pPVH OT-to NTS circuit. They show that 3rd icv leptin induces c-Fos in PVH OT neurons that are labeled by NTS tracer injection, and that OT antagonist (forebrain ventricle; hindbrain delivery not described) attenuates leptin-induced anorexia. Other hormones expressed by PVH neurons may also contribute to hypothalamic leptin signaling induced feeding inhibition via their projections to NTS including, but not limited to thyrotropin releasing hormone and CRF (Ahima et al., 1996; Harris et al., 2001; Uehara et al., 1998). The intake inhibitory effect of hypothalamic leucine infusion is attenuated by hindbrain icv OT antagonist (Blouet et al., 2009). Similarly, the food intake suppressive effects of nesfatin-1, a food intake suppressing peptide localized to pPVH neurons that also express OT, is reversed by hindbrain icv OT antagonist delivery (Maejima et al., 2009). Collectively, these findings suggest a mechanism by which hindbrain (NTS) processing is required to mediate the intake inhibitory effects of various hypothalamic energy status signals, and that the conduit of such communication involves descending OT communication. More work is required to substantiate this hypothesis.

IV. Ascending Hindbrain Projections to Midbrain/Forebrain Neurons are Essential for the Neural Control of Food Intake

The literature on the neurobiology of food reward focuses appropriately on gustatory signaling from the oral cavity and its synaptic connections to mesolimbic structures associated with reward processing (MRS) [see (Norgren et al., 2006) for review]. The ascending central gustatory communication begins with neurons in the rostral NTS that receive cranial nerve afferents that innervate taste receptor cells. In the rodent, these neurons project in turn to the parabrachial nucleus (PBN). PBN axons then convey gustatory afferent activity to the medial extension of the thalamic trigeminal relay, or along a separate pathway to the LH, central nucleus of the amygdala (CeA), and the bed nucleus of the stria terminalis (BNST) (Norgren, 1976, 1978; Norgren et al., 2006). The gustatory region of PBN projects to limbic structures - CeA and BNST - that are considered to be primarily responsible for subsequently engaging the nucleus accumbens (NAc) to modulate dopamine signaling (Norgren et al., 2006). While these discoveries define the feed-forward neural circuitry that promotes food intake, largely unexplored are the intake inhibitory feedback signals that accumulate as consumption of a meal progresses and affect MRS neural processing. It is likely that these inhibitory signals are communicated via a CNS relay between GI-derived vagally mediated satiation signals and nuclei of the MRS. The recent discovery of monosynaptic connections from the caudal NTS to the ventral tegmental area (VTA) and NAc (Alhadeff et al., 2011; Dossat et al., 2011; Rinaman, 2010) provides such a conduit.

The processing and integration of gastric- and intestinal-inhibitory signals beginning within the NTS in turn counteracts the intake stimulatory gustatory signals from the oral cavity. The notion that post-ingestive signals modulate gustatory reward has been established for some time (Babcock et al., 1985; Carr and Simon, 1984; Sclafani and Ackroff, 2004); however, the mediating systems and neuronal pathways remain poorly characterized. One possibility is that the neuronal circuitry required for GI inhibitory control of feeding is endemic to the caudal brainstem. Evidence for this hypothesis comes from studies using chronic decerebrate rats showing that the intake inhibitory response to gastric nutrient preloads, as well as intestinally derived satiation signals is equivalent in neurologically intact rats and in rats with neural processing restricted to the isolated caudal brainstem (Grill and Smith, 1988; Hayes et al., 2008; Seeley et al., 1994). As decerebrate rats are unable to engage in appetitive feeding, these findings establish a forebrain-independent neural circuitry that mediates the intake inhibitory effects on GI satiation signals on consummatory phase of feeding control. These data do not address, however, neural control of the appetitive phase of feeding, but they do offer a foundation for an additional hypothesis to explain how GI-derived satiation signals may modulate gustatory reward: namely, that processing and integration of GI satiation signals within the NTS (or hindbrain more broadly), modulate the higher-order cognitive-centers of the brain involved in the appetitive control of feeding.

The idea that GI-derived vagally mediated satiation signals can modulate the rewarding value of food via direct NTS-relayed monosynaptic projections to nuclei of the MRS (i.e. VTA and NAc) is novel and provides a rapid temporal modulation of within-meal food intake control (see Figure 1c). Monosynaptic connections from the caudal NTS to the MRS are established (Alhadeff et al., 2011; Dossat et al., 2011; Ricardo and Koh, 1978; Rinaman, 2010). Among the possible phenotypes of NTS neurons considered, GLP-1 synthesizing preproglucagon (PPG; precursor peptide for GLP-1) neurons stand out as an intriguing possibility for a number of reasons, including that these NTS PPG neurons are physiologically required for the normal control of food intake and body weight (Barrera, 2011). While immunocytochemisty and in situ hybridization confirm the presence of GLP-1R mRNA and GLP-1 immunopositive fibers, respectively, in both the VTA and NAc (Merchenthaler et al., 1999; Rinaman, 2010), only recently have two, independent labs provided direct evidence that NTS PPG neurons are in fact projecting monosynaptically to the MRS (Alhadeff et al., 2011; Dossat et al., 2011). Double IHC for retrograde tracers injected in the VTA, NAc core, or NAc shell together with PPG-expressing neurons in the caudal NTS provide a previously unidentified mechanism by which central GLP-1 can regulate food intake. Moreover, these GLP-1-to-MRS projections are physiologically relevant for food intake control, as blockade of GLP-1R in the VTA (Alhadeff et al., 2011) or NAc core (Alhadeff et al., 2011; Dossat et al., 2011) increased food intake. The GLP-1R signaling in either the VTA or NAc was shown to preferentially reduced intake of palatable foods and did so in the absence of nausea (Alhadeff et al., 2011). Given that within-meal vagally mediated satiation signals (i.e. CCK and gastric distension) activate NTS PPG neurons (Hayes et al., 2009a; Hisadome et al., 2011; Vrang et al., 2003), these findings provide a mechanism by which hindbrain processing of satiation signals could communicate with higher-order regions of the forebrain, and subsequently modulate food intake by decreasing the rewarding value of the ongoing meal. Of course, PPG neurons are not the sole phenotype of NTS neuron that project to reward processing associated structures. Indeed, work from Aston-Jones’ laboratory has concluded that the primary source of noradrenergic afferents to the NAc originates in the A2 region of the NTS (Delfs et al., 1998). Further identification of NTS neuron phenotypes whose axons project to forebrain nuclei will be useful in better characterizing the mechanisms by which hindbrain processing of energy status signals modulates appetitive processes of food intake control.

Conclusions

The case is made that hindbrain nuclei perform integrations that are essential to the control of behavioral, autonomic, and endocrine responses that are marshaled to collectively control energy balance. These hindbrain-mediated response controls involve endemic hindbrain circuits, as well as the participation of midbrain and forebrain nuclei and peripheral circuits. Thus, the hindbrain is viewed here as one key element in the multi-site, neuroanatomically distributed control of energy balance function.

Acknowledgements

We thank Dr. S. Kanoski for his critical reading of the manuscript and Samantha Fortin for developing the concept of the figure. We acknowledge our research support from the NIH; DK21397 (HJG), and DK085435 (MRH).

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  1. Ahima RS, Prabakaran D, Mantzoros C, Qu D, Lowell B, Maratos-Flier E, Flier JS. Role of leptin in the neuroendocrine response to fasting. Nature. 1996;382:250–252. doi: 10.1038/382250a0. [DOI] [PubMed] [Google Scholar]
  2. Ahren B. Autonomic regulation of islet hormone secretion--implications for health and disease. Diabetologia. 2000;43:393–410. doi: 10.1007/s001250051322. [DOI] [PubMed] [Google Scholar]
  3. Alhadeff AL, Rupprecht LE, Hayes MR. GLP-1 neurons in the nucleus of the solitary tract project directly to the ventral tegmental area and nucleus accumbens to control for food intake. Endocrinology. 2011 doi: 10.1210/en.2011-1443. In Press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Andrew SF, Dinh TT, Ritter S. Localized glucoprivation of hindbrain sites elicits corticosterone and glucagon secretion. Am J Physiol Regul Integr Comp Physiol. 2007;292:R1792–1798. doi: 10.1152/ajpregu.00777.2006. [DOI] [PubMed] [Google Scholar]
  5. Appleyard SM, Marks D, Kobayashi K, Okano H, Low MJ, Andresen MC. Visceral afferents directly activate catecholamine neurons in the solitary tract nucleus. J Neurosci. 2007;27:13292–13302. doi: 10.1523/JNEUROSCI.3502-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Arluison M, Quignon M, Nguyen P, Thorens B, Leloup C, Penicaud L. Distribution and anatomical localization of the glucose transporter 2 (GLUT2) in the adult rat brain--an immunohistochemical study. J Chem Neuroanat. 2004;28:117–136. doi: 10.1016/j.jchemneu.2004.05.009. [DOI] [PubMed] [Google Scholar]
  7. Babcock AM, Livosky M, Avery DD. Cholecystokinin and bombesin suppress operant responding for food reward. Pharmacol Biochem Behav. 1985;22:893–895. doi: 10.1016/0091-3057(85)90543-x. [DOI] [PubMed] [Google Scholar]
  8. Bachman ES, Dhillon H, Zhang CY, Cinti S, Bianco AC, Kobilka BK, Lowell BB. betaAR signaling required for diet-induced thermogenesis and obesity resistance. Science. 2002;297:843–845. doi: 10.1126/science.1073160. [DOI] [PubMed] [Google Scholar]
  9. Bado A, Levasseur S, Attoub S, Kermorgant S, Laigneau JP, Bortoluzzi MN, Moizo L, Lehy T, Guerre-Millo M, Le Marchand-Brustel Y, et al. The stomach is a source of leptin. Nature. 1998;394:790–793. doi: 10.1038/29547. [DOI] [PubMed] [Google Scholar]
  10. Bady I, Marty N, Dallaporta M, Emery M, Gyger J, Tarussio D, Foretz M, Thorens B. Evidence from glut2-null mice that glucose is a critical physiological regulator of feeding. Diabetes. 2006;55:988–995. doi: 10.2337/diabetes.55.04.06.db05-1386. [DOI] [PubMed] [Google Scholar]
  11. Bamshad M, Song CK, Bartness TJ. CNS origins of the sympathetic nervous system outflow to brown adipose tissue. Am J Physiol. 1999;276:R1569–1578. doi: 10.1152/ajpregu.1999.276.6.R1569. [DOI] [PubMed] [Google Scholar]
  12. Banihashemi L, Rinaman L. Noradrenergic inputs to the bed nucleus of the stria terminalis and paraventricular nucleus of the hypothalamus underlie hypothalamic-pituitary-adrenal axis but not hypophagic or conditioned avoidance responses to systemic yohimbine. J Neurosci. 2006;26:11442–11453. doi: 10.1523/JNEUROSCI.3561-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Barrera JG. Hyperphagia and increased fat accumulation in two models of chronic CNS glucagon-like Peptide-1 loss of function. J Neurosci. 2011;31:3904–3913. doi: 10.1523/JNEUROSCI.2212-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Batterham RL, ffytche DH, Rosenthal JM, Zelaya FO, Barker GJ, Withers DJ, Williams SC. PYY modulation of cortical and hypothalamic brain areas predicts feeding behaviour in humans. Nature. 2007;450:106–109. doi: 10.1038/nature06212. [DOI] [PubMed] [Google Scholar]
  15. Bernard C. Lecons de Physiologie Experimentale Appliquee a la Medecine. Bailliere; Paris: 1855. [Google Scholar]
  16. Berthoud HR. The caudal brainstem and the control of food intake and energy balance. Plenum; New York: 2004. [Google Scholar]
  17. Berthoud HR. The vagus nerve, food intake and obesity. Regul Pept. 2008;149:15–25. doi: 10.1016/j.regpep.2007.08.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Berthoud HR, Blackshaw LA, Brookes SJ, Grundy D. Neuroanatomy of extrinsic afferents supplying the gastrointestinal tract. Neurogastroenterol Motil 16 Suppl. 2004;1:28–33. doi: 10.1111/j.1743-3150.2004.00471.x. [DOI] [PubMed] [Google Scholar]
  19. Berthoud HR, Mogenson GJ. Ingestive behavior after intracerebral and intracerebroventricular infusions of glucose and 2-deoxy-D-glucose. Am J Physiol. 1977;233:R127–133. doi: 10.1152/ajpregu.1977.233.3.R127. [DOI] [PubMed] [Google Scholar]
  20. Berthoud HR, Munzberg H. The lateral hypothalamus as integrator of metabolic and environmental needs: from electrical self-stimulation to optogenetics. Physiol Behav. 2011;104:29–39. doi: 10.1016/j.physbeh.2011.04.051. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Berthoud HR, Sutton GM, Townsend RL, Patterson LM, Zheng H. Brainstem mechanisms integrating gut-derived satiety signals and descending forebrain information in the control of meal size. Physiol Behav. 2006;89:517–524. doi: 10.1016/j.physbeh.2006.08.018. [DOI] [PubMed] [Google Scholar]
  22. Bjorbaek C, Buchholz RM, Davis SM, Bates SH, Pierroz DD, Gu H, Neel BG, Myers MG, Jr., Flier JS. Divergent roles of SHP-2 in ERK activation by leptin receptors. J Biol Chem. 2001;276:4747–4755. doi: 10.1074/jbc.M007439200. [DOI] [PubMed] [Google Scholar]
  23. Blessing W. The Lower Brainstem and Bodily Homeostasis. Oxford Press; New York: 1997. [Google Scholar]
  24. Blevins JE, Eakin TJ, Murphy JA, Schwartz MW, Baskin DG. Oxytocin innervation of caudal brainstem nuclei activated by cholecystokinin. Brain Res. 2003;993:30–41. doi: 10.1016/j.brainres.2003.08.036. [DOI] [PubMed] [Google Scholar]
  25. Blevins JE, Schwartz MW, Baskin DG. Evidence that paraventricular nucleus oxytocin neurons link hypothalamic leptin action to caudal brain stem nuclei controlling meal size. Am J Physiol Regul Integr Comp Physiol. 2004;287:R87–96. doi: 10.1152/ajpregu.00604.2003. [DOI] [PubMed] [Google Scholar]
  26. Blouet C, Jo YH, Li X, Schwartz GJ. Mediobasal hypothalamic leucine sensing regulates food intake through activation of a hypothalamus-brainstem circuit. J Neurosci. 2009;29:8302–8311. doi: 10.1523/JNEUROSCI.1668-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Borg WP, Sherwin RS, During MJ, Borg MA, Shulman GI. Local ventromedial hypothalamus glucopenia triggers counterregulatory hormone release. Diabetes. 1995;44:180–184. doi: 10.2337/diab.44.2.180. [DOI] [PubMed] [Google Scholar]
  28. Boyle CN, Rossier MM, Lutz TA. Influence of high-fat feeding, diet-induced obesity, and hyperamylinemia on the sensitivity to acute amylin. Physiol Behav. 2011;104:20–28. doi: 10.1016/j.physbeh.2011.04.044. [DOI] [PubMed] [Google Scholar]
  29. Bozkurt A, Oktar BK, Kurtel H, Alican I, Coskun T, Yegen BC. Capsaicin-sensitive vagal fibres and 5-HT3-, gastrin releasing peptide- and cholecystokinin A-receptors are involved in distension-induced inhibition of gastric emptying in the rat. Regul Pept. 1999;83:81–86. doi: 10.1016/s0167-0115(99)00050-6. [DOI] [PubMed] [Google Scholar]
  30. Broberger C, Johansen J, Johansson C, Schalling M, Hokfelt T. The neuropeptide Y/agouti gene-related protein (AGRP) brain circuitry in normal, anorectic, and monosodium glutamate-treated mice. Proc Natl Acad Sci U S A. 1998;95:15043–15048. doi: 10.1073/pnas.95.25.15043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Burdyga G, Spiller D, Morris R, Lal S, Thompson DG, Saeed S, Dimaline R, Varro A, Dockray GJ. Expression of the leptin receptor in rat and human nodose ganglion neurones. Neuroscience. 2002;109:339–347. doi: 10.1016/s0306-4522(01)00474-2. [DOI] [PubMed] [Google Scholar]
  32. Burton-Freeman B, Gietzen DW, Schneeman BO. Cholecystokinin and serotonin receptors in the regulation of fat-induced satiety in rats. Am J Physiol. 1999;276:R429–434. doi: 10.1152/ajpregu.1999.276.2.R429. [DOI] [PubMed] [Google Scholar]
  33. Buyse M, Ovesjo ML, Goiot H, Guilmeau S, Peranzi G, Moizo L, Walker F, Lewin MJ, Meister B, Bado A. Expression and regulation of leptin receptor proteins in afferent and efferent neurons of the vagus nerve. Eur J Neurosci. 2001;14:64–72. doi: 10.1046/j.0953-816x.2001.01628.x. [DOI] [PubMed] [Google Scholar]
  34. Cao WH, Madden CJ, Morrison SF. Inhibition of brown adipose tissue thermogenesis by neurons in the ventrolateral medulla and in the nucleus tractus solitarius. Am J Physiol Regul Integr Comp Physiol. 2010;299:R277–290. doi: 10.1152/ajpregu.00039.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Carr KD, Simon EJ. Potentiation of reward by hunger is opioid mediated. Brain Res. 1984;297:369–373. doi: 10.1016/0006-8993(84)90578-x. [DOI] [PubMed] [Google Scholar]
  36. Castonguay TW, Upton DE, Leung PM, Stern JS. Meal patterns in the genetically obese Zucker rat: a reexamination. Physiol Behav. 1982;28:911–916. doi: 10.1016/0031-9384(82)90213-x. [DOI] [PubMed] [Google Scholar]
  37. Chaudhri O, Small C, Bloom S. Gastrointestinal hormones regulating appetite. Philos Trans R Soc Lond B Biol Sci. 2006;361:1187–1209. doi: 10.1098/rstb.2006.1856. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Chen Z, Travers JB. Inactivation of amino acid receptors in medullary reticular formation modulates and suppresses ingestion and rejection responses in the awake rat. Am J Physiol Regul Integr Comp Physiol. 2003;285:R68–83. doi: 10.1152/ajpregu.00054.2003. [DOI] [PubMed] [Google Scholar]
  39. Chen Z, Travers SP, Travers JB. Muscimol infusions in the brain stem reticular formation reversibly block ingestion in the awake rat. Am J Physiol Regul Integr Comp Physiol. 2001;280:R1085–1094. doi: 10.1152/ajpregu.2001.280.4.R1085. [DOI] [PubMed] [Google Scholar]
  40. Ciriello J, McMurray JC, Babic T, de Oliveira CV. Collateral axonal projections from hypothalamic hypocretin neurons to cardiovascular sites in nucleus ambiguus and nucleus tractus solitarius. Brain Res. 2003;991:133–141. doi: 10.1016/j.brainres.2003.08.016. [DOI] [PubMed] [Google Scholar]
  41. Coons EE, Levak M, Miller NE. Lateral hypothalamus: learning of food-seeking response motivated by electrical stimulation. Science. 1965;150:1320–1321. doi: 10.1126/science.150.3701.1320. [DOI] [PubMed] [Google Scholar]
  42. Craig W. Appetites and aversions as constituents of instincts. Biology Bulletin. 1918;34:91–107. [Google Scholar]
  43. Cui RJ, Li X, Appleyard SM. Ghrelin inhibits visceral afferent activation of catecholamine neurons in the solitary tract nucleus. J Neurosci. 2011;31:3484–3492. doi: 10.1523/JNEUROSCI.3187-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Cunningham ET, Jr., Sawchenko PE. Anatomical specificity of noradrenergic inputs to the paraventricular and supraoptic nuclei of the rat hypothalamus. J Comp Neurol. 1988;274:60–76. doi: 10.1002/cne.902740107. [DOI] [PubMed] [Google Scholar]
  45. Dallaporta M, Himmi T, Perrin J, Orsini JC. Solitary tract nucleus sensitivity to moderate changes in glucose level. Neuroreport. 1999;10:2657–2660. doi: 10.1097/00001756-199908200-00040. [DOI] [PubMed] [Google Scholar]
  46. Daniels D, Yee DK, Faulconbridge LF, Fluharty SJ. Divergent behavioral roles of angiotensin receptor intracellular signaling cascades. Endocrinology. 2005;146:5552–5560. doi: 10.1210/en.2005-0774. [DOI] [PubMed] [Google Scholar]
  47. Darling RA, Ritter S. 2-Deoxy-D-glucose, but not mercaptoacetate, increases food intake in decerebrate rats. Am J Physiol Regul Integr Comp Physiol. 2009;297:R382–386. doi: 10.1152/ajpregu.90827.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Daughters RS, Hofbauer RD, Grossman AW, Marshall AM, Brown EM, Hartman BK, Faris PL. Ondansetron attenuates CCK induced satiety and c-fos labeling in the dorsal medulla. Peptides. 2001;22:1331–1338. doi: 10.1016/s0196-9781(01)00460-0. [DOI] [PubMed] [Google Scholar]
  49. Day HE, Campeau S, Watson SJ, Jr., Akil H. Distribution of alpha 1a-, alpha 1b- and alpha 1d-adrenergic receptor mRNA in the rat brain and spinal cord. J Chem Neuroanat. 1997;13:115–139. doi: 10.1016/s0891-0618(97)00042-2. [DOI] [PubMed] [Google Scholar]
  50. Delfs JM, Zhu Y, Druhan JP, Aston-Jones GS. Origin of noradrenergic afferents to the shell subregion of the nucleus accumbens: anterograde and retrograde tract-tracing studies in the rat. Brain Res. 1998;806:127–140. doi: 10.1016/s0006-8993(98)00672-6. [DOI] [PubMed] [Google Scholar]
  51. Dossat AM, Lilly N, Kay K, Williams DL. Glucagon-like Peptide 1 receptors in nucleus accumbens affect food intake. J Neurosci. 2011;31:14453–14457. doi: 10.1523/JNEUROSCI.3262-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Fan W, Ellacott KL, Halatchev IG, Takahashi K, Yu P, Cone RD. Cholecystokinin-mediated suppression of feeding involves the brainstem melanocortin system. Nat Neurosci. 2004;7:335–336. doi: 10.1038/nn1214. [DOI] [PubMed] [Google Scholar]
  53. Faulconbridge LF, Cummings DE, Kaplan JM, Grill HJ. Hyperphagic effects of brainstem ghrelin administration. Diabetes. 2003;52:2260–2265. doi: 10.2337/diabetes.52.9.2260. [DOI] [PubMed] [Google Scholar]
  54. Feldberg W, Pyke D, Stubbs WA. Hyperglycaemia: imitating Claude Bernard's piqure with drugs. J Auton Nerv Syst. 1985;14:213–228. doi: 10.1016/0165-1838(85)90111-0. [DOI] [PubMed] [Google Scholar]
  55. Flynn FW, Berridge KC, Grill HJ. Pre- and postabsorptive insulin secretion in chronic decerebrate rats. Am J Physiol. 1986;250:R539–548. doi: 10.1152/ajpregu.1986.250.4.R539. [DOI] [PubMed] [Google Scholar]
  56. Flynn FW, Grill HJ. Insulin elicits ingestion in decerebrate rats. Science. 1983;221:188–190. doi: 10.1126/science.6344221. [DOI] [PubMed] [Google Scholar]
  57. Gaetani S, Fu J, Cassano T, Dipasquale P, Romano A, Righetti L, Cianci S, Laconca L, Giannini E, Scaccianoce S, et al. The fat-induced satiety factor oleoylethanolamide suppresses feeding through central release of oxytocin. J Neurosci. 2010;30:8096–8101. doi: 10.1523/JNEUROSCI.0036-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Gao S, Kinzig KP, Aja S, Scott KA, Keung W, Kelly S, Strynadka K, Chohnan S, Smith WW, Tamashiro KL, et al. Leptin activates hypothalamic acetyl-CoA carboxylase to inhibit food intake. Proc Natl Acad Sci U S A. 2007;104:17358–17363. doi: 10.1073/pnas.0708385104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Geerling JC, Shin JW, Chimenti PC, Loewy AD. Paraventricular hypothalamic nucleus: axonal projections to the brainstem. J Comp Neurol. 2010;518:1460–1499. doi: 10.1002/cne.22283. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Glatzle J, Sternini C, Robin C, Zittel TT, Wong H, Reeve JR, Jr., Raybould HE. Expression of 5-HT3 receptors in the rat gastrointestinal tract. Gastroenterology. 2002;123:217–226. doi: 10.1053/gast.2002.34245. [DOI] [PubMed] [Google Scholar]
  61. Grill HJ. Leptin and the systems neuroscience of meal size control. Front Neuroendocrinol. 2010;31:61–78. doi: 10.1016/j.yfrne.2009.10.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Grill HJ, Norgren R. The taste reactivity test. II. Mimetic responses to gustatory stimuli in chronic thalamic and chronic decerebrate rats. Brain Res. 1978;143:281–297. doi: 10.1016/0006-8993(78)90569-3. [DOI] [PubMed] [Google Scholar]
  63. Grill HJ, Schwartz MW, Kaplan JM, Foxhall JS, Breininger J, Baskin DG. Evidence that the caudal brainstem is a target for the inhibitory effect of leptin on food intake. Endocrinology. 2002;143:239–246. doi: 10.1210/endo.143.1.8589. [DOI] [PubMed] [Google Scholar]
  64. Grill HJ, Smith GP. Cholecystokinin decreases sucrose intake in chronic decerebrate rats. Am J Physiol. 1988;254:R853–856. doi: 10.1152/ajpregu.1988.254.6.R853. [DOI] [PubMed] [Google Scholar]
  65. Gutierrez R, Carmena JM, Nicolelis MA, Simon SA. Orbitofrontal ensemble activity monitors licking and distinguishes among natural rewards. J Neurophysiol. 2006;95:119–133. doi: 10.1152/jn.00467.2005. [DOI] [PubMed] [Google Scholar]
  66. Harris M, Aschkenasi C, Elias CF, Chandrankunnel A, Nillni EA, Bjoorbaek C, Elmquist JK, Flier JS, Hollenberg AN. Transcriptional regulation of the thyrotropin-releasing hormone gene by leptin and melanocortin signaling. J Clin Invest. 2001;107:111–120. doi: 10.1172/JCI10741. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Havel PJ, Taborsky GJ., Jr. The contribution of the autonomic nervous system to changes of glucagon and insulin secretion during hypoglycemic stress. Endocr Rev. 1989;10:332–350. doi: 10.1210/edrv-10-3-332. [DOI] [PubMed] [Google Scholar]
  68. Hayes MR, Bradley L, Grill HJ. Endogenous hindbrain glucagon-like peptide-1 receptor activation contributes to the control of food intake by mediating gastric satiation signaling. Endocrinology. 2009a;150:2654–2659. doi: 10.1210/en.2008-1479. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Hayes MR, Chory FM, Gallagher CA, Covasa M. Serotonin type-3 receptors mediate cholecystokinin-induced satiation through gastric distension. Am J Physiol Regul Integr Comp Physiol. 2006;291:R115–123. doi: 10.1152/ajpregu.00002.2006. [DOI] [PubMed] [Google Scholar]
  70. Hayes MR, Covasa M. CCK and 5-HT act synergistically to suppress food intake through simultaneous activation of CCK-1 and 5-HT3 receptors. Peptides. 2005;26:2322–2330. doi: 10.1016/j.peptides.2005.03.045. [DOI] [PubMed] [Google Scholar]
  71. Hayes MR, Covasa M. Gastric distension enhances CCK-induced Fos-like immunoreactivity in the dorsal hindbrain by activating 5-HT3 receptors. Brain Res. 2006;1088:120–130. doi: 10.1016/j.brainres.2006.03.018. [DOI] [PubMed] [Google Scholar]
  72. Hayes MR, De Jonghe BC, Kanoski SE. Role of the glucagon-like-peptide-1 receptor in the control of energy balance. Physiol Behav. 2010a;100:503–510. doi: 10.1016/j.physbeh.2010.02.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Hayes MR, Leichner TM, Zhao S, Lee GS, Chowansky A, Zimmer D, De Jonghe BC, Kanoski SE, Grill HJ, Bence KK. Intracellular signals mediating the food intake-suppressive effects of hindbrain glucagon-like peptide-1 receptor activation. Cell Metab. 2011;13:320–330. doi: 10.1016/j.cmet.2011.02.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Hayes MR, Moore RL, Shah SM, Covasa M. 5-HT3 receptors participate in CCK-induced suppression of food intake by delaying gastric emptying. Am J Physiol Regul Integr Comp Physiol. 2004;287:R817–823. doi: 10.1152/ajpregu.00295.2004. [DOI] [PubMed] [Google Scholar]
  75. Hayes MR, Skibicka KP, Bence KK, Grill HJ. Dorsal hindbrain 5'-adenosine monophosphate-activated protein kinase as an intracellular mediator of energy balance. Endocrinology. 2009b;150:2175–2182. doi: 10.1210/en.2008-1319. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Hayes MR, Skibicka KP, Grill HJ. Caudal brainstem processing is sufficient for behavioral, sympathetic, and parasympathetic responses driven by peripheral and hindbrain glucagon-like-peptide-1 receptor stimulation. Endocrinology. 2008;149:4059–4068. doi: 10.1210/en.2007-1743. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Hayes MR, Skibicka KP, Leichner TM, Guarnieri DJ, DiLeone RJ, Bence KK, Grill HJ. Endogenous leptin signaling in the caudal nucleus tractus solitarius and area postrema is required for energy balance regulation. Cell Metab11. 2010b:77–83. doi: 10.1016/j.cmet.2009.10.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Herman JP, Cullinan WE. Neurocircuitry of stress: central control of the hypothalamo-pituitary-adrenocortical axis. Trends Neurosci. 1997;20:78–84. doi: 10.1016/s0166-2236(96)10069-2. [DOI] [PubMed] [Google Scholar]
  79. Hermann GE, Van Meter MJ, Rood JC, Rogers RC. Proteinase-activated receptors in the nucleus of the solitary tract: evidence for glial-neural interactions in autonomic control of the stomach. J Neurosci. 2009;29:9292–9300. doi: 10.1523/JNEUROSCI.6063-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Hinney A, Schmidt A, Nottebom K, Heibult O, Becker I, Ziegler A, Gerber G, Sina M, Gorg T, Mayer H, et al. Several mutations in the melanocortin-4 receptor gene including a nonsense and a frameshift mutation associated with dominantly inherited obesity in humans. J Clin Endocrinol Metab. 1999;84:1483–1486. doi: 10.1210/jcem.84.4.5728. [DOI] [PubMed] [Google Scholar]
  81. Hisadome K, Reimann F, Gribble FM, Trapp S. CCK Stimulation of GLP-1 Neurons Involves {alpha}1-Adrenoceptor-Mediated Increase in Glutamatergic Synaptic Inputs. Diabetes. 2011 doi: 10.2337/db11-0489. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Ho A, Chin A. Circadian feeding and drinking patterns of genetically obese mice fed solid chow diet. Physiol Behav. 1988;43:651–656. doi: 10.1016/0031-9384(88)90221-1. [DOI] [PubMed] [Google Scholar]
  83. Holst JJ. The physiology of glucagon-like peptide 1. Physiol Rev. 2007;87:1409–1439. doi: 10.1152/physrev.00034.2006. [DOI] [PubMed] [Google Scholar]
  84. Hudson B, Ritter S. Hindbrain catecholamine neurons mediate consummatory responses to glucoprivation. Physiol Behav. 2004;82:241–250. doi: 10.1016/j.physbeh.2004.03.032. [DOI] [PubMed] [Google Scholar]
  85. Huo L, Maeng L, Bjorbaek C, Grill HJ. Leptin and the control of food intake: neurons in the nucleus of the solitary tract are activated by both gastric distension and leptin. Endocrinology. 2007;148:2189–2197. doi: 10.1210/en.2006-1572. [DOI] [PubMed] [Google Scholar]
  86. Huszar D, Lynch CA, Fairchild-Huntress V, Dunmore JH, Fang Q, Berkemeier LR, Gu W, Kesterson RA, Boston BA, Cone RD, et al. Targeted disruption of the melanocortin-4 receptor results in obesity in mice. Cell. 1997;88:131–141. doi: 10.1016/s0092-8674(00)81865-6. [DOI] [PubMed] [Google Scholar]
  87. Jiang C, Fogel R, Zhang X. Lateral hypothalamus modulates gut-sensitive neurons in the dorsal vagal complex. Brain Res. 2003;980:31–47. doi: 10.1016/s0006-8993(03)02844-0. [DOI] [PubMed] [Google Scholar]
  88. Jordan SD, Konner AC, Bruning JC. Sensing the fuels: glucose and lipid signaling in the CNS controlling energy homeostasis. Cell Mol Life Sci. 2010;67:3255–3273. doi: 10.1007/s00018-010-0414-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Kanoski SE, Fortin SM, Arnold M, Grill HJ, Hayes MR. Peripheral and central GLP-1 receptor populations mediate the anorectic effects of peripherally administered GLP-1 receptor agonists, liraglutide and exendin-4. Endocrinology. 2011a;152:3103–3112. doi: 10.1210/en.2011-0174. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Kanoski SE, Hayes MR, Greenwald HS, Fortin SM, Gianessi CA, Gilbert JR, Grill HJ. Hippocampal Leptin Signaling Reduces Food Intake and Modulates Food-Related Memory Processing. Neuropsychopharmacology. 2011b doi: 10.1038/npp.2011.70. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Kanoski SE, Zhao S, Guarnieri DJ, DiLeone RJ, Yan J, De Jonghe BC, Bence KK, Hayes MR, Grill HJ. Leptin receptor signaling in the medial nucleus tractus solitarius reduces meal size and potentiates intestinal satiation signals. Am J Physiol: Endocrinol Metab. 2012 doi: 10.1152/ajpendo.00205.2012. In Press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Kaplan JM, Grill HJ. Swallowing during ongoing fluid ingestion in the rat. Brain Res. 1989;499:63–80. doi: 10.1016/0006-8993(89)91135-9. [DOI] [PubMed] [Google Scholar]
  93. Kenny PJ. Common cellular and molecular mechanisms in obesity and drug addiction. Nat Rev Neurosci. 2011;12:638–651. doi: 10.1038/nrn3105. [DOI] [PubMed] [Google Scholar]
  94. Khan AM, Ponzio TA, Sanchez-Watts G, Stanley BG, Hatton GI, Watts AG. Catecholaminergic control of mitogen-activated protein kinase signaling in paraventricular neuroendocrine neurons in vivo and in vitro: a proposed role during glycemic challenges. J Neurosci. 2007;27:7344–7360. doi: 10.1523/JNEUROSCI.0873-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Kishi T, Aschkenasi CJ, Lee CE, Mountjoy KG, Saper CB, Elmquist JK. Expression of melanocortin 4 receptor mRNA in the central nervous system of the rat. J Comp Neurol. 2003;457:213–235. doi: 10.1002/cne.10454. [DOI] [PubMed] [Google Scholar]
  96. Lam TK, Pocai A, Gutierrez-Juarez R, Obici S, Bryan J, Aguilar-Bryan L, Schwartz GJ, Rossetti L. Hypothalamic sensing of circulating fatty acids is required for glucose homeostasis. Nat Med. 2005;11:320–327. doi: 10.1038/nm1201. [DOI] [PubMed] [Google Scholar]
  97. Levin BE. Metabolic sensing neurons and the control of energy homeostasis. Physiol Behav. 2006;89:486–489. doi: 10.1016/j.physbeh.2006.07.003. [DOI] [PubMed] [Google Scholar]
  98. Liu H, Kishi T, Roseberry AG, Cai X, Lee CE, Montez JM, Friedman JM, Elmquist JK. Transgenic mice expressing green fluorescent protein under the control of the melanocortin-4 receptor promoter. J Neurosci. 2003;23:7143–7154. doi: 10.1523/JNEUROSCI.23-18-07143.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Maejima Y, Sedbazar U, Suyama S, Kohno D, Onaka T, Takano E, Yoshida N, Koike M, Uchiyama Y, Fujiwara K, et al. Nesfatin-1-regulated oxytocinergic signaling in the paraventricular nucleus causes anorexia through a leptin-independent melanocortin pathway. Cell Metab. 2009;10:355–365. doi: 10.1016/j.cmet.2009.09.002. [DOI] [PubMed] [Google Scholar]
  100. Malik S, McGlone F, Bedrossian D, Dagher A. Ghrelin modulates brain activity in areas that control appetitive behavior. Cell Metab. 2008;7:400–409. doi: 10.1016/j.cmet.2008.03.007. [DOI] [PubMed] [Google Scholar]
  101. Marty N, Dallaporta M, Foretz M, Emery M, Tarussio D, Bady I, Binnert C, Beermann F, Thorens B. Regulation of glucagon secretion by glucose transporter type 2 (glut2) and astrocyte-dependent glucose sensors. J Clin Invest. 2005;115:3545–3553. doi: 10.1172/JCI26309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Marty N, Dallaporta M, Thorens B. Brain glucose sensing, counterregulation, and energy homeostasis. Physiology (Bethesda) 2007;22:241–251. doi: 10.1152/physiol.00010.2007. [DOI] [PubMed] [Google Scholar]
  103. Mathis C, Moran TH, Schwartz GJ. Load-sensitive rat gastric vagal afferents encode volume but not gastric nutrients. Am J Physiol. 1998;274:R280–286. doi: 10.1152/ajpregu.1998.274.2.R280. [DOI] [PubMed] [Google Scholar]
  104. Mayer J. Glucostatic mechanism of regulation of food intake. N Engl J Med. 1953;249:13–16. doi: 10.1056/NEJM195307022490104. [DOI] [PubMed] [Google Scholar]
  105. Mazda T, Yamamoto H, Fujimura M, Fujimiya M. Gastric distension-induced release of 5-HT stimulates c-fos expression in specific brain nuclei via 5-HT3 receptors in conscious rats. Am J Physiol Gastrointest Liver Physiol. 2004;287:G228–235. doi: 10.1152/ajpgi.00373.2003. [DOI] [PubMed] [Google Scholar]
  106. McDougal DH, Hermann GE, Rogers RC. Vagal afferent stimulation activates astrocytes in the nucleus of the solitary tract via AMPA receptors: evidence of an atypical neural-glial interaction in the brainstem. J Neurosci. 2011;31:14037–14045. doi: 10.1523/JNEUROSCI.2855-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Merchenthaler I, Lane M, Shughrue P. Distribution of pre-proglucagon and glucagon-like peptide-1 receptor messenger RNAs in the rat central nervous system. J Comp Neurol. 1999;403:261–280. doi: 10.1002/(sici)1096-9861(19990111)403:2<261::aid-cne8>3.0.co;2-5. [DOI] [PubMed] [Google Scholar]
  108. Minokoshi Y, Alquier T, Furukawa N, Kim YB, Lee A, Xue B, Mu J, Foufelle F, Ferre P, Birnbaum MJ, et al. AMP-kinase regulates food intake by responding to hormonal and nutrient signals in the hypothalamus. Nature. 2004;428:569–574. doi: 10.1038/nature02440. [DOI] [PubMed] [Google Scholar]
  109. Mizuno Y, Oomura Y. Glucose responding neurons in the nucleus tractus solitarius of the rat: in vitro study. Brain Res. 1984;307:109–116. doi: 10.1016/0006-8993(84)90466-9. [DOI] [PubMed] [Google Scholar]
  110. Moran TH. Gut peptide signaling in the controls of food intake. Obesity (Silver Spring) 2006;14(Suppl 5):250S–253S. doi: 10.1038/oby.2006.318. [DOI] [PubMed] [Google Scholar]
  111. Moran TH, Ladenheim EE. Adiposity signaling and meal size control. Physiol Behav. 2011;103:21–24. doi: 10.1016/j.physbeh.2010.11.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Moran TH, Ladenheim EE, Schwartz GJ. Within-meal gut feedback signaling. Int J Obes Relat Metab Disord 25 Suppl. 2001;5:S39–41. doi: 10.1038/sj.ijo.0801910. [DOI] [PubMed] [Google Scholar]
  113. Moran TH, McHugh PR. Cholecystokinin suppresses food intake by inhibiting gastric emptying. Am J Physiol. 1982;242:R491–497. doi: 10.1152/ajpregu.1982.242.5.R491. [DOI] [PubMed] [Google Scholar]
  114. Morrison SF. 2010 Carl Ludwig Distinguished Lectureship of the APS Neural Control and Autonomic Regulation Section: Central neural pathways for thermoregulatory cold defense. J Appl Physiol. 2011;110:1137–1149. doi: 10.1152/japplphysiol.01227.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Morton GJ, Blevins JE, Williams DL, Niswender KD, Gelling RW, Rhodes CJ, Baskin DG, Schwartz MW. Leptin action in the forebrain regulates the hindbrain response to satiety signals. J Clin Invest. 2005;115:703–710. doi: 10.1172/JCI200522081. [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Mounien L, Marty N, Tarussio D, Metref S, Genoux D, Preitner F, Foretz M, Thorens B. Glut2-dependent glucose-sensing controls thermoregulation by enhancing the leptin sensitivity of NPY and POMC neurons. FASEB J. 2010;24:1747–1758. doi: 10.1096/fj.09-144923. [DOI] [PubMed] [Google Scholar]
  117. Nautiyal KM, Dailey M, Brito N, Brito MN, Harris RB, Bartness TJ, Grill HJ. Energetic responses to cold temperatures in rats lacking forebraincaudal brain stem connections. Am J Physiol Regul Integr Comp Physiol. 2008;295:R789–798. doi: 10.1152/ajpregu.90394.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Neary NM, Small CJ, Druce MR, Park AJ, Ellis SM, Semjonous NM, Dakin CL, Filipsson K, Wang F, Kent AS, et al. Peptide YY3-36 and glucagon-like peptide-17-36 inhibit food intake additively. Endocrinology. 2005;146:5120–5127. doi: 10.1210/en.2005-0237. [DOI] [PubMed] [Google Scholar]
  119. Norgren R. Taste pathways to hypothalamus and amygdala. J Comp Neurol. 1976;166:17–30. doi: 10.1002/cne.901660103. [DOI] [PubMed] [Google Scholar]
  120. Norgren R. Projections from the nucleus of the solitary tract in the rat. Neuroscience. 1978;3:207–218. doi: 10.1016/0306-4522(78)90102-1. [DOI] [PubMed] [Google Scholar]
  121. Norgren R, Hajnal A, Mungarndee SS. Gustatory reward and the nucleus accumbens. Physiol Behav. 2006;89:531–535. doi: 10.1016/j.physbeh.2006.05.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Obici S, Feng Z, Arduini A, Conti R, Rossetti L. Inhibition of hypothalamic carnitine palmitoyltransferase-1 decreases food intake and glucose production. Nat Med. 2003;9:756–761. doi: 10.1038/nm873. [DOI] [PubMed] [Google Scholar]
  123. Obici S, Feng Z, Morgan K, Stein D, Karkanias G, Rossetti L. Central administration of oleic acid inhibits glucose production and food intake. Diabetes. 2002;51:271–275. doi: 10.2337/diabetes.51.2.271. [DOI] [PubMed] [Google Scholar]
  124. Olson BR, Drutarosky MD, Stricker EM, Verbalis JG. Brain oxytocin receptor antagonism blunts the effects of anorexigenic treatments in rats: evidence for central oxytocin inhibition of food intake. Endocrinology. 1991;129:785–791. doi: 10.1210/endo-129-2-785. [DOI] [PubMed] [Google Scholar]
  125. Ouellet V, Routhier-Labadie A, Bellemare W, Lakhal-Chaieb L, Turcotte E, Carpentier AC, Richard D. Outdoor temperature, age, sex, body mass index, and diabetic status determine the prevalence, mass, and glucose-uptake activity of 18F-FDG-detected BAT in humans. J Clin Endocrinol Metab. 2011;96:192–199. doi: 10.1210/jc.2010-0989. [DOI] [PubMed] [Google Scholar]
  126. Overduin J, Frayo RS, Grill HJ, Kaplan JM, Cummings DE. Role of the duodenum and macronutrient type in ghrelin regulation. Endocrinology. 2005;146:845–850. doi: 10.1210/en.2004-0609. [DOI] [PubMed] [Google Scholar]
  127. Patterson CM, Leshan RL, Jones JC, Myers MG., Jr. Molecular mapping of mouse brain regions innervated by leptin receptor-expressing cells. Brain Res. 2011;1378:18–28. doi: 10.1016/j.brainres.2011.01.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Peters JH, McDougall SJ, Fawley JA, Andresen MC. TRPV1 marks synaptic segregation of multiple convergent afferents at the rat medial solitary tract nucleus. PLoS One. 2011;6:e25015. doi: 10.1371/journal.pone.0025015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Peters JH, McDougall SJ, Kellett DO, Jordan D, Llewellyn-Smith IJ, Andresen MC. Oxytocin enhances cranial visceral afferent synaptic transmission to the solitary tract nucleus. J Neurosci. 2008;28:11731–11740. doi: 10.1523/JNEUROSCI.3419-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Peters JH, McKay BM, Simasko SM, Ritter RC. Leptin-induced satiation mediated by abdominal vagal afferents. Am J Physiol Regul Integr Comp Physiol. 2005a;288:R879–884. doi: 10.1152/ajpregu.00716.2004. [DOI] [PubMed] [Google Scholar]
  131. Peters JH, Ritter RC, Simasko SM. Leptin and CCK selectively activate vagal afferent neurons innervating the stomach and duodenum. Am J Physiol Regul Integr Comp Physiol. 2005b doi: 10.1152/ajpregu.00811.2005. [DOI] [PubMed] [Google Scholar]
  132. Phillips RJ, Powley TL. Gastric volume rather than nutrient content inhibits food intake. Am J Physiol. 1996;271:R766–769. doi: 10.1152/ajpregu.1996.271.3.R766. [DOI] [PubMed] [Google Scholar]
  133. Pocai A, Obici S, Schwartz GJ, Rossetti L. A brain-liver circuit regulates glucose homeostasis. Cell Metab. 2005;1:53–61. doi: 10.1016/j.cmet.2004.11.001. [DOI] [PubMed] [Google Scholar]
  134. Potes CS, Lutz TA. Brainstem mechanisms of amylin-induced anorexia. Physiol Behav. 2010;100:511–518. doi: 10.1016/j.physbeh.2010.03.001. [DOI] [PubMed] [Google Scholar]
  135. Powley TL. Vagal circuitry mediating cephalic-phase responses to food. Appetite. 2000;34:184–188. doi: 10.1006/appe.1999.0279. [DOI] [PubMed] [Google Scholar]
  136. Powley TL, Phillips RJ. Gastric satiation is volumetric, intestinal satiation is nutritive. Physiol Behav. 2004;82:69–74. doi: 10.1016/j.physbeh.2004.04.037. [DOI] [PubMed] [Google Scholar]
  137. Puskas N, Papp RS, Gallatz K, Palkovits M. Interactions between orexin-immunoreactive fibers and adrenaline or noradrenaline-expressing neurons of the lower brainstem in rats and mice. Peptides. 2010;31:1589–1597. doi: 10.1016/j.peptides.2010.04.020. [DOI] [PubMed] [Google Scholar]
  138. Reimann F, Habib AM, Tolhurst G, Parker HE, Rogers GJ, Gribble FM. Glucose sensing in L cells: a primary cell study. Cell Metab. 2008;8:532–539. doi: 10.1016/j.cmet.2008.11.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  139. Ricardo JA, Koh ET. Anatomical evidence of direct projections from the nucleus of the solitary tract to the hypothalamus, amygdala, and other forebrain structures in the rat. Brain Res. 1978;153:1–26. doi: 10.1016/0006-8993(78)91125-3. [DOI] [PubMed] [Google Scholar]
  140. Rinaman L. Oxytocinergic inputs to the nucleus of the solitary tract and dorsal motor nucleus of the vagus in neonatal rats. J Comp Neurol. 1998;399:101–109. doi: 10.1002/(sici)1096-9861(19980914)399:1<101::aid-cne8>3.0.co;2-5. [DOI] [PubMed] [Google Scholar]
  141. Rinaman L. Hindbrain noradrenergic lesions attenuate anorexia and alter central cFos expression in rats after gastric viscerosensory stimulation. J Neurosci. 2003;23:10084–10092. doi: 10.1523/JNEUROSCI.23-31-10084.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Rinaman L. Ascending projections from the caudal visceral nucleus of the solitary tract to brain regions involved in food intake and energy expenditure. Brain Res. 2010;1350:18–34. doi: 10.1016/j.brainres.2010.03.059. [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Rinaman L, Card JP, Schwaber JS, Miselis RR. Ultrastructural demonstration of a gastric monosynaptic vagal circuit in the nucleus of the solitary tract in rat. J Neurosci. 1989;9:1985–1996. doi: 10.1523/JNEUROSCI.09-06-01985.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Ritter RC. Gastrointestinal mechanisms of satiation for food. Physiol Behav. 2004;81:249–273. doi: 10.1016/j.physbeh.2004.02.012. [DOI] [PubMed] [Google Scholar]
  145. Ritter S, Bugarith K, Dinh TT. Immunotoxic destruction of distinct catecholamine subgroups produces selective impairment of glucoregulatory responses and neuronal activation. J Comp Neurol. 2001;432:197–216. doi: 10.1002/cne.1097. [DOI] [PubMed] [Google Scholar]
  146. Ritter S, Dinh TT, Li AJ. Hindbrain catecholamine neurons control multiple glucoregulatory responses. Physiol Behav. 2006;89:490–500. doi: 10.1016/j.physbeh.2006.05.036. [DOI] [PubMed] [Google Scholar]
  147. Ritter S, Dinh TT, Zhang Y. Localization of hindbrain glucoreceptive sites controlling food intake and blood glucose. Brain Res. 2000;856:37–47. doi: 10.1016/s0006-8993(99)02327-6. [DOI] [PubMed] [Google Scholar]
  148. Ritter S, Li AJ, Wang Q, Dinh TT. Minireview: The Value of Looking Backward: The Essential Role of the Hindbrain in Counterregulatory Responses to Glucose Deficit. Endocrinology. 2011 doi: 10.1210/en.2010-1458. [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Ritter S, Llewellyn-Smith I, Dinh TT. Subgroups of hindbrain catecholamine neurons are selectively activated by 2-deoxy-D-glucose induced metabolic challenge. Brain Res. 1998;805:41–54. doi: 10.1016/s0006-8993(98)00655-6. [DOI] [PubMed] [Google Scholar]
  150. Ritter S, Watts AG, Dinh TT, Sanchez-Watts G, Pedrow C. Immunotoxin lesion of hypothalamically projecting norepinephrine and epinephrine neurons differentially affects circadian and stressor-stimulated corticosterone secretion. Endocrinology. 2003;144:1357–1367. doi: 10.1210/en.2002-221076. [DOI] [PubMed] [Google Scholar]
  151. Rogers RC, Kita H, Butcher LL, Novin D. Afferent projections to the dorsal motor nucleus of the vagus. Brain Res Bull. 1980;5:365–373. doi: 10.1016/s0361-9230(80)80006-2. [DOI] [PubMed] [Google Scholar]
  152. Rosenbaum M, Sy M, Pavlovich K, Leibel RL, Hirsch J. Leptin reverses weight loss-induced changes in regional neural activity responses to visual food stimuli. J Clin Invest. 2008;118:2583–2591. doi: 10.1172/JCI35055. [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Savastano DM, Carelle M, Covasa M. Serotonin-Type 3 Receptors Mediate Intestinal Polycose- and Glucose- Induced Suppression of Intake. Am J Physiol Regul Integr Comp Physiol. 2005;288:R1499–1508. doi: 10.1152/ajpregu.00745.2004. [DOI] [PubMed] [Google Scholar]
  154. Sawchenko PE. Toward a new neurobiology of energy balance, appetite, and obesity: the anatomists weigh in. J Comp Neurol. 1998;402:435–441. [PubMed] [Google Scholar]
  155. Sawchenko PE, Swanson LW. Immunohistochemical identification of neurons in the paraventricular nucleus of the hypothalamus that project to the medulla or to the spinal cord in the rat. J Comp Neurol. 1982;205:260–272. doi: 10.1002/cne.902050306. [DOI] [PubMed] [Google Scholar]
  156. Schwartz GJ. Brainstem integrative function in the central nervous system control of food intake. Forum Nutr. 2010;63:141–151. doi: 10.1159/000264402. [DOI] [PubMed] [Google Scholar]
  157. Schwartz GJ, McHugh PR, Moran TH. Gastric loads and cholecystokinin synergistically stimulate rat gastric vagal afferents. Am J Physiol. 1993;265:R872–876. doi: 10.1152/ajpregu.1993.265.4.R872. [DOI] [PubMed] [Google Scholar]
  158. Schwartz GJ, Moran TH. Sub-diaphragmatic vagal afferent integration of meal-related gastrointestinal signals. Neurosci Biobehav Rev. 1996;20:47–56. doi: 10.1016/0149-7634(95)00039-h. [DOI] [PubMed] [Google Scholar]
  159. Schwartz GJ, Netterville LA, McHugh PR, Moran TH. Gastric loads potentiate inhibition of food intake produced by a cholecystokinin analogue. Am J Physiol. 1991;261:R1141–1146. doi: 10.1152/ajpregu.1991.261.5.R1141. [DOI] [PubMed] [Google Scholar]
  160. Sclafani A, Ackroff K. The relationship between food reward and satiation revisited. Physiol Behav. 2004;82:89–95. doi: 10.1016/j.physbeh.2004.04.045. [DOI] [PubMed] [Google Scholar]
  161. Scott MM, Lachey JL, Sternson SM, Lee CE, Elias CF, Friedman JM, Elmquist JK. Leptin targets in the mouse brain. J Comp Neurol. 2009;514:518–532. doi: 10.1002/cne.22025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  162. Seeley RJ, Grill HJ, Kaplan JM. Neurological dissociation of gastrointestinal and metabolic contributions to meal size control. Behav Neurosci. 1994;108:347–352. [PubMed] [Google Scholar]
  163. Seeley RJ, Kaplan JM, Grill HJ. Effect of occluding the pylorus on intraoral intake: a test of the gastric hypothesis of meal termination. Physiol Behav. 1995;58:245–249. doi: 10.1016/0031-9384(95)00054-m. [DOI] [PubMed] [Google Scholar]
  164. Siegel EG, Trimble ER, Renold AE, Berthoud HR. Importance of preabsorptive insulin release on oral glucose tolerance: studies in pancreatic islet transplanted rats. Gut. 1980;21:1002–1009. doi: 10.1136/gut.21.11.1002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  165. Skibicka KP, Grill HJ. Energetic responses are triggered by caudal brainstem melanocortin receptor stimulation and mediated by local sympathetic effector circuits. Endocrinology. 2008;149:3605–3616. doi: 10.1210/en.2007-1754. [DOI] [PMC free article] [PubMed] [Google Scholar]
  166. Skibicka KP, Grill HJ. Hindbrain leptin stimulation induces anorexia and hyperthermia mediated by hindbrain melanocortin receptors. Endocrinology. 2009a;150:1705–1711. doi: 10.1210/en.2008-1316. [DOI] [PMC free article] [PubMed] [Google Scholar]
  167. Skibicka KP, Grill HJ. Hypothalamic and hindbrain nuclei contribute to the anorexic, thermogenic and cardiovascular action of melanocortins. Endocrinology. 2009b doi: 10.1210/en.2009-0804. [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Smith GP. The direct and indirect controls of meal size. Neurosci Biobehav Rev. 1996;20:41–46. doi: 10.1016/0149-7634(95)00038-g. [DOI] [PubMed] [Google Scholar]
  169. Smith GP, Jerome C, Norgren R. Afferent axons in abdominal vagus mediate satiety effect of cholecystokinin in rats. Am J Physiol. 1985;249:R638–641. doi: 10.1152/ajpregu.1985.249.5.R638. [DOI] [PubMed] [Google Scholar]
  170. Spanswick D, Smith MA, Groppi VE, Logan SD, Ashford ML. Leptin inhibits hypothalamic neurons by activation of ATP-sensitive potassium channels. Nature. 1997;390:521–525. doi: 10.1038/37379. [DOI] [PubMed] [Google Scholar]
  171. Steiner JE. The gustofacial response: observation on normal and anencephalic newborn infants. Symp Oral Sens Percept. 1973:254–278. [PubMed] [Google Scholar]
  172. Sumal KK, Blessing WW, Joh TH, Reis DJ, Pickel VM. Synaptic interaction of vagal afferents and catecholaminergic neurons in the rat nucleus tractus solitarius. Brain Res. 1983;277:31–40. doi: 10.1016/0006-8993(83)90904-6. [DOI] [PubMed] [Google Scholar]
  173. Sutton GM, Duos B, Patterson LM, Berthoud HR. Melanocortinergic modulation of cholecystokinin-induced suppression of feeding through extracellular signal-regulated kinase signaling in rat solitary nucleus. Endocrinology. 2005;146:3739–3747. doi: 10.1210/en.2005-0562. [DOI] [PubMed] [Google Scholar]
  174. Sutton GM, Patterson LM, Berthoud HR. Extracellular signal-regulated kinase 1/2 signaling pathway in solitary nucleus mediates cholecystokinin-induced suppression of food intake in rats. J Neurosci. 2004;24:10240–10247. doi: 10.1523/JNEUROSCI.2764-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  175. Teff KL, Engelman K. Oral sensory stimulation improves glucose tolerance in humans: effects on insulin, C-peptide, and glucagon. Am J Physiol. 1996;270:R1371–1379. doi: 10.1152/ajpregu.1996.270.6.R1371. [DOI] [PubMed] [Google Scholar]
  176. Thorens B. Central control of glucose homeostasis: the brain--endocrine pancreas axis. Diabetes Metab 36 Suppl. 2010;3:S45–49. doi: 10.1016/S1262-3636(10)70466-3. [DOI] [PubMed] [Google Scholar]
  177. Travagli RA, Hermann GE, Browning KN, Rogers RC. Musings on the wanderer: what's new in our understanding of vago-vagal reflexes? III. Activity-dependent plasticity in vago-vagal reflexes controlling the stomach. Am J Physiol Gastrointest Liver Physiol. 2003;284:G180–187. doi: 10.1152/ajpgi.00413.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  178. Uehara Y, Shimizu H, Ohtani K, Sato N, Mori M. Hypothalamic corticotropin-releasing hormone is a mediator of the anorexigenic effect of leptin. Diabetes. 1998;47:890–893. doi: 10.2337/diabetes.47.6.890. [DOI] [PubMed] [Google Scholar]
  179. van de Wall EH, Duffy P, Ritter RC. CCK enhances response to gastric distension by acting on capsaicin insensitive vagal afferents. Am J Physiol Regul Integr Comp Physiol. 2005;289:R695–703. doi: 10.1152/ajpregu.00809.2004. [DOI] [PubMed] [Google Scholar]
  180. Vaughan CH, Haskell-Luevano C, Andreasen A, Rowland NE. Effects of oral preload, CCK or bombesin administration on short term food intake of melanocortin 4-receptor knockout (MC4RKO) mice. Peptides. 2006;27:3226–3233. doi: 10.1016/j.peptides.2006.08.002. [DOI] [PubMed] [Google Scholar]
  181. Venugopal S, Boulant JA, Chen Z, Travers JB. Intrinsic membrane properties of pre-oromotor neurons in the intermediate zone of the medullary reticular formation. Neuroscience. 2010;168:31–47. doi: 10.1016/j.neuroscience.2010.03.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  182. Vrang N, Phifer CB, Corkern MM, Berthoud HR. Gastric distension induces c-Fos in medullary GLP-1/2-containing neurons. Am J Physiol Regul Integr Comp Physiol. 2003;285:R470–478. doi: 10.1152/ajpregu.00732.2002. [DOI] [PubMed] [Google Scholar]
  183. Wang FB, Powley TL. Topographic inventories of vagal afferents in gastrointestinal muscle. J Comp Neurol. 2000;421:302–324. [PubMed] [Google Scholar]
  184. Williams DL, Kaplan JM, Grill HJ. The role of the dorsal vagal complex and the vagus nerve in feeding effects of melanocortin-3/4 receptor stimulation. Endocrinology. 2000;141:1332–1337. doi: 10.1210/endo.141.4.7410. [DOI] [PubMed] [Google Scholar]
  185. Yettefti K, Orsini JC, Perrin J. Characteristics of glycemia-sensitive neurons in the nucleus tractus solitarii: possible involvement in nutritional regulation. Physiol Behav. 1997;61:93–100. doi: 10.1016/s0031-9384(96)00358-7. [DOI] [PubMed] [Google Scholar]
  186. Yoshida-Yoneda E, TJ OL, Wei JY, Vigna SR, Tache Y. Peripheral bombesin induces gastric vagal afferent activation in rats. Am J Physiol. 1996;271:R1584–1593. doi: 10.1152/ajpregu.1996.271.6.R1584. [DOI] [PubMed] [Google Scholar]
  187. Zhang Y, Kerman IA, Laque A, Nguyen P, Faouzi M, Louis GW, Jones JC, Rhodes C, Munzberg H. Leptin-receptor-expressing neurons in the dorsomedial hypothalamus and median preoptic area regulate sympathetic brown adipose tissue circuits. J Neurosci. 2011;31:1873–1884. doi: 10.1523/JNEUROSCI.3223-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  188. Zhao S, Kanoski SE, Yan J, Grill HJ, Hayes MR. Hindbrain leptin and glucagon-like-peptide-1 receptor signaling interact to suppress food intake in an additive manner. Int J Obes (Lond) 2012 doi: 10.1038/ijo.2011.265. [DOI] [PMC free article] [PubMed] [Google Scholar]
  189. Zheng H, Patterson LM, Berthoud HR. Orexin-A projections to the caudal medulla and orexin-induced c-Fos expression, food intake, and autonomic function. J Comp Neurol. 2005;485:127–142. doi: 10.1002/cne.20515. [DOI] [PubMed] [Google Scholar]
  190. Zheng H, Patterson LM, Rhodes CJ, Louis GW, Skibicka KP, Grill HJ, Myers MG, Jr., Berthoud HR. A potential role for hypothalamomedullary POMC projections in leptin-induced suppression of food intake. Am J Physiol Regul Integr Comp Physiol. 2010;298:R720–728. doi: 10.1152/ajpregu.00619.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  191. Zigman JM, Jones JE, Lee CE, Saper CB, Elmquist JK. Expression of ghrelin receptor mRNA in the rat and the mouse brain. J Comp Neurol. 2006;494:528–548. doi: 10.1002/cne.20823. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES