Next Article in Journal
Soluble Fraction from Lysate of a High Concentration Multi-Strain Probiotic Formulation Inhibits TGF-β1-Induced Intestinal Fibrosis on CCD-18Co Cells
Next Article in Special Issue
Gastrointestinal Contributions to the Postprandial Experience
Previous Article in Journal
The Western Dietary Pattern Combined with Vancomycin-Mediated Changes to the Gut Microbiome Exacerbates Colitis Severity and Colon Tumorigenesis
Previous Article in Special Issue
Gastrointestinal Vagal Afferents and Food Intake: Relevance of Circadian Rhythms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nutrient-Induced Cellular Mechanisms of Gut Hormone Secretion

Wellcome Trust-MRC Institute of Metabolic Science Metabolic Research Laboratories, Cambridge University, Addenbrookes Hospital, Cambridge CB2 0QQ, UK
*
Authors to whom correspondence should be addressed.
Nutrients 2021, 13(3), 883; https://doi.org/10.3390/nu13030883
Submission received: 18 January 2021 / Revised: 27 February 2021 / Accepted: 5 March 2021 / Published: 9 March 2021
(This article belongs to the Special Issue Appetite and Satiety Control-Gut Mechanisms)

Abstract

:
The gastrointestinal tract can assess the nutrient composition of ingested food. The nutrient-sensing mechanisms in specialised epithelial cells lining the gastrointestinal tract, the enteroendocrine cells, trigger the release of gut hormones that provide important local and central feedback signals to regulate nutrient utilisation and feeding behaviour. The evidence for nutrient-stimulated secretion of two of the most studied gut hormones, glucagon-like peptide 1 (GLP-1) and glucose-dependent insulinotropic polypeptide (GIP), along with the known cellular mechanisms in enteroendocrine cells recruited by nutrients, will be the focus of this review. The mechanisms involved range from electrogenic transporters, ion channel modulation and nutrient-activated G-protein coupled receptors that converge on the release machinery controlling hormone secretion. Elucidation of these mechanisms will provide much needed insight into postprandial physiology and identify tractable dietary approaches to potentially manage nutrition and satiety by altering the secreted gut hormone profile.

Graphical Abstract

1. Introduction

The food we eat is composed of water and macronutrients including carbohydrates, fats, and proteins. These nutrients trigger physiological responses to initiate digestion, absorption, and metabolism of nutrients to allow for their biochemical utilisation in the body. Furthermore, nutrients activate neuronal and hormonal signalling to the brain to regulate food intake and appetite. The gastrointestinal tract plays a key role in mediating the physiological effects induced by ingested nutrients. It has long been long known that specialised cells lining the gut epithelium can sense changes in luminal content and respond by releasing chemicals. Bayliss and Starling [1] described the first gut hormone secretin, and demonstrated its release following delivery of acidic solutions into the small intestine. Similarly, nutrients ingested or liberated following digestion can stimulate hormone secretion from enteroendocrine cells (EECs), which are specialised gut epithelial cells that reside within the polarized absorptive epithelial layer. The anatomy of “open” type EECs, with a slender apical process that extends to the intestinal lumen and a basolateral surface facing the interstitial space and circulatory system, link luminal composition to a variety of secreted chemical signals thus making EECs prime candidates to serve as intestinal nutrient sensors. EECs may also form additional extensions that interact with local neuronal and glial cells [2,3] to further expand the range of physiological responses to detected nutrients.
Gut responses triggered by nutrients extend beyond detection of the physical presence of substances within the intestinal lumen. Although classical studies demonstrated osmotic pressure within the stomach as an important factor in determining the rate of gastric emptying into the duodenum [4], the addition of physiological or hyperosmotic solutions of sodium chloride into the intestine were not sufficient to trigger gut hormone secretion [5], supporting the notion that nutrient-stimulated hormone release involves specific mechanisms. Moreover, nutrient-stimulated release can be disrupted using pharmacological and genetic approaches targeting transporter and carrier proteins as well as luminal and epithelial enzymes involved in digestion and absorption. This chapter will review the cellular mechanisms recruited by various nutrients to stimulate hormone secretion from the gastrointestinal tract, with a focus on mechanisms within EECs that release two important gut hormones implicated in glucose homeostasis and appetite regulation: glucagon-like peptide 1 (GLP-1) and glucose-dependent insulinotropic polypeptide (formerly known as gastric inhibitory peptide, GIP).

2. Enteroendocrine Cells That Release GLP-1 and GIP

EECs are a collection of endocrine cells of the gastrointestinal tract, releasing over 30 different hormones into the bloodstream [6] that act locally, peripherally or centrally to initiate physiological responses. Although EECs are sparse and account for only 1% of the total epithelial cell number, they are scattered along the entire gastrointestinal tract from the stomach to the rectum covering a substantial area and thus together comprise the largest endocrine organ in the body. EECs arise from local intestinal stem cells and are in a continuous state of cell turnover, being replaced every 3–5 days in the small intestine [7].
Collectively, EECs secrete a range of hormones that regulate glucose homeostasis, gut motility, appetite, adiposity, and epithelial cell proliferation. Notably, GLP-1 and GIP act as incretin hormones, amplifying insulin secretion following oral glucose administration [8,9], and account for 50–70% of total postprandial insulin secretion. Exaggerated incretin hormone release following bariatric surgery contributes to the beneficial outcomes of weight loss [10] and glycaemic control [11] and mimetics of GLP-1 are effective treatments for Type 2 diabetes [12], with some agents additionally licenced to treat obesity.
EECs are traditionally categorised into distinct cell types based on their hormonal signature. For instance, EECs that release GIP or GLP-1 are traditionally classified as K- and L-cells, respectively. However, immunohistochemical studies [13,14] and transcriptomic profiling of different EEC populations [15,16,17] have revealed an unexpected degree of overlap between EECs within the proximal small intestine, including those expressing GLP-1 and GIP. Indeed, the data suggest that individual EECs can express a much broader range of gut hormones than originally believed, which may be exploited in future therapeutic strategies. Although individual gut hormones are produced by overlapping populations of EECs, they each have a distinct longitudinal distribution along the gut [18,19,20] with the highest number of GIP-producing K-cells being found in the proximal small intestine, predominantly the duodenum [21,22], and GLP-1-producing L-cells more broadly located along the gut but increasing in numbers more distally with the highest density of L-cells in the distal small intestine and colon [23] (Figure 1).
EECs in the proximal gut are well placed to respond acutely to incoming nutrient loads and are postulated to contribute more than distally located EECs to nutrient-driven satiety [24]. Most ingested nutrients are absorbed within the proximal small intestine and compared with the distal gut, the upper small intestine receives more vagal afferent innervation, which forms part of a neural circuit that mediates satiety [25,26]. The physiological roles of nutrient-sensing mechanisms in EECs of the distal gut remain elusive, although colonic EECs may respond to locally produced microbial products [27] and lipid metabolites [28] and provide signals reflective of long-term dietary history. However, the beneficial metabolic effects of increased nutrient exposure and recruitment of distal EECs following bariatric surgery suggest these mechanisms may be exploited for effective weight and blood glucose control.
This review will focus on mechanisms of GLP-1 and GIP secretion and suppression of appetite and food intake, but it is worth mentioning other co-released gut hormones from EECs that also mediate physiological effects on food intake and may act in concert with GIP and GLP-1.
Cholecystokinin (CCK) is a peptide hormone secreted by a subset of EECs, classically designated as I-cells [29]; though significant protein levels of CCK are also produced in the brain and peripheral nervous system [30,31]. CCK plays an important role facilitating digestion in the small intestine by stimulating bile release from the gallbladder and enzyme secretion from the pancreas [32], and CCK also reduces food intake when administered peripherally [33,34] or centrally [35] via a number of mechanisms [36]. CCK is produced along the entire small intestine, with the highest density of CCK hormone-expressing cells in the duodenum [19,20,37]. Several transcriptomic [16,17,37] and immunohistochemical [37,38] studies demonstrated co-expression of CCK with other gut hormones including GIP, GLP-1, secretin and neurotensin. A recent transcriptomic analysis of all EEC populations in the small intestine revealed overlap of CCK expression in the majority of EEC cell types defined by hormonal expression profile [15]. The co-expression of CCK and incretin hormones within the same individual EECs suggests similar nutrient-sensing mechanisms can trigger multiple hormone release. However, carbohydrates were a modest secretagogue of CCK release despite being a potent stimulus for GIP and GLP-1 secretion [39,40,41]. Within EECs expressing multiple gut hormones, storage vesicles containing only CCK have been reported [42]; however, possible mechanisms to selectively mobilise specific vesicular pools remain to be elucidated.
Secretin is another peptide hormone transcriptionally co-expressed with a variety of gut peptides [15], including GLP-1 and GIP [16,17,37,43], which switches on during EEC maturation [44]. Levels of secretin are elevated postprandially and play roles in gastric acid secretion, pancreatic bicarbonate release [45] and promoting satiety. Peripheral administration of secretin decreased food intake in rats [46,47] which was mediated by vagal afferent signalling [48]. Secretin producing EECs (S-cells) are found throughout the small intestine, with the highest immunoreactivity in the duodenum [19,20,37] and are also found in the colon of adult and developing mice [49].
Xenin [50] is a 25 amino acid length neurotensin-like peptide released from GIP-expressing K-cells [51], although of questionable physiological significance as it is produced from a cytoplasmic coat protein [52] with no clear evidence of how it might reach the lumen of secretory vesicles. Secretion of xenin is elevated after a meal and possibly triggered by the anticipation of food [53]. Xenin has been reported to enhance GIP-mediated insulin secretion [54] via activation of cholinergic neurons innervating β cells. Intravenous (IV) injection of synthetic xenin stimulated jejunal motility in dogs [55] and increased contraction frequency in humans [56]. Intracerebroventricular (ICV) injection of xenin reduced food intake and weight gain in mice and this effect was abolished in neurotensin receptor 1 (Ntsr1)-deficient mice [57,58]. Neurotensin (NTS) is a peptide hormone widely distributed in the central and peripheral nervous system and expressed in a subpopulation of EECs (N-cells). Both glucose and fat triggered NTS release [59,60,61] and centrally administered NTS reduces appetite [57,62,63], with peripheral satiety effects mediated primarily by Ntsr1 receptors located on vagal afferent neurons [58,64]. The greatest expression of NTS protein is found in the ileum [19,37,65] and co-localises with a number of gut hormones including the incretin hormones GLP-1 and GIP [16,17,37,66]. Interestingly, NTS was localised to a population of vesicles distinct from those staining for GLP-1 in the distal ileum [66], which may result from EECs expressing Gcg and Nts at different times during development and maturation. Although the independent release of NTS from this distinct pool of vesicles may be possible, GLP-1 and NTS were found to be co-secreted across a range of different stimuli [66].
Oxyntomodulin (OXM) is a circulating gut hormone produced from the same proglucagon precursor peptide as GLP-1, and can activate both GLP-1 and glucagon receptors (GLP1R, GCGR, respectively) [67]. ICV and intraperitoneal (IP) injections of oxyntomodulin in rats inhibit food intake and promote weight loss [68,69] and in mouse models, both GCGR and GLP1R activity were shown to contribute to the weight loss phenotype [70]. In a randomised double-blind placebo controlled cross-over study, IV oxyntomodulin administration reduced energy intake and significantly reduced hunger scores [71]. Long acting peptides combining GCGR and GLP1R activity are in clinical trials for the treatment of type 2 diabetes and obesity [72].
Peptide YY (PYY) hormone is co-located with GLP-1 in L-cells, and is found at highest levels in the ileum and colon [18,19,20] where it is co-released with GLP-1 [73,74,75]. IP or IV administration of PYY3-36 suppressed food intake in rodents and humans [76,77], analogous to the effects of GLP-1. Interestingly, direct stimulation of gut hormone release from distal L-cells, which increased both GLP-1 and PYY levels, reduced food intake as a result of PYY signalling through Y2 receptors [78]. However, not all studies have been able to reproduce the anorexigenic actions of exogenously administered PYY [79,80,81]. In studies describing a PYY-induced reduction in feeding, mechanisms proposed to mediate this effect include the vagal-brainstem-hypothalamic circuit [25,82,83], inhibition of gastric acid secretion [84,85] and delayed gastrointestinal motility/small intestinal transit [86,87,88].
Another hormonal product co-expressed in GLP-1 expressing L-cells of the distal colon and rectum is insulin-like peptide 5 (INSL5). INSL5 protein is co-stored in the same vesicles and co-released with GLP-1 and PYY following stimulation [74]. Unlike GLP-1 and PYY, however, INSL5 appears to exert orexigenic actions, as peripheral injection of INSL5 increased food intake in mice, an effect lost in mice deficient in its cognate receptor, relaxin family peptide receptor 4, Rxfp4 [89]. The overall importance of INSL5 in regulating feeding behaviour is unclear as ablation of Insl5 expression in mice resulted in no obvious feeding deficit [90] and a subtle orexigenic effect of selective distal L-cell stimulation became apparent only when the opposing and overriding anorexigenic PYY effect was blocked with a Y2R-inhibitor [78]. The physiological rationale for distal EECs co-releasing hormones with opposing actions on food intake has yet to be reconciled.

3. Models to Study Nutrient-Sensing Mechanisms in the Gut

A variety of in vitro and ex vivo experimental models have been developed to study the mechanisms regulating EECs function. In combination with in vivo models and clinical studies that measure food intake following ingestion of nutrients, a comprehensive understanding of nutrient-stimulated responses in the gastrointestinal tract has been revealed.
Studies utilising intestinal cell line models of EECs provided initial insights into nutrient-sensing mechanisms in EECs. The most commonly used murine and human models for GLP-1 secretion are GLUTag [91] and NCI-H716 cells [92,93], respectively. GLUTag cells, derived from oncogenic tumours from the large bowel, respond to a range of nutrient and hormonal stimuli [94], whereas NCI-H716 cells, derived from a poorly differentiated adenocarcinoma of the human caecum, also respond to a range of nutrient stimuli [95] but possess altered regulation of proglucagon gene expression [96]. The secretin tumour cell line, STC-1 [97], derived from a mouse small intestinal neuroendocrine carcinoma, secretes a variety of small intestinal gut hormones including GIP, GLP-1, secretin and CCK. Subclones of STC-1 cells have been generated to produce lines with increased GIP expression or secretion [98,99,100]. Conflicting reports on the sensitivity of STC-1 cells to glucose emphasized the need for studies using primary intestinal cell models. Initial attempts to culture primary EECs involved elutriation to purify and enrich the K- or L-cell population from canine intestinal epithelia [101,102] or the use of fetal rat intestinal tissue [103]. Optimised culturing techniques, including enrichment of intestinal crypts or the deep folds of the intestinal epithelium have permitted studies from primary cultures of adult mouse [104] and human [105] intestinal tissue. The development of transgenic mouse models labelling K- and L-cell populations [104,106] has allowed transcriptomic and single-cell recording approaches to be applied to specific EEC populations and advanced our understanding of the signalling pathways recruited following specific nutrient exposure. Renewable cell culture technologies such as intestinal organoids [107] have been used to study EECs and have confirmed many nutrient-stimulated signalling mechanisms described in primary intestinal preparations [108] but more importantly provide a means to interrogate nutrient-sensing mechanisms in human EEC populations by labelling specific populations of EECs with fluorescent proteins under the control of hormonal promoters [109,110].
Two-dimensional cultures of these in vitro cell models permit access to EECs for membrane recordings and intracellular measurements. They also allow nutrients to bind targets on EECs that may be inaccessible from the luminal side of the intestine, by-passing nutrient absorptive and transport mechanisms. Three-dimensional organoid models maintain a polarized epithelium in culture, and as the luminal epithelial surface is directed towards the central organoid domain, exogenously applied nutrients may still by-pass transport and absorptive pathways. To investigate nutrient-sensing mechanisms under more physiological settings, ex vivo preparations such as Ussing chambers [111,112,113] and vascular perfused intestinal models have been used by several laboratories [114,115]. Properly prepared, both models retain tight junctions between epithelial cells and maintain the polarity of the epithelial layer. The location of receptors and sites of action of nutrients can be determined in these models by application of nutrients to the isolated apical or basolateral surface. Hormone secretion from L-cells has been monitored in these models either by collection of perfusate or media from the basolateral side combined with immunoassays [112], or by measurement of trans-epithelial short circuit currents [113].
Finally, in vivo models allow assessment of nutrient ingestion or infusion and effects on food intake and feeding behaviour. In rodents and humans, diets may be altered to control for macronutrient composition, but the palatability and other sensory attributes of food (e.g., odour) should be considered, as inputs from lingual sensors and olfactory centres can also impact feeding behaviour. Nutrients may also be delivered directly into the intestine by oral gavage or by insertion of a nasoenteral feeding tube into the stomach, duodenum, or jejunum. Alternatively, catheters may be surgically inserted to directly infuse nutrients into the lumen of a specified intestinal region, which however, if infused distal to the sphincter of Oddi, may not mix physiologically with pancreatic exocrine secretions and bile, thus altering the digestion of macronutrients. Studies involving human participants are relevant in translating our understanding of the cellular mechanisms underlying nutrient-sensing but can be particularly challenging. Measurements of hunger are subjective and dependent on previous eating habits. Moreover, multiple factors such as alterations in gut motility may unexpectedly modify gastrointestinal responses to nutrients. For instance, the effect of nutrients on hunger scores was attenuated in older participants [116].

4. Cellular Mechanisms of Nutrient-Induced Gut Hormone Secretion

The suppression of food intake following oral ingestion or direct infusion of nutrients into the intestine is associated with a rise in the release of GLP-1 and GIP [117,118,119]. Nutrient-induced gut hormone secretion entails recruitment of specific cellular mechanisms that are influenced by rates of digestion and absorption as well as expression of nutrient-specific transporters and receptors.

4.1. Carbohydrates

Complex carbohydrates cannot be absorbed across the intestinal wall and therefore must be broken down to monosaccharides before they are transported across cell membranes. Amylases from saliva and pancreatic secretions initiate enzymatic digestion and a combination of hydrolases expressed on enterocytes complete the breakdown of ingested carbohydrates to monosaccharides.
Mechanisms of glucose-stimulated GLP-1 and GIP secretion have been intensively investigated as glucose is a potent secretagogue and there is interest in understanding the role of both incretin hormones in the pathology and therapeutics of Type 2 diabetes (T2D). In humans, glucose can be detected in the proximal duodenum within 5 min of ingestion of liquid glucose load, correlating with the time for first rapid elevation of GLP-1 and GIP concentrations in the bloodstream [120]. The early rise in GIP levels is readily attributed to arrival of nutrients in the duodenum with its high local density of GIP-expressing K-cells. However, the rapid rise in GLP-1 occurs well before glucose can reach the distal portion of the gut, and most of the glucose ingested is absorbed in the proximal small intestine with very little passing more distally to where the majority of GLP-1 expressing L-cells reside [120]. Measurements of glucose concentrations in the distal gut of several species after a meal confirm low levels of glucose [121]; however, levels at the ileo-caecal junction can rise as high as 10 mM after a meal [122]. The most likely mechanism to account for the rapid phase of GLP-1 secretion is the activation of proximal GLP-1 expressing cells [123], rather than recruitment of neuronal or humoral signals from the proximal to distal intestine, as patients with ileal resections maintained the rapid phase of GLP-1 secretion [124].
Glucose-stimulated release of GLP-1 and GIP from EECs requires absorption of carbohydrates. Monosaccharides are transported into cells by active and facilitative glucose transporters. Active glucose transporters, including the sodium-glucose linked transporter 1 (SGLT1), are located on the apical surface of the intestinal epithelial layer [125,126,127] and Sglt1 is expressed in GIP- and GLP-1-releasing EECs [104,106,110] (Figure 2). Sodium-coupled active transporters carry glucose into cells, against its concentration gradient, along with sodium ions (Na+) down the Na+ concentration gradient established by basolateral Na/K-ATPase activity. The movement of 2 Na+ ions per glucose molecule by SGLT1 produces a small electrogenic current which is sufficient to depolarise the cell membrane potential and trigger action potential firing in GLUTag cells [128]. The sensitivity of L-cells to release GLP-1 in response to glucose closely mirrors the binding potency of SGLT1 for glucose [129] supporting the role of SGLT1 as the primary glucose-sensing mechanism for GLP-1 expressing cells. Non-metabolisable sugars such as alpha-methyl-glucopyranoside and 3-O-methylglucose [130,131] are also substrates for SGLT1 and stimulate GIP and GLP-1 secretion in vivo [115,132] and activating SGLT1 using various substrates reduced food intake [24]. Further supporting the essential role of SGLT1 in glucose-stimulated incretin hormone release, the SGLT1 inhibitor phloridzin or knock-out of Sglt1 reduced glucose-triggered GLP-1 and GIP release [115,132,133]. However, whilst glucose triggered GIP-secretion is essentially absent in Sglt1-knock-out mice, the profile of GLP-1 release after an oral glucose challenge in Sglt1-deficient mice is complex as there is significantly reduced GLP-1 release at early time points (<15 min), but elevated GLP-1 release at later times (1–2 h) [134]. The amplified delayed phase of GLP-1 release in this model is thought to be due to reduced glucose absorption in the proximal gut and delivery of more glucose to the distal intestine that seems to activate SGLT1-independent glucose-sensing mechanisms, possibly involving fermentation to short chain fatty acids. Another sodium-glucose co-transporter, SGLT3, has been described in GLUTag cells [128] and enteric neurons of the submucosal and myenteric plexus [135], but whether SGLT3 contributes to glucose-sensing in EECs is not clear. Interestingly, the human variant of SGLT3 has been reported to have lost its capacity to transport glucose and operates as a glucose-sensitive sodium channel [135]. Facilitative glucose transporters are differentially expressed along the gastrointestinal tract [125] and generally localised to the basolateral surface to facilitate glucose efflux into the bloodstream. Transient apical translocation of glucose transporters has also been implicated in the absorption of glucose into intestinal cells [136,137]. Transient GLUT2 insertion into the brush border shortly after high glucose exposure would allow rapid glucose uptake when SGLT1 capacity is saturated and maintain the Na+ concentration gradient for other cellular processes, but this hypothesis remains controversial [138,139].
Glucose-sensitive tissues, including enteroendocrine K- and L-cells, express glucokinase (GCK), usually the rate-limiting enzyme in the breakdown of glucose [14,17,104,106,140,141]. The relatively low affinity of GCK for glucose links glycolytic fluxes to physiologically relevant extracellular glucose concentrations. However, patients with inactivating mutations in GCK did not exhibit impaired GLP-1 or GIP secretion following an oral glucose challenge [142], suggesting GCK is not the primary glucose sensor for the gut. Potassium channels sensitive to ATP (KATP channels) are described in a number of glucose-responsive tissues as a mechanism to link the nutrient status of a cell to membrane electrical excitability. The generation of ATP following catabolism of glucose increases the intracellular ATP to ADP ratio, leading to the closure of KATP channels and depolarisation of the membrane potential that may trigger action potential firing or activation of voltage-gated calcium channels to facilitate Ca2+ entry into the cell [143]. KATP channels are composed of a pore forming subunit (Kir6.1/2) and a sulphonylurea receptor (SUR1/2), and both Kir6.2 and SUR1 are highly expressed in GIP- and GLP-1-expressing EECs [104,106,144]. In electrophysiological recordings from GLP-1 expressing cells, glucose depolarises the cell membrane potential, triggers an increase in electrical activity and stimulates GLP-1 secretion [104,110,145]. Closure of KATP channels by tolbutamide stimulated a similar increase in action potential firing frequency and diazoxide, a KATP channel opener, blocked the effects of high glucose on cell excitability and GLP-1 release [104,145,146]. Much less is known about the electrical activity of GIP-expressing K-cells, but tolbutamide stimulated an increase in GIP secretion in primary small intestinal cultures [106] and in STC-1 cells high glucose triggered membrane depolarisation and extracellular Ca2+ entry to stimulate CCK release [147]. However, the finding of functional KATP channels in EECs does not imply that they are involved in glucose-triggered hormone release, and evidence that KATP channels regulate glucose-stimulated incretin hormone secretion in vivo is lacking. Mice deficient in Kir6.2 expression maintained elevated GIP release after an oral glucose load [148] and sulfonylureas had no effect on peak GLP-1 and GIP responses to an oral glucose tolerance test in human participants [149,150]. This discrepancy around the impact of KATP channels on regulating gut hormone secretion could be a result of low resting KATP channel activity in vivo compared to in vitro culturing conditions.
Other carbohydrates, such as galactose, are handled by the intestine in a similar way to glucose. However, some other dietary carbohydrates, such as fructose, can recruit different mechanisms in EECs. Fructose is not as satiating as glucose in rats [24,26] and monkeys [151] but was found to have similar or greater satiating potential in humans [152,153]. Fructose can stimulate GLP-1 and GIP secretion [60,106,115,128]; though some studies found fructose did not stimulate GIP [60,132] or GLP-1 [154] release. Fructose is not a substrate for SGLT1 but is transported into cells by the apical transporter GLUT5 [155] and possibly GLUT2. GLUT5 is abundant in the gut epithelium, with higher expression in the proximal small intestine [156]. GIP- and GLP-1 expressing cells of the intestine expressed Slc2a5, the gene encoding GLUT5 [104,106]. Fructose is fully metabolised, like glucose, and can recruit metabolic pathways such as KATP channel closure to induce gut hormone secretion. In GLUTag cells, fructose triggered membrane depolarisation and action potential firing, and a decrease in conductance consistent with closure of KATP channels [128]. Furthermore, fructose stimulated GLP-1 secretion was abolished by treatment with the KATP channel opener diazoxide [60], but other studies have found that mice lacking the pore-forming subunit Kir6.2 maintained fructose-stimulated GIP and GLP-1 release [157].
Sweet taste receptors (STRs) expressed in the gut can also detect glucose and other natural or synthetic sweeteners; however, a number of conflicting reports in the field have prevented a consensus view on their glucose-sensing role in EECs. Sweet taste receptors were first described in oral lingual taste cells [158] and comprise heterodimeric GPCRs from the type 1 taste receptor family, T1R2 and T1R3 [159]. STRs in the tongue couple to α-gustducin and activate the calcium-sensitive transient receptor potential cation channel subfamily M member 5 (TrpM5) to allow Na+ entry into cells and trigger membrane depolarisation. Components of the STR signalling pathway have been identified in isolated cells of the intestinal epithelium and co-localised with GLP-1 and GIP [160,161,162,163]; however, it is unclear if all elements are expressed in the same EEC to reconstitute functional STR signalling [164]. Furthermore, expression of the genes encoding the STR subunits Tas1R2 and Tas1R3 were not readily detectable in GIP or GLP-1-expressing cells [104,106,110]. Agonists of STRs, such as sucralose, stimulated GLP-1 release from GLUTag [165] and NCI-H716 cells [162], but did not enhance GIP or GLP-1 secretion from murine primary intestinal epithelial cultures [104,106]. Artificial sweeteners were also unable to replicate glucose-stimulated GIP [163] or GLP-1 levels [166] in rats and humans, and although mice lacking α-gustducin or T1R3 exhibited reduced GIP and GLP-1 responses [161,162], inhibiting STRs with gurmarin or blocking TrpM5 channels did not inhibit luminal glucose stimulated GIP and GLP-1 secretion [167]. Further work is needed to clarify the role, if any, of STRs, and the accompanying transduction pathway in gut endocrine cells.

4.2. Proteins

Dietary proteins are broken down to small peptides by enzymes in the stomach and pancreas, then further digested by peptidases on the brush border to smaller di- and tripeptides and free amino acids. Amino acids are primarily recycled to produce proteins in the body but can also be used to generate energy when carbohydrate or lipid stores are depleted. In comparison to carbohydrates, the number of different products that arise from the breakdown of ingested proteins is vast; however, specific amino acids or small peptides can elicit specific responses from EECs. In classical studies, intraduodenal infusion of specific amino acids was more potent at stimulating GIP over CCK secretion [168]. The specificity of responses could permit targeting of peptide/amino acid-sensing pathways in incretin hormone releasing cells to manage food intake and satiety.
Several combinations of peptides or protein hydrolysates have been shown to stimulate hormone secretion from EECs through a variety of mechanisms. Peptones, consisting of a soluble mix of amino acids and peptides derived from partial protein hydrolysis, potently stimulated GLP-1 and GIP secretion from GLUTag and STC-1 cells [100,169], which involved upregulation of proglucagon gene expression but not alteration in KATP channel activity; arguing against metabolic production of ATP from dietary protein as the primary mechanism for peptide-stimulated gut hormone secretion. In NCI-H716 cells, meat hydrolysates triggered GLP-1 release by recruiting the MAPK signalling pathway but did not alter proglucagon gene expression [170], perhaps highlighting species variation in protein/peptide-sensing in the gut.
In murine primary colonic L-cells, the non-metabolisable dipeptide glycine-sarcosine (Gly-Sar) stimulated GLP-1 secretion [171], likely involving the H+/peptide co-transporter, PEPT1, as higher pH, an inhibitor of PEPT1 (4-aminomethylbenzoic acid) and knock-out of Pept1 abolished dipeptide stimulated GLP-1 release and intracellular Ca2+ responses in vitro (Figure 3). Transport by PEPT1 is electrogenic and activation of PEPT1 has been shown to depolarise the cell membrane potential and activate voltage-gated calcium channels, which can contribute to hormone secretion from EECs [172]. Further work is required, however, to ascertain the role of PEPT1 as a dipeptide sensor in EECs in vivo. Similarly, the role of another electrogenic amino-acid transporter located at the apical site of the intestinal epithelium, B(0)AT-1 (Slc6a19), needs to be further explored. B(0)AT-1 knock-out mice showed elevated GIP and GLP-1 responses upon refeeding compared to their wild type littermates, which at least for GLP-1 might reflect increased delivery of nutrients to the more distal intestine [173].
The calcium sensing receptor, CaSR, is expressed in L-cells of the small intestine and colon and also contributes to the stimulatory responses to amino acids, dipeptides and peptones in L-cells [171,174]. Agonists of CaSR triggered GLP-1 release and selective CaSR antagonists reduced peptone-triggered GLP-1 release from rodent primary intestinal cultures [171,175]. This role of CaSR in amino acid-sensing in L-cells is comparable to its involvement in peptone triggered release of CCK from STC-1 and primary CCK-releasing EECs [176] supporting a common mechanism of amino acid sensing across EEC populations. CaSR responds to a broad range of amino acids and in other intestinal epithelial cell models the receptor couples to the phosphatidylinositol pathway resulting in elevation of intracellular Ca2+ [177]. In murine primary L-cells, however, peptone triggered GLP-1 release was inhibited by nifedipine and lanthanum, suggesting roles for L-type Ca2+ channels and TRP channels, respectively, as alternative signalling pathways downstream of CaSR recruitment [174]. In the perfused rat small intestinal model, CaSR agonists were shown to access basolaterally located receptors to trigger GLP-1 release [178].
There are several other GPCRs that have been implicated in amino acid responses in EECs. The lysophosphatidic acid receptor 5 (LPAR5, also known as GPR92/93), is highly expressed in the intestinal mucosal layer particularly in the duodenum [179]. This promiscuous receptor couples to various G-proteins, including Gαq and Gα12/13 [180]. LPAR5 is proposed to mediate peptone-sensing in CCK-expressing EECs [179]; however, expression of Lpar5 transcript was not detected in colonic L-cells and GLP-1 secretion was not impaired in primary cultures from Lpar5-deficient mice [171] suggesting LPAR5 is not a major contributor to amino acid-sensing in L-cells.
GPR142 is a Gαq-coupled receptor activated by aromatic amino acids such as tryptophan and phenylalanine. In vivo, oral dosing of aromatic amino acids increased plasma GIP levels which were abolished in Gpr142 knock-out animals [181,182]. Interestingly, aromatic amino acids also evoked a rapid rise in plasma GLP-1 levels after oral gavage but this was maintained in Gpr142-deficient mice [182]. GPR142 seems also not to be required for gut hormone responses to dietary protein, as mixed protein delivered orally to mice triggered a robust increase in GIP and GLP-1 secretion that was maintained in Gpr142-deficient mice [181].
Taste receptors composed of T1R1/T1R3 subunits, which form the umami taste receptor, respond to a broad spectrum of aliphatic amino acids including L-glutamate [159,183]. This receptor is allosterically modulated by purine nucleotides, such as inosine-5-monophosphate (IMP), to potentiate amino acid responses. However, even though genes encoding the components of the umami taste receptor, Tas1R1 and Tas1R3, were not enriched in GLP-1 or GIP-expressing cells [104,106,110], umami receptor dependent GLP-2 secretion has been reported [184]. In lingual taste cells, residual sensitivity to oral glutamate and IMP was observed in double Tas1R1/Tas1R3 knock-outs [185] suggesting other receptors may be involved in mediating glutamate-triggered taste responses. Alternative candidate receptors including ionotropic and metabotropic glutamate receptors (mGluR1/4) have been identified in taste cells [186,187]; even though expression and a defensive role of mGluR4 activation in the duodenal mucosa has been demonstrated [188], the question of whether similar mechanisms are utilised in EECs remains to be studied.
GPRC6A is another receptor that responds to a broad spectrum of amino acids, particularly basic ones [189]. GPRC6A is activated by multiple ligands besides amino acids, including the hormones osteocalcin and testosterone, and is allosterically modulated by Ca2+ [190]. GPRC6A colocalised with GLP-1 expressing cells in the small intestine [191], though the majority of GPRC6A-positive cells were not immunoreactive for GLP-1. However, in another study, Gprc6a expression was low in primary L-cells [171] and of all the GPCRs associated with amino acid detection, Gprc6a was found at lowest abundance in STC-1 cells [192]. Although GPRC6A contributed to ornithine triggered GLP-1 release from GLUTag cells, ornithine did not stimulate GLP-1 secretion from primary L-cells [193].
More work is needed in the area of amino acid sensing mechanisms in the gut as several amino acids exert potent effects on gastrointestinal function and feeding behaviour, but the exact mechanisms remain to be fully elucidated. Tryptophan, for instance, can potently suppress energy intake and inhibit gastric emptying [194]. Although GLP-1 release from EECs can reduce food intake and slow gastric emptying rates, only modest changes in GLP-1 levels were reported by several groups following tryptophan administration [195,196] and studies that reported an increase in GLP-1 levels following tryptophan administration found GPR142 not to be involved [182]. GIP secretion was markedly increased after an oral tryptophan ingestion [181], but GIP is proposed to have the opposite effect and increase the gastric emptying rate [197]. Another amino acid with an incompletely characterised mechanism of action is glutamine, which triggers incretin hormone secretion from GLUTag, primary L- and K-cells and in human participants [104,106,198,199,200]. A sodium-dependent electrogenic transporter may be involved in glutamine sensing as both glutamine and asparagine triggered depolarising currents when applied to GLUTag cells [198] and transcripts for several electrogenic amino acid transporters are expressed in GLUTag cells. However, under non-electrogenic conditions where the cell membrane potential is pre-emptively depolarised to negate the effect of small currents generated by electrogenic transporters, glutamine could still enhance GLP-1 release [198]. Intracellular calcium and cAMP levels are elevated in GLP-1 expressing cells following glutamine application [201] suggesting the activation of additional pathways. The rise in intracellular calcium was dependent on extracellular Na+ and Ca2+, so recruitment of voltage-gated calcium channels following activation of a sodium-dependent electrogenic transporter is possible. CaSR may also be involved as the CaSR inhibitor Calhex 231 reduced glutamine stimulated GIP and GLP-1 release [174,175]. Studies in other endocrine cell models have demonstrated that CaSR can also couple to Gαs to stimulate adenylyl cyclase activity and cAMP production [202].

4.3. Fats

Fats are highly effective at suppressing energy intake and appetite [203,204], stimulating GIP secretion [118], and eliciting sustained GLP-1 release [118,205]. The stimulatory effects of dietary lipids on incretin hormone secretion were found to be dependent on fatty acid chain length and saturation [206,207,208,209]; however, other studies demonstrated the importance of metabolism and absorption for fat-stimulated GLP-1 and GIP secretion [210].
The processing of lipids after ingestion is more involved than carbohydrates or proteins. Dietary lipids in the form of triglycerides, cholesterol, phospholipids and fat-soluble vitamins are first emulsified by bile salts to promote efficient hydrolysis and absorption. Triglycerides are broken down to monoglycerides and fatty acids by pancreatic lipases released into the intestinal lumen. Hydrolysis of triglycerides is essential for EEC lipid-sensing as orlistat, a lipase inhibitor, attenuated postprandial GIP and GLP-1 secretion [211,212,213,214]. The breakdown products generated by lipases aggregate with bile acids to form micelles, facilitating uptake by enterocytes. Within enterocytes, triglycerides are re-synthesised from fatty acids and monoglycerides, and packaged with lipoproteins and other lipids to form chylomicrons. Chylomicrons are released from enterocytes from the basolateral surface and enter the lymphatic system through the central lacteal of the villus. Physiological concentrations of chylomicrons [215] significantly stimulated GLP-1 and GIP secretion in murine and human duodenal cultures [216]. The formation of chylomicrons seems to be required for lipid-stimulated release of incretin hormones as pluronic L-81, a surfactant that inhibits chylomicron synthesis [217], impaired lipid-stimulated release of GLP-1 and GIP [218,219]. Furthermore, murine knock-outs of genes responsible for re-esterification of absorbed free-fatty acids and monoglycerides prior to their incorporation into chylomicrons, monocaylglyceride-acyltransferase-2 (MGAT2) and diacylglyceride-acyltransferase-1 (DGAT1), reduced secretion of GIP from the upper GI tract following an oral triglyceride load [220]. However, MGAT2 and DGAT1 knock-out animals exhibited delayed GLP-1 responses to an oral triglyceride load, likely because impairment of fat absorption increased delivery of lipids more distally to regions of the intestine with a higher density of GLP-1-producing L-cells, leading to activation of other non-chylomicron-mediated mechanisms of GLP-1 release [221]. A neuronally mediated circuit has also been implicated in regulating feeding responses following lipid ingestion, as vagal deafferentation blocked the ability of oleate and corn oil infusions to reduce food intake [222] and silencing neuronal activity with tetracaine or capsaicin prevented duodenal lipid suppression of feeding in sham-fed rats [223,224].
The cellular mechanisms underlying lipid-stimulated GLP-1 and GIP secretion involve G-protein coupled receptors of the free-fatty acid (FFA) receptor family (Figure 4). The free-fatty acid receptor 1 (FFA1), known previously as the orphan receptor GPR40 [225,226], is highly expressed in K- and L-cells [104,106,110] and FFA1 colocalised with gut hormones GIP and GLP-1 [227]. FFA1 binds medium to long chain fatty acids and mediates lipid-induced incretin hormone release, as mice lacking Ffar1 expression exhibited impaired release of GIP and GLP-1 following consumption of a high fat diet [227]. FFA1 couples primarily to Gαq proteins, which recruit phosphatidylinositol signalling pathways, but with some ligands has also been suggested to couple to Gαs proteins to elevate cAMP levels [228]. Heterologous expression studies suggested that Gαi/o proteins may also contribute to the downstream signalling associated with FFA1 activation [229]. In pancreatic β cells, activation of FFA1 was associated with activation of transient receptor potential cation channel subfamily C member 3, TrpC3 [230] and in mouse GLP-1 expressing cells TrpC3 activation downstream of FFA1 generated depolarising currents to increase electrical excitability and GLP-1 secretion [108]. A similar increase in electrical excitability following FFA1 activation was observed in human GLP-1 expressing L-cells, but the mechanism did not appear to involve TrpC3 channels [110]. The other known GPCR of the free-fatty acid receptor family responsive to long-chain fatty acids is FFA4, formerly known as GPR120 [231]. FFA4 has been identified in the intestine, particularly in GIP-expressing K-cells [106,232,233] and GLP-1-expressing L-cells [104,110]. FFA4 is activated by unsaturated long-chain fatty acids such as α-linoleate, docosahexaenoic acid (DHA), palmitoleate and oleate. Activation of FFA4 by long-chain fatty acids promotes secretion of GLP-1 [231] and knock-down of Ffar4 in STC-1 cells reduced GLP-1 secretion induced by α-linolenic acid [231]. Mice deficient in Ffar4 or treated with a pharmacological inhibitor of FFA4 exhibited reduced GIP responses to an oral lard oil challenge [233]. However, the relative contributions of FFA1 and FFA4 in mediating lipid-stimulated GIP secretion remain incompletely resolved. One study found that knock-out of Ffar1 but not Ffar4 reduced GIP release following an oral olive oil challenge, although double knock-out of both receptors was more effective at reducing lipid-stimulated GIP secretion than either single knock-out [234]. By contrast, another study reported that mice deficient in Ffar1 or Ffar4 exhibited reduced GIP secretion following oral corn oil ingestion, with a greater reduction in Ffar1 knock-out animals [235]. In this study, it was noted that the simultaneous reduction in CCK secretion further impaired GIP secretion, likely downstream of reduced gallbladder contraction, as GIP levels were partially restored following exogenous CCK replacement, particularly in the Ffar4 knock-out group. Like FFA1, FFA4 couples to Gαq-dependent pathways to mediate GLP-1 secretion, but Gαi/o pathways have been implicated in other FFA4 expressing cells including gastric ghrelin-secreting cells [236] and pancreatic δ cells [237]. Recruitment of protein kinase C ζ, which is not a target of diacylglycerol downstream of FFA1/4 activation, was implicated in oleic acid-induced GLP-1 release from GLUTag cells [238], although details of the pathway remain uncertain. Recruitment of the transient receptor potential channel TrpM5 has been linked to FFA4 activation by linoleic acid to stimulate CCK release from STC-1 cells [239], mirroring the involvement of TrpC3 in L-cells downstream of FFA1 activation [108]. Although the reported recruitment of different Trp channels downstream of FFA1 and FFA4 might reflect functional differences between these receptors, the possibility of cell line or species-specific responses cannot be ruled out.
Another GPCR involved in lipid sensing in the gastrointestinal tract is GPR119, which is also enriched in K- and L-cells [104,106,110,234,240]. A number of endogenous lipid substrates bind GPR119 receptors including: oleoylethanolamide (OEA), a lipid amide synthesized in the small intestine during absorption of dietary fats [241]; 2-oleoylglycerol (2-OG) and other 2-monoacylglycerols, natural digestion products of intestinal triacylglycerol digestion [242]; N-oleoyldopamine [243] and lysophosphatidylcholine [244]. The selectivity of the above-mentioned lipid-derivatives for GPR119 still needs to be addressed as OEA was able to suppress food intake in Gpr119-deficient mice suggesting other targets of OEA [245]. It has been reported, for example, that OEA is an endogenous substrate for the nuclear receptor PPAR-α which can also influence satiety and lipid uptake [241,246,247]. Application of OEA to various in vitro L-cell models, however, triggered GPR119-dependent GLP-1 secretion [240]. Oral administration of OEA to humans increased plasma GLP-1 and GIP levels [242], and small synthetic agonists of GPR119 elevated GIP and GLP-1 levels in mice [248] and decreased food intake in rats [249]. Whilst it is possible that local production of OEA in the small intestine modulates GPR119 activity [250], it is unclear if sufficient OEA is generated to activate the receptor [251]. Nevertheless, GPR119 seems to be an important sensor of ingested fat, as there was reduced GIP secretion following an oral triglyceride challenge in Gpr119 knock-out animals, [234], and impaired oil-triggered GLP-1 release in mice lacking Gpr119 in L-cells [252]. The GLP-1 response to GPR119 agonism is more prominent in the distal gut, as it was impaired in mice lacking L-cells in the terminal ileum and large intestine [28]. Mirroring this finding, GPR119 agonists were more effective at triggering GLP-1 release in vitro from murine colonic than small intestinal cultures, and elevated cAMP in ∼70% of colonic L-cells but only 50% of small intestinal L-cells [252]. By contrast, GPR119 agonism was a relatively poor stimulus of GLP-1 secretion from human colonic cultures [105], although this might reflect species-specific differences in the receptor or in vitro preparation. In a clinical study, multiple dosing of a GPR119 agonist JNJ-38431055 in subjects with T2D increased plasma GLP-1 and GIP levels, but this did not translate into improvements in 24 h blood glucose control [253]. GPR119 couples to Gαs proteins to increase cAMP levels [234,252]. Thus, in conjunction with Gαq-signalling pathways downstream of FFA1/4, the magnitude of gut hormone secretion following fat ingestion is amplified beyond activation of a single GPCR. This feature may underlie the effectiveness of fats in stimulating robust incretin hormone responses.
As mentioned above, lipid absorption is critical for lipid-induced GIP secretion, and impairment of luminal fat digestion or epithelial absorption reduced GIP release. Receptors for free fatty acids (FFA1) are located on the basolateral face of EECs [254] so free-fatty acids can only access receptors to trigger hormone release following their absorption across the epithelium. GPR119, by contrast, is reported to be located apically as well as basolaterally [255].
Free fatty acids can enter enterocytes by the fatty acid transporter CD36. Expression of CD36 is described in oral taste bud cells [256,257] and intestine, though levels are higher in the proximal than distal intestine [258]. CD36 is an integral membrane protein which mediates the cellular uptake of long-chain fatty acids and has been linked to lingual and small intestinal fat detection and transport [241,259,260]. Mice lacking CD36 exhibit reduced chylomicron formation [261] and reduced fatty acid and cholesterol uptake in the proximal, but not distal intestine [262]. CD36 knock-out mice are also insensitive to the suppression of food intake mediated by duodenal lipid infusion [241]. Such participation of CD36 in lipid signalling mechanisms is likely indirect. As discussed above, transepithelial transport of fatty acids is necessary to target basolaterally located free-fatty acid receptors, and CD36 was also postulated to assist in FFA4 signalling by accumulating long-chain fatty acids in the vicinity of low affinity FFA4 receptors [263]. In addition, uptake of fatty acids is required for the synthesis of OEA and its precursor N-oleoyl-phosphatidylethanolamine (NOPE), and mice lacking CD36 had reduced OEA production and OEA-induced satiety [241].

4.4. Other: Bile Acids

Bile acids are released into the gut lumen following fat detection in the duodenum and release of CCK from EECs which stimulates gall bladder contraction. Bile acids contribute to the emulsification of dietary lipids, aiding digestion, and can also activate selective G-protein coupled bile acid receptors (GPBAR1) also known as Takeda G-protein coupled receptor 5 (TGR5) [264,265], as well as the nuclear farnesoid X receptor, FXR [266].
Early studies in anesthetised dogs found that intra-ileal infusion of bile increased secretion of proglucagon gene products [267,268]. EEC cell line models and mouse primary intestinal epithelial cultures confirmed that bile acids stimulated GLP-1 release [269,270,271] alongside clinical studies showing that luminal bile acids increased GLP-1 and PYY release [272,273]. Intraluminal infusion of bile into human participants and rodents also triggered a strong GIP secretory response [274,275].
Bile acid triggered gut hormone release is dependent on GPBAR1 as selective agonists of the receptor, designed with preferential activity for GPBAR1 over the nuclear bile acid receptor FXR, triggered GLP-1 release and mice lacking the receptor exhibited abrogated GLP-1 secretory responses to bile acids [112,269,276]. Ex vivo intestinal models demonstrated that only basolaterally accessible selective GPBAR1 agonists were able to elicit GLP-1 release [112,277], indicating that absorption of bile acids is critical for their effectiveness in triggering gut hormone secretion. Conjugated bile acids require transport into cells by the apical sodium-dependent bile acid transporter, ASBT [278], which is predominantly expressed in the terminal ileum. Consequently, blocking the absorption of bile acids with an inhibitor of ASBT blocked the effectiveness of bile acids to trigger GLP-1 release [112,277]. In the colon where expression of ASTB is lower, bile acids gain access to basolaterally located GPBAR1 following deconjugation and dehydroxylation by gut bacteria, which improves bile acid permeability and potency at GPBAR1 [265] and negates the dependence of bile acid-triggered GLP-1 release on ASBT [277]. Binding to GPBAR1 activates adenylyl cyclase cAMP production [112,270,271] downstream of Gαs proteins, and elevations in intracellular Ca2+ have also been reported in cell models of GLP-1 releasing cells following application of bile acids [112,269]. The activation of GPBAR1 in GLP-1 expressing cells increased evoked action potential activity and increased calcium currents through L-type voltage-gated calcium channels [108]. Bile acids may also increase the size of the L-cell population in the intestinal epithelium [276]. By contrast, FXR receptor activity inhibited GLP-1 production in L-cells [279], and the consequences of simultaneous activation of intestinal GPBAR1 and FXR receptors by bile acids still need to be reconciled. Bile acids may also contribute to reducing food intake by GLP-1 independent mechanisms: for example, GPBAR1 expression was found on inhibitory motor neurons in the myenteric plexus, corresponding with reduced spontaneous intestinal contractile activity following bile acid application [280].

4.5. Other: Short-Chain Fatty Acids (SCFAs)

Colonic fermentation of undigested dietary fibre by gut bacteria produces high concentrations of SCFAs in the intestinal lumen. In humans, SCFAs were most abundant in the caecum and ascending colon, whereas levels of circulating SCFA were substantially lower [27].
SCFA signalling is mediated by G-protein coupled receptors of the free-fatty acid receptor family. Free-fatty acid receptor 2 (FFA2), formerly known as GPR43 [281,282], is activated by SCFAs of 2–4 carbon lengths but acetate (C2) and propionate (C3) are the most potent. Ffar2 expression is highest in the distal gut and using an antibody raised against rat FFA2, the receptor co-localised with PYY, and presumably GLP-1, in rats [283] and humans [284]. FFA2 was also described in serotonin-containing mast cells but not enterochromaffin cells, clarifying a mechanism for SCFA stimulation of colonic motility [285]. An Ffar2-reporter mouse line confirmed Ffar2 expression in intestinal mast cells; however, there was weak association with EECs [286]. FFA2 is postulated to couple to G-proteins of the Gαq/11 and Gαi/o family, though the pathways recruited in EECs are incompletely elucidated. Activation of FFA2 produces a rise in intracellular calcium in murine L-cells [286,287], presumably by canonical Gαq-coupled pathways involving phospholipase C-dependent production of inositol-3-phosphate and Ca2+ release from intracellular stores. The free-fatty acid receptor 3 (FFA3), formerly known as GPR41 [288], is the other member of the FFA receptor family responsive to SCFAs. It preferentially binds SCFAs of three to five carbon lengths and couples to G-proteins of the Gαi/o family. FFA3 is localised to the colonic mucosa and co-localises with PYY [289]. In an Ffar3-reporter mouse model, a widespread pattern of Ffar3 expression was described in EECs, including GLP-1 and GIP-expressing cells, as well as peripheral neuronal ganglia associated with enteric, sensory and autonomic signalling [286,290]. The cellular mechanism of FFA3 signalling in EECs has not been elucidated, though in neurons has been shown to involve inhibition of voltage-gated calcium channels by a Gβγ-mediated mechanism [291]. Other receptors responsive to SCFAs include the olfactory receptor subfamily 51E (OR51E1 in human, Olfr558 in mouse; OR51E2 in human, Olfr78 in mouse), and GPR109A, also known as the niacin receptor.
In murine primary colonic cultures, SCFAs stimulated GLP-1 release through FFA2-dependent [292] or both FFA2 and FFA3-dependent mechanisms [287]. A selective FFA2 agonist, Compound 1, also stimulated GLP-1 secretion [293]. SCFA-stimulated GLP-1 release did not involve pertussis toxin-sensitive pathways and SCFAs triggered a rise in intracellular calcium in GLP-1-producing EECs, suggesting Gαq-coupled pathways are responsible for mediating the stimulatory effects of SCFAs on EECs. Furthermore, a selective inhibitor of Gαq signalling FR900359 abolished SCFA-triggered GLP-1 release and an FFA2 agonist with biased activity for Gαi-signalling, AZ1729, did not trigger GLP-1 secretion [294]. The molecular details of FFA3 involvement in stimulating GLP-1 release are still unclear, given that FFA3 seems to signal exclusively through inhibitory Gαi/o-coupled pathways. In an isolated perfused colon model, basolaterally applied SCFAs triggered release of GLP-1 which was more pronounced following enhancement of cAMP levels [295]. Given the high production of SCFA luminally, localisation of receptors on the basolateral surface may be a mechanism preventing continuous activation or saturation of SCFA receptors. However, the receptors involved in enhancing GLP-1 release following SCFA administration in this ex vivo model were not clear as selective FFA2 or FFA3 agonists and antagonists had no effect. Perhaps another SCFA receptor is involved or synergistic activation of both FFA2 and FFA3 is required for mediating the effects of SCFAs on GLP-1 expressing cells. Basolateral, rather than apical sensing of SCFA on GLP-1 secreting cells is, however, further supported by the correlation of circulating rather than faecal SCFA concentrations with GLP-1 in fasting humans [296].
As mentioned previously, FFA3 co-localised with GIP in the proximal small intestine [286]. SCFAs can arise in the proximal small intestine from fermentation by oral microbiota [297] but at much lower concentrations compared with the distal intestine, or from pathological bacterial overgrowth. Generation of SCFAs in the proximal small intestine is associated with suppression of GIP release by an FFA3-dependent mechanism [298].
Prolonged SCFA exposure can also alter gene expression, including genes encoding gut hormones. Proglucagon expression increased in STC-1 cells after 24 h incubation with SCFAs and in the colon of rats fed chow diets enriched with fermentable fibre [299]. In human cell line models of EECs, there was little increase in proglucagon expression following SCFA treatment, but a pronounced increase in PYY expression following butyrate treatment which was mediated by histone deacetylase (HDAC) inhibition and FFA2 [300]. SCFA have also been reported to alter the number of GLP-1 producing L-cells, but whereas an increase in L-cell number was observed following chronic SCFA exposure in mouse and human intestinal organoids in vitro [301], SCFA appeared to suppress L-cell number and function in germ free mice recolonised with SCFA producing bacteria [302,303].

5. Effect of GLP-1 and GIP on Food Intake and Weight Loss

The consensus view is that exogenous administration of GLP-1 reduces energy intake and appetite in rodents [304] and humans [305,306,307]. The anorexigenic actions of GLP-1 extend to long-acting mimetics such as exenatide, liraglutide and semaglutide [308,309,310,311] and persist in Type 2 diabetic and obese individuals, thus supporting the use of GLP-1 receptor agonists in weight management therapies.
The effect of GIP on food intake is more controversial. Administration of GIP or a long-acting analogue acyl-GIP, at doses sufficient to produce positive metabolic effects, had no effect on body weight in mice [312] or hunger scores and food ingestion in humans [197], and in a recent double-blind crossover study, there was no significant change in food intake in overweight/obese individuals given GIP along with an IV glucose infusion that mimicked the blood glucose rise after an oral glucose challenge [313]. Animal models with diminished GIP activity including Gipr knockout mice [314], enteroendocrine K-cell ablation [315], and GIPR antagonism [316,317] show reduced body weight gain associated with diet-induced obesity. By contrast, overexpression of GIP in mice also led to decreased energy intake and reduced weight gain associated with diet-induced obesity [318]. Interpretation of these studies may be complicated by confounding factors such as GIP stimulation of pancreatic insulin and glucagon secretion, which can themselves modulate appetite [319,320], and compensatory changes in the absence of functional GIP signalling [321]. Furthermore, GIP has a complicated pharmacology as certain peptides declared to be antagonists of GIPR in fact display partial agonist activity [322,323] and interspecies differences in GIPR pharmacology exist [322]. Recent studies have favoured a reduction of food intake as a consequence of pharmacological GIPR agonism [324], but further studies will be required to fully elucidate the molecular mechanisms governing GIP action on food intake.
Despite the controversy surrounding GIP action on food intake, there is growing interest in developing agonists that target two or more receptors involved in metabolic control, including receptors for the incretin hormones GLP-1 and GIP. Studies in mice demonstrated that co-administration of GIP and GLP-1 analogues enhanced body weight loss and improved glycaemic control in obese mice [312,325,326]. Dual GIP/GLP-1 agonists are in development as potential new anti-diabetic and anti-obesity treatments. The pharmacological reasoning for this approach is to amplify the beneficial metabolic effects whilst lowering doses and adverse side effect profiles of the individual components. Promising improvements in blood glucose control were obtained with the acylated unimolecular dual GIP/GLP-1 agonist RG7697/NNC0090-2746 in healthy [327] and T2D patients [328] and significant reductions in body weight were achieved but the magnitude was possibly similar to liraglutide treatment alone [329]. However, another dual GIP/GLP-1 agonist tirzepatide (LY3298176) displayed improved efficacy over the selective GLP-1 receptor agonist dulaglutide [330,331].

6. Gut Hormone-Mediated Mechanisms of Satiety

Infusion of various nutrients into the small intestine is associated with greater suppression of food intake than nutrients that are delivered intravenously [332,333], strongly supporting a gut-derived mechanism for satiety. GLP-1 and GIP released from the gastrointestinal tract can initiate satiety signals through a variety of mechanisms ranging from direct modulation of gastrointestinal function to recruitment of central circuits involved in feeding behaviours.
The actions of GLP-1 and GIP on gastric emptying and gastric acid secretion are well studied. Briefly, postprandial GLP-1 is associated with delayed gastric emptying [334] whereas GIP was found to have the opposite effect of increasing the gastric emptying rate [197]. Vagal afferent fibres were necessary for mediating GLP-1 effects on gastric emptying [335] along with cholinergic signalling and GLP1R activation [336], which is consistent with GLP1R expression described in the gastric antrum and pylorus [337]. Other gut hormones such as CCK and PYY also control gastric emptying rates [338]. Vagal innervation was also necessary in mediating the inhibitory effects of GLP-1 on gastric acid secretion [339] as vagotomised patients did not exhibit an inhibitory response to GLP-1 [340]. Peripheral infusion of GIP also inhibited gastric acid secretion, albeit at supraphysiological concentrations [341] and involved a reduction in postprandial gastrin release.
Gut hormones also internally regulate intestinal motility and luminal transit in response to a high nutrient load. Infusion of physiological levels of GLP-1 slowed intestinal motility in the fed and fasted state in rats [342] and humans [342,343], and was reversed by the GLP1R antagonist Exendin(9–39) [342]. Exogenous GIP also reduced intestinal transit in mice, which was blocked by a somatostatin receptor antagonist [344], but by comparison GLP-1 was more potent than GIP in abolishing myoelectric activity in the small bowel of rats [345].
Overall, by modulating various aspects of gastrointestinal function (gastric emptying, gastric acid secretion and intestinal motility), GLP-1 and GIP can slow the digestion of nutrients allowing more efficient nutrient breakdown and absorption. This serves to increase exposure of proximal EECs to nutrients. The slowing of gastrointestinal activity can also increase distension of the stomach and intestinal wall, and consequent activation of stretch-activated mechanoreceptors directly signals to the brain to control feeding and enhances the perception of fullness [346].
Neuronal mechanisms play an integral role in mediating the satiating effects of GLP-1 and possibly GIP. The effect of peripherally administered GLP-1 on food intake was ablated after bilateral subdiaphragmatic total vagotomy [25] in rats, and the suppression of short-term food intake following intraduodenal nutrient infusion was eliminated by selective vagal rhizotomy of the celiac branch, which abolishes afferent vagal inputs to the intestine [26]. Both findings highlight the significance of vagal signalling in appetite control. Vagal afferent neurons project peripherally to visceral organs, including much of the small intestine and the proximal third of the large intestine. A subpopulation of vagal afferent neurons express receptors for GLP-1 [337,347,348] but single-cell transcriptomics studies were unable to resolve expression of receptors for GIP in nodose neurons. GLP-1 directly stimulated activity in vagal afferent neurons [349] and augmented Ca2+ responses in GLP1R-expressing nodose neurons in vitro [337]. Vagal afferent neurons project to two brainstem regions: the area postrema and the nucleus of the solitary tract (NTS). GLP1R expression, staining and ligand binding have been detected in a number of brain regions, including the area postrema, ventromedial hypothalamus, arcuate and paraventricular nuclei [350,351,352,353]. Sensory afferent inputs into the hindbrain are essential for mediating the suppression of feeding by nutrients as chemical destruction of these neurons by capsaicin abolished this inhibition [354,355]. Some neurons in the NTS also produce and release GLP-1 [356] but these neurons do not express GLP1R or respond to exogenous GLP-1 [357,358]. Although vagal afferent signalling is important in mediating some effects of GLP-1 on feeding control, the route of intestinal GLP-1 to activate vagal afferent neurons is unclear. Vagal afferents, particularly neurons that express gut hormone receptors, do not appear to innervate the basolateral surface of L-cells [348,359], although Glp1r-expressing enteric neurons have been identified [337], which may be involved in relaying signals. Diffusion of GLP-1 in the vicinity of its site of release from L-cells may contribute to non-synaptic activation of local nerve endings, and could act in tandem with other mediators released from EECs such as glutamate [360] or ATP [361] to exert its effects on food intake.
Central injection of GLP-1 into the cerebral ventricles [304,362,363,364] or hypothalamus [365] reduces food intake. These centrally mediated effects of GLP-1 were dependent on GLP1R signalling as they were not observed in Glp1r KO animals [366] or in the presence of the GLP1R antagonist exendin(9–39) [362,367]. Whilst some of the central GLP1R sites are involved in the satiating effects of peripherally supplied GLP1R agonists, it is becoming clear that activation of GLP-1 expressing neurons in the NTS recruits additional anorexic circuits [358].
There is increasing support for a central mechanism mediating food intake reduction by GIP. Central administration of GIP decreased food intake and body weight and combined with GLP-1 produced a synergistic reduction of food consumption [367]. Furthermore, Gipr expression has been described in the arcuate, dorsomedial, and paraventricular nuclei in the hypothalamus and activation of Gipr expressing cells in this region suppressed food intake [368]. Many questions remain, including whether circulating GIP released from K-cells in the proximal small intestine activates centrally expressed GIPR, which are also found in the area postrema [368], as well as the identity and relative importance of the circuitry mediating GIP-mediated suppression of food intake.
Besides its role in mediating homeostatic feeding, GLP-1 is also implicated in circuits driving food reward and motivation. Activation of GLP-1 producing preproglucagon neurons in the NTS seems to be of minor importance for homeostatic feeding control, but becomes relevant under stress, one of which is overeating, presumably involving stomach stretching [369]. In this context it is important to note that Glp1r is expressed by neurons in the ventral tegmental area (VTA) and nucleus accumbens, brain regions associated with reward and desire. Exendin-4, a GLP-1 mimetic, or GLP-1 injected peripherally reduced palatable food intake and reward-motivated behaviour and a similar effect was observed when exendin-4 was injected directly into the VTA [370,371]. In agreement with these findings, injection of the GLP1R antagonist, Exendin(9–39), into the nucleus accumbens increased meal size and palatability of sucrose solutions [372].

7. Future Directions

There are still many unresolved issues in our understanding of nutrient-induced gut hormone secretory mechanisms, including the role of GLP-1 produced in the distal GI tract. GLP-1 is produced in large quantities in the distal small intestine and colon, regions of the intestine not typically reached by ingested nutrients, where its release may follow different regulatory mechanisms including control by metabolites of gut microbiota and neurohormonal pathways. Recent studies identified the possible importance of distal L-cells in mediating responses to lipid-derivatives, melanocortin receptor 4 agonists and lipopolysaccharides [28]. Another question is whether therapeutic activation of GPCRs in EECs could trigger sufficient gut hormone release to produce significant metabolic benefits or satiation. Injectable GLP1 receptor agonists and the levels of GLP-1 achieved following bariatric surgery are far in excess of levels expected in a normal healthy postprandial state. The feasibility of targeting nutrient-sensing receptors in EECs also needs to be properly assessed. Many of these receptors respond to a range of ligands but the potential importance of biased downstream signalling is not known. Finally, desensitisation of nutrient-sensing receptors in EECs has not been investigated in-depth, but could arise from the prolonged exposure of EECs to saturating concentrations of nutrients in the postprandial state or with altered dietary status such as in models of diet-induced obesity, which have been shown to alter expression levels of peptide hormone receptors expressed on vagal afferent neurons [373]. Further complications or, alternatively considered, treatment options arise from cross talk between different EECs, notably the apparent chronic paracrine inhibition of L-cells by somatostatin [374] released from nearby D-cells, which themselves express a number of nutrient sensing receptors, as shown by RNAseq for D-cells in the stomach [375].

8. Conclusions

Nutrient absorption and stimulation of gut hormone secretion occur predominantly in the proximal small intestine. However, as demonstrated by the effectiveness of bariatric surgery to deliver nutrients more distally along the gut, distal GLP-1 producing L-cells represent a vast source of endogenous GLP-1 that can potentially be exploited to suppress appetite and increase insulin secretion. Future work is needed to reveal the physiological importance of distal GLP-1 producing L-cells and identify strategies to recruit these cell populations for improved glucose homeostasis and appetite regulation.

Author Contributions

V.B.L., F.M.G., F.R. wrote, revised, and approved the submitted version of this review. All authors have read and agreed to the published version of the manuscript.

Funding

V.B.L., F.M.G. and F.R. were funded by a Wellcome Joint Investigator Award to FR and FMG (106262/Z/14/Z and 106263/Z/14/Z) and the Metabolic Diseases Unit (MRC_MC_UU_12012/3).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The FR/FMG lab receives additional grant support from AstraZeneca and Eli Lilly for unrelated work. FMG is a consultant for Kallyope (New York, NY, USA).

References

  1. Bayliss, W.M.; Starling, E.H. The mechanism of pancreatic secretion. J. Physiol. 1902, 28, 325–353. [Google Scholar] [CrossRef]
  2. Bohórquez, D.V.; Chandra, R.; Samsa, L.A.; Vigna, S.R.; Liddle, R.A. Characterization of basal pseudopod-like processes in ileal and colonic PYY cells. J. Mol. Histol. 2010, 42, 3–13. [Google Scholar] [CrossRef] [Green Version]
  3. Bohórquez, D.V.; Samsa, L.A.; Roholt, A.; Medicetty, S.; Chandra, R.; Liddle, R.A. An Enteroendocrine cell—enteric glia connection revealed by 3D electron microscopy. PLoS ONE 2014, 9, e89881. [Google Scholar] [CrossRef] [Green Version]
  4. McSwiney, B.A.; Spurrell, W.R. Influence of osmotic pressure upon the emptying time of the stomach. J. Physiol. 1933, 79, 437–442. [Google Scholar] [CrossRef] [Green Version]
  5. Knapper, J.; Heath, A.; Fletcher, J.; Morgan, L.; Marks, V. GIP and GLP-1(7–36) amide secretion in response to intraduodenal infusions of nutrients in pigs. Comp. Biochem. Physiol. Part C Pharm. Toxicol. Endocrinol. 1995, 111, 445–450. [Google Scholar] [CrossRef]
  6. Rehfeld, J.F. The new biology of gastrointestinal hormones. Physiol. Rev. 1998, 78, 1087–1108. [Google Scholar] [CrossRef] [Green Version]
  7. Umar, S. Intestinal stem cells. Curr. Gastroenterol. Rep. 2010, 12, 340–348. [Google Scholar] [CrossRef] [PubMed]
  8. Creutzfeldt, M.W. The incretin concept today. Diabetologia 1979, 16, 75–85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Elrick, H.; Stimmler, L.; Hlad, C.J.; Arai, Y. Plasma insulin response to oral and intravenous glucose administration. Clin. Endocrinol. Metab. 1964, 24, 1076–1082. [Google Scholar] [CrossRef]
  10. Le Roux, C.W.; Welbourn, R.; Werling, M.; Osborne, A.; Kokkinos, A.; Laurenius, A.; Lönroth, H.; Fändriks, L.; Ghatei, M.A.; Bloom, S.R.; et al. Gut hormones as mediators of appetite and weight loss after Roux-en-Y gastric bypass. Ann. Surg. 2007, 246, 780–785. [Google Scholar] [CrossRef] [PubMed]
  11. Jørgensen, N.B.; Dirksen, C.; Bojsen-Møller, K.N.; Jacobsen, S.H.; Worm, D.; Hansen, D.L.; Kristiansen, V.B.; Naver, L.; Madsbad, S.; Holst, J.J. Exaggerated glucagon-like peptide 1 response is important for improved β-cell function and glucose tolerance after Roux-en-Y gastric bypass in patients with Type 2 diabetes. Diabetes 2013, 62, 3044–3052. [Google Scholar] [CrossRef] [Green Version]
  12. Hansen, K.B.; Vilsbøll, T.; Knop, F.K. Incretin mimetics: A novel therapeutic option for patients with type 2 diabetes—A review. Diabetes Metab. Syndr. Obes. Targets 2010, 3, 155–163. [Google Scholar]
  13. Mortensen, K.; Christensen, L.L.; Holst, J.J.; Orskov, C. GLP-1 and GIP are colocalized in a subset of endocrine cells in the small intestine. Regul. Pept. 2003, 114, 189–196. [Google Scholar] [CrossRef]
  14. Theodorakis, M.J.; Carlson, O.; Michopoulos, S.; Doyle, M.E.; Juhaszova, M.; Petraki, K.; Egan, J.M. Human duodenal enteroendocrine cells: Source of both incretin peptides, GLP-1 and GIP. Am. J. Physiol. Metab. 2006, 290, E550–E559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Haber, A.L.; Biton, M.; Rogel, N.; Herbst, R.H.; Shekhar, K.; Smillie, C.; Burgin, G.; DeLorey, T.M.; Howitt, M.R.; Katz, Y.; et al. A single-cell survey of the small intestinal epithelium. Nat. Cell Biol. 2017, 551, 333–339. [Google Scholar] [CrossRef]
  16. Egerod, K.L.; Engelstoft, M.S.; Grunddal, K.V.; Nøhr, M.K.; Secher, A.; Sakata, I.; Pedersen, J.; Windeløv, J.A.; Füchtbauer, E.-M.; Olsen, J.; et al. A major lineage of enteroendocrine cells coexpress CCK, secretin, GIP, GLP-1, PYY, and neurotensin but not somatostatin. Endocrinology 2012, 153, 5782–5795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Habib, A.M.; Richards, P.; Cairns, L.S.; Rogers, G.J.; Bannon, C.A.M.; Parker, H.E.; Morley, T.C.E.; Yeo, G.S.H.; Reimann, F.; Gribble, F.M.; et al. Overlap of endocrine hormone expression in the mouse intestine revealed by transcriptional profiling and flow cytometry. Endocrinology 2012, 153, 3054–3065. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Jorsal, T.; Rhee, N.A.; Pedersen, J.; Wahlgren, C.D.; Mortensen, B.; Jepsen, S.L.; Jelsing, J.; Dalbøge, L.S.; Vilmann, P.; Hassan, H.; et al. Enteroendocrine K and L cells in healthy and Type 2 diabetic individuals. Diabetologia 2018, 61, 284–294. [Google Scholar] [CrossRef] [PubMed]
  19. Roberts, G.P.; Larraufie, P.; Richards, P.; Kay, R.G.; Galvin, S.G.; Miedzybrodzka, E.L.; Leiter, A.; Li, H.J.; Glass, L.L.; Ma, M.K.; et al. Comparison of human and murine enteroendocrine cells by transcriptomic and peptidomic profiling. Diabetes 2019, 68, 1062–1072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Sjölund, K.; Sandén, G.; Håkanson, R.; Sundler, F. Endocrine cells in human intestine: An immunocytochemical study. Gastroenterology 1983, 85, 1120–1130. [Google Scholar] [CrossRef]
  21. Bryant, M.G.; Bloom, S.R.; Polak, J.M.; Hobbs, S.; Domschke, W.; Mitznegg, P.; Ruppin, H.; Demling, L. Measurement of gut hormonal peptides in biopsies from human stomach and proximal small intestine. Gut 1983, 24, 114–119. [Google Scholar] [CrossRef]
  22. Buchan, A.M.J.; Polak, J.M.; Capella, C.; Solcia, E.; Pearse, A.G.E. Electronimmunocytochemical evidence for the K cell localization of gastric inhibitory polypeptide (GIP) im man. Histochem. Cell Biol. 1978, 56, 37–44. [Google Scholar] [CrossRef] [PubMed]
  23. Eissele, R.; Göke, R.; Willemer, S.; Harthus, H.-P.; Vermeer, H.; Arnold, R.; Göke, B. Glucagon-like peptide-1 cells in the gastrointestinal tract and pancreas of rat, pig and man. Eur. J. Clin. Investig. 1992, 22, 283–291. [Google Scholar] [CrossRef] [PubMed]
  24. Meyer, J.H.; Hlinka, M.; Tabrizi, Y.; Dimaso, N.; Raybould, H.E. Chemical specificities and intestinal distributions of nutrient-driven satiety. Am. J. Physiol. Integr. Comp. Physiol. 1998, 275, R1293–R1307. [Google Scholar] [CrossRef] [PubMed]
  25. Abbott, C.R.; Monteiro, M.; Small, C.J.; Sajedi, A.; Smith, K.L.; Parkinson, J.R.; Ghatei, M.A.; Bloom, S.R. The inhibitory effects of peripheral administration of peptide YY3–36 and glucagon-like peptide-1 on food intake are attenuated by ablation of the vagal–brainstem–hypothalamic pathway. Brain Res. 2005, 1044, 127–131. [Google Scholar] [CrossRef] [PubMed]
  26. Walls, E.K.; Phillips, R.J.; Wang, F.B.; Holst, M.C.; Powley, T.L. Suppression of meal size by intestinal nutrients is eliminated by celiac vagal deafferentation. Am. J. Physiol. Integr. Comp. Physiol. 1995, 269, R1410–R1419. [Google Scholar] [CrossRef] [PubMed]
  27. Cummings, J.H.; Pomare, E.W.; Branch, W.J.; Naylor, C.P.; Macfarlane, G.T. Short chain fatty acids in human large intestine, portal, hepatic and venous blood. Gut 1987, 28, 1221–1227. [Google Scholar] [CrossRef] [Green Version]
  28. Panaro, B.L.; Yusta, B.; Matthews, D.; Koehler, J.A.; Song, Y.; Sandoval, D.A.; Drucker, D.J. Intestine-selective reduction of Gcg expression reveals the importance of the distal gut for GLP-1 secretion. Mol. Metab. 2020, 37, 100990. [Google Scholar] [CrossRef]
  29. Buffa, R.; Solcia, E.; Go, V.L.W. Immunohistochemical identification of the cholecystokinin cell in the intestinal mucosa. Gastroenterology 1976, 70, 528–532. [Google Scholar] [CrossRef]
  30. Hökfelt, T.; Herrera-Marschitz, M.; Seroogy, K.; Ju, G.; A Staines, W.; Holets, V.; Schalling, M.; Ungerstedt, U.; Post, C.; Rehfeld, J.F.; et al. Immunohistochemical studies on cholecystokinin (CCK)-immunoreactive neurons in the rat using sequence specific antisera and with special reference to the caudate nucleus and primary sensory neurons. J. Chem. Neuroanat. 1988, 1, 11–51. [Google Scholar]
  31. Rehfeld, J. Immunochemical studies on cholecystokinin. II: Distribution and molecular heterogeneity in the central nervous system and small intestine of man and hog. J. Biol. Chem. 1978, 253, 4022–4030. [Google Scholar] [CrossRef]
  32. Kerstens, P.; Lamers, C.; Jansen, J.; De Jong, A.; Hessels, M.; Hafkenscheid, J. Physiological plasma concentrations of cholecystokinin stimulate pancreatic enzyme secretion and gallbladder contraction in man. Life Sci. 1985, 36, 565–569. [Google Scholar] [CrossRef]
  33. Gibbs, J.; Smith, G.P. Cholecystokinin and satiety in rats and rhesus monkeys. Am. J. Clin. Nutr. 1977, 30, 758–761. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Gibbs, J.; Young, R.C.; Smith, G.P. Cholecystokinin decreases food intake in rats. J. Comp. Physiol. Psychol. 1973, 84, 488–495. [Google Scholar] [CrossRef]
  35. Saito, A.; Williams, J.A.; Goldfine, I.D. Alterations in brain cholecystokinin receptors after fasting. Nat. Cell Biol. 1981, 289, 599–600. [Google Scholar] [CrossRef]
  36. Little, T.J.; Horowitz, M.; Feinle-Bisset, C. Role of cholecystokinin in appetite control and body weight regulation. Obes. Rev. 2005, 6, 297–306. [Google Scholar] [CrossRef] [PubMed]
  37. Sykaras, A.G.; Demenis, C.; Cheng, L.; Pisitkun, T.; McLaughlin, J.T.; Fenton, R.A.; Smith, C.P. Duodenal CCK cells from male mice express multiple hormones including ghrelin. Endocrinology 2014, 155, 3339–3351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Roth, K.A.; Kim, S.; Gordon, J.I. Immunocytochemical studies suggest two pathways for enteroendocrine cell differentiation in the colon. Am. J. Physiol. Liver Physiol. 1992, 263, G174–G180. [Google Scholar] [CrossRef]
  39. Gerspach, A.C.; Steinert, R.E.; Schönenberger, L.; Graber-Maier, A.; Beglinger, C. The role of the gut sweet taste receptor in regulating GLP-1, PYY, and CCK release in humans. Am. J. Physiol. Metab. 2011, 301, E317–E325. [Google Scholar] [CrossRef]
  40. Hopman, W.P.M.; Jansen, J.B.M.J.; Lamers, C.B.H.W. Comparative Study of the effects of equal amounts of fat, protein, and starch on plasma cholecystokinin in man. Scand. J. Gastroenterol. 1985, 20, 843–847. [Google Scholar] [CrossRef]
  41. Liddle, R.A.; Goldfine, I.D.; Rosen, M.S.; Taplitz, R.A.; Williams, J.A. Cholecystokinin bioactivity in human plasma. Molecular forms, responses to feeding, and relationship to gallbladder contraction. J. Clin. Investig. 1985, 75, 1144–1152. [Google Scholar] [CrossRef]
  42. Fothergill, L.J.; Callaghan, B.; Hunne, B.; Bravo, D.M.; Furness, J.B. Costorage of enteroendocrine hormones evaluated at the cell and subcellular levels in male mice. Endocrinology 2017, 158, 2113–2123. [Google Scholar] [CrossRef]
  43. Rindi, G.; Ratineau, C.; Ronco, A.; E Candusso, M.; Tsai, M.; Leiter, A.B. Targeted ablation of secretin-producing cells in transgenic mice reveals a common differentiation pathway with multiple enteroendocrine cell lineages in the small intestine. Development 1999, 126, 4149–4156. [Google Scholar]
  44. Beumer, J.; Artegiani, B.; Post, Y.; Reimann, F.; Gribble, F.; Nguyen, T.N.; Zeng, H.; Born, M.V.D.; Van Es, J.H.; Clevers, H.; et al. Enteroendocrine cells switch hormone expression along the crypt-to-villus BMP signalling gradient. Nat. Cell Biol. 2018, 20, 909–916. [Google Scholar] [CrossRef]
  45. Afroze, S.; Meng, F.; Jensen, K.; McDaniel, K.; Rahal, K.; Onori, P.; Gaudio, E.; Alpini, G.; Glaser, S.S. The physiological roles of secretin and its receptor. Ann. Transl. Med. 2013, 1, 29. [Google Scholar] [PubMed]
  46. Cheng, C.Y.Y.; Chu, J.Y.S.; Chow, B.K.C. Central and peripheral administration of secretin inhibits food intake in mice through the activation of the melanocortin system. Neuropsychopharmacology 2010, 36, 459–471. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Motojima, Y.; Kawasaki, M.; Matsuura, T.; Saito, R.; Yoshimura, M.; Hashimoto, H.; Ueno, H.; Maruyama, T.; Suzuki, H.; Ohnishi, H.; et al. Effects of peripherally administered cholecystokinin-8 and secretin on feeding/drinking and oxytocin-mRFP1 fluorescence in transgenic rats. Neurosci. Res. 2016, 109, 63–69. [Google Scholar] [CrossRef] [PubMed]
  48. Chu, J.Y.S.; Cheng, C.Y.Y.; Sekar, R.; Chow, B.K.C. Vagal afferent mediates the anorectic effect of peripheral secretin. PLoS ONE 2013, 8, e64859. [Google Scholar] [CrossRef] [Green Version]
  49. Lopez, M.J.; Upchurch, B.H.; Rindi, G.; Leiter, A.B. Studies in Transgenic mice reveal potential relationships between secretin-producing cells and other endocrine cell types. J. Biol. Chem. 1995, 270, 885–891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Feurle, G.; Hamscher, G.; Kusiek, R.; Meyer, H.; Metzger, J. Identification of xenin, a xenopsin-related peptide, in the human gastric mucosa and its effect on exocrine pancreatic secretion. J. Biol. Chem. 1992, 267, 22305–22309. [Google Scholar] [CrossRef]
  51. Anlauf, M.; Weihe, E.; Hartschuh, W.; Hamscher, G.; Feurle, G.E. Localization of xenin-immunoreactive cells in the duodenal mucosa of humans and various mammals. J. Histochem. Cytochem. 2000, 48, 1617–1626. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Feurle, G. Xenin—A review. Peptides 1998, 19, 609–615. [Google Scholar] [CrossRef]
  53. Feurle, G.E.; Ikonomu, S.; Partoulas, G.; Stoschus, B.; Hamscher, G. Xenin plasma concentrations during modified sham feeding and during meals of different composition demonstrated by radioimmunoassay and chromatography. Regul. Pept. 2003, 111, 153–159. [Google Scholar] [CrossRef]
  54. Wice, B.M.; Wang, S.; Crimmins, D.L.; Diggs-Andrews, K.A.; Althage, M.C.; Ford, E.L.; Tran, H.; Ohlendorf, M.; Griest, T.A.; Wang, Q.; et al. Xenin-25 potentiates glucose-dependent insulinotropic polypeptide action via a novel cholinergic relay mechanism. J. Biol. Chem. 2010, 285, 19842–19853. [Google Scholar] [CrossRef] [Green Version]
  55. Feurle, G.E.; Heger, M.; Niebergall-Roth, E.; Teyssen, S.; Fried, M.; Eberle, C.; Singer, M.V.; Hamscher, G. Gastroenteropancreatic effects of xenin in the dog. J. Pept. Res. 2009, 49, 324–330. [Google Scholar] [CrossRef] [PubMed]
  56. Feurle, G.E.; Pfeiffer, A.; Schmidt, T.; Dominguez-Munoz, E.; Malfertheiner, P.; Hamscher, G. Phase III of the migrating motor complex: Associated with endogenous xenin plasma peaks and induced by exogenous xenin. Neurogastroenterol. Motil. 2001, 13, 237–246. [Google Scholar] [CrossRef]
  57. Cooke, J.H.; Patterson, M.; Patel, S.R.; Smith, K.L.; Ghatei, M.A.; Bloom, S.R.; Murphy, K.G. Peripheral and central administration of xenin and neurotensin suppress food intake in rodents. Obesity 2009, 17, 1135–1143. [Google Scholar] [CrossRef]
  58. Kim, E.R.; Mizuno, T.M. Role of neurotensin receptor 1 in the regulation of food intake by neuromedins and neuromedin-related peptides. Neurosci. Lett. 2010, 468, 64–67. [Google Scholar] [CrossRef]
  59. Dakka, T.; Cuber, J.C.; Chayvialle, J.A. Functional coupling between the active transport of glucose and the secretion of intestinal neurotensin in rats. J. Physiol. 1993, 469, 753–765. [Google Scholar] [CrossRef] [Green Version]
  60. Kuhre, R.E.; Gribble, F.M.; Hartmann, B.; Reimann, F.; Windeløv, J.A.; Rehfeld, J.F.; Holst, J.J. Fructose stimulates GLP-1 but not GIP secretion in mice, rats, and humans. Am. J. Physiol. Liver Physiol. 2014, 306, G622–G630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Walker, J.P.; Fujimura, M.; Sakamoto, T.; Greeley, G.H.; Townsend, C.M.; Thompson, J.C. Importance of the ileum in neurotensin released by fat. Surgery 1985, 98, 224–229. [Google Scholar]
  62. Hawkins, M.F. Central nervous system neurotensin and feeding. Physiol. Behav. 1986, 36, 1–8. [Google Scholar] [CrossRef]
  63. Luttinger, D.; King, R.A.; Sheppard, D.; Strupp, J.; Nemeroff, C.B.; Prange, A.J. The effect of neurotensin on food consumption in the rat. Eur. J. Pharm. 1982, 81, 499–503. [Google Scholar] [CrossRef]
  64. Remaury, A.; Vita, N.; Gendreau, S.; Jung, M.; Arnone, M.; Poncelet, M.; Culouscou, J.-M.; Le Fur, G.; Soubrié, P.; Caput, D.; et al. Targeted inactivation of the neurotensin type 1 receptor reveals its role in body temperature control and feeding behavior but not in analgesia. Brain Res. 2002, 953, 63–72. [Google Scholar] [CrossRef]
  65. Polak, J.M.; Sullivan, S.N.; Bloom, S.R.; Buchan, A.M.J.; Facer, P.; Brown, M.R.; Pearse, A.G.E. Specific localisation of neurotensin to the N cell in human intestine by radioimmunoassay and immunocytochemistry. Nat. Cell Biol. 1977, 270, 183–184. [Google Scholar] [CrossRef] [PubMed]
  66. Grunddal, K.V.; Ratner, C.F.; Svendsen, B.; Sommer, F.; Engelstoft, M.S.; Madsen, A.N.; Pedersen, J.; Nøhr, M.K.; Egerod, K.L.; Nawrocki, A.R.; et al. Neurotensin is coexpressed, coreleased, and acts together with GLP-1 and PYY in enteroendocrine control of metabolism. Endocrinology 2016, 157, 176–194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Pocai, A.; Carrington, P.E.; Adams, J.R.; Wright, M.; Eiermann, G.; Zhu, L.; Du, X.; Petrov, A.; Lassman, M.E.; Jiang, G.; et al. Glucagon-like peptide 1/glucagon receptor dual agonism reverses obesity in mice. Diabetes 2009, 58, 2258–2266. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Dakin, C.L.; Gunn, I.; Small, C.J.; Edwards, C.M.B.; Hay, D.L.; Smith, D.M.; Ghatei, M.A.; Bloom, S.R. Oxyntomodulin inhibits food intake in the rat. Endocrinology 2001, 142, 4244–4250. [Google Scholar] [CrossRef]
  69. Dakin, C.L.; Small, C.J.; Batterham, R.L.; Neary, N.M.; Cohen, M.A.; Patterson, M.; Ghatei, M.A.; Bloom, S.R. Peripheral oxyntomodulin reduces food intake and body weight gain in rats. Endocrinology 2004, 145, 2687–2695. [Google Scholar] [CrossRef] [Green Version]
  70. Kosinski, J.R.; Hubert, J.; Carrington, P.E.; Chicchi, G.G.; Mu, J.; Miller, C.; Cao, J.; Bianchi, E.; Pessi, A.; Sinharoy, R.; et al. The glucagon receptor is involved in mediating the body weight-lowering effects of oxyntomodulin. Obesity 2012, 20, 1566–1571. [Google Scholar] [CrossRef] [Green Version]
  71. Cohen, M.A.; Ellis, S.M.; Le Roux, C.W.; Batterham, R.L.; Park, A.; Patterson, M.M.; Frost, G.S.; Ghatei, M.A.; Bloom, S.R. Oxyntomodulin suppresses appetite and reduces food intake in humans. J. Clin. Endocrinol. Metab. 2003, 88, 4696–4701. [Google Scholar] [CrossRef]
  72. Parker, V.E.R.; Robertson, D.; Wang, T.; Hornigold, D.C.; Petrone, M.; Cooper, A.T.; Posch, M.G.; Heise, T.; Plum-Moerschel, L.; Schlichthaar, H.; et al. Efficacy, safety, and mechanistic insights of cotadutide, a dual receptor glucagon-like peptide-1 and glucagon agonist. J. Clin. Endocrinol. Metab. 2019, 105, 803–820. [Google Scholar] [CrossRef] [PubMed]
  73. Adrian, T.; Ferri, G.-L.; Bacarese-Hamilton, A.; Fuessl, H.; Polak, J.; Bloom, S. Human distribution and release of a putative new gut hormone, peptide YY. Gastroenterology 1985, 89, 1070–1077. [Google Scholar] [CrossRef]
  74. Billing, L.J.; Smith, C.A.; Larraufie, P.; Goldspink, D.A.; Galvin, S.; Kay, R.G.; Howe, J.D.; Walker, R.; Pruna, M.; Glass, L.; et al. Co-storage and release of insulin-like peptide-5, glucagon-like peptide-1 and peptide YY from murine and human colonic enteroendocrine cells. Mol. Metab. 2018, 16, 65–75. [Google Scholar] [CrossRef] [PubMed]
  75. Pedersen-Bjergaard, U.; Høt, U.; Kelbæk, H.; Schifter, S.; Rehfeld, J.F.; Faber, J.; Christensen, N.J. Influence of meal composition on postprandial peripheral plasma concentrations of vasoactive peptides in man. Scand. J. Clin. Lab. Investig. 1996, 56, 497–503. [Google Scholar] [CrossRef]
  76. Batterham, R.L.; Cowley, M.A.; Small, C.J.; Herzog, H.; Cohen, M.A.; Dakin, C.L.; Wren, A.M.; Brynes, A.E.; Low, M.J.; Ghatei, M.A.; et al. Gut hormone PYY3–36 physiologically inhibits food intake. Nat. Cell Biol. 2002, 418, 650–654. [Google Scholar] [CrossRef]
  77. Halatchev, I.G.; Ellacott, K.L.J.; Fan, W.; Cone, R.D. Peptide YY3–36Inhibits food intake in mice through a melanocortin-4 Receptor-independent mechanism. Endocrinology 2004, 145, 2585–2590. [Google Scholar] [CrossRef] [Green Version]
  78. Lewis, J.E.; Miedzybrodzka, E.L.; Foreman, R.E.; Woodward, O.R.M.; Kay, R.G.; Goldspink, D.A.; Gribble, F.M.; Reimann, F. Selective stimulation of colonic L cells improves metabolic outcomes in mice. Diabetologia 2020, 63, 1396–1407. [Google Scholar] [CrossRef]
  79. Babu, M.; Purhonen, A.; Bansiewicz, T.; Mäkelä, K.; Walkowiak, J.; Miettinen, P.; Herzig, K. Effect of total colectomy and PYY infusion on food intake and body weight in rats. Regul. Pept. 2005, 131, 29–33. [Google Scholar] [CrossRef] [PubMed]
  80. Boggiano, M.M.; Chandler, P.C.; Oswald, K.D.; Rodgers, R.J.; Blundell, J.E.; Ishii, Y.; Beattie, A.H.; Holch, P.; Allison, D.B.; Schindler, M.; et al. PYY3–36 as an anti-obesity drug target. Obes. Rev. 2005, 6, 307–322. [Google Scholar] [CrossRef]
  81. Tschöp, M.; Castañeda, T.R.; Joost, H.G.; Thöne-Reineke, C.; Ortmann, S.; Klaus, S.; Hagan, M.M.; Chandler, P.C.; Oswald, K.D.; Benoit, S.C.; et al. Does gut hormone PYY3–36 decrease food intake in rodents? Nat. Cell Biol. 2004, 430, 1–3. [Google Scholar] [CrossRef]
  82. Fu-Cheng, X.; Anini, Y.; Chariot, J.; Castex, N.; Galmiche, J.-P.; Rozé, C. Mechanisms of peptide YY release induced by an intraduodenal meal in rats: Neural regulation by proximal gut. Pflügers Arch. 1997, 433, 571–579. [Google Scholar] [CrossRef]
  83. Koda, S.; Date, Y.; Murakami, N.; Shimbara, T.; Hanada, T.; Toshinai, K.; Niijima, A.; Furuya, M.; Inomata, N.; Osuye, K.; et al. The role of the vagal nerve in peripheral PYY3–36-induced feeding reduction in rats. Endocrinology 2005, 146, 2369–2375. [Google Scholar] [CrossRef] [Green Version]
  84. Yang, H.; Tache, Y. PYY in brain stem nuclei induces vagal stimulation of gastric acid secretion in rats. Am. J. Physiol. Liver Physiol. 1995, 268, G943–G948. [Google Scholar] [CrossRef]
  85. Zeng, N.; Walsh, J.; Kang, T.; Wu, S.; Sachs, G. Peptide YY inhibition of rat gastric enterochromaffin-like cell function. Gastroenterology 1997, 112, 127–135. [Google Scholar] [CrossRef]
  86. Lin, H.C.; Zhao, X.T.; Wang, L.; Wong, H. Fat-induced ileal brake in the dog depends on peptide YY. Gastroenterology 1996, 110, 1491–1495. [Google Scholar] [CrossRef] [PubMed]
  87. Savage, A.P.; E Adrian, T.; Carolan, G.; Chatterjee, V.K.; Bloom, S.R. Effects of peptide YY (PYY) on mouth to caecum intestinal transit time and on the rate of gastric emptying in healthy volunteers. Gut 1987, 28, 166–170. [Google Scholar] [CrossRef]
  88. Wiley, J.W.; Lu, Y.; Owyang, C. Mechanism of action of peptide YY to inhibit gastric motility. Gastroenterology 1991, 100, 865–872. [Google Scholar] [CrossRef]
  89. Grosse, J.; Heffron, H.; Burling, K.; Hossain, M.A.; Habib, A.M.; Rogers, G.J.; Richards, P.; Larder, R.; Rimmington, D.; Adriaenssens, A.A.; et al. Insulin-like peptide 5 is an orexigenic gastrointestinal hormone. Proc. Natl. Acad. Sci. USA 2014, 111, 11133–11138. [Google Scholar] [CrossRef] [Green Version]
  90. Lee, Y.S.; De Vadder, F.; Tremaroli, V.; Wichmann, A.; Mithieux, G.; Bäckhed, F. Insulin-like peptide 5 is a microbially regulated peptide that promotes hepatic glucose production. Mol. Metab. 2016, 5, 263–270. [Google Scholar] [CrossRef]
  91. Drucker, D.J.; Jin, T.; Asa, S.L.; Young, T.A.; Brubaker, P.L. Activation of proglucagon gene transcription by protein kinase—A in a novel mouse enteroendocrine cell line. Mol. Endocrinol. 1994, 8, 1646–1655. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. De Bruïne, A.P.; Dinjens, W.N.M.; Pijls, M.M.J.; Linden, E.P.M.V.D.; Rousch, M.J.M.; Moerkerk, P.T.; De Goeij, A.F.P.M.; Bosnian, F.T. NCI-H716 cells as a model for endocrine differentiation in colorectal cancer. Virchows Arch. B 1992, 62, 311–320. [Google Scholar] [CrossRef]
  93. Park, J.G.; Oie, H.K.; Sugarbaker, P.H.; Henslee, J.G.; Chen, T.R.; Johnson, B.E.; Gazdar, A. Characteristics of cell lines estab-lished from human colorectal carcinoma. Cancer Res. 1987, 47, 6710–6718. [Google Scholar] [PubMed]
  94. Brubaker, P.L.; Schloos, J.; Drucker, D.J. Regulation of glucagon-like peptide-1 synthesis and secretion in the GLUTag Enteroendocrine cell line. Endocrinology 1998, 139, 4108–4114. [Google Scholar] [CrossRef] [PubMed]
  95. Reimer, R.A.; Darimont, C.; Gremlich, S.; Nicolas-Métral, V.; Rüegg, U.T.; Macé, K. A human cellular model for studying the regulation of glucagon-like peptide-1 secretion. Endocrinology 2001, 142, 4522–4528. [Google Scholar] [CrossRef]
  96. Cao, X.; Flock, G.; Choi, C.; Irwin, D.M.; Drucker, D.J. Aberrant regulation of human intestinal proglucagon gene expression in the NCI-H716 cell line. Endocrinology 2003, 144, 2025–2033. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Rindi, G.; Grant, S.G.; Yiangou, Y.; Ghatei, M.A.; Bloom, S.R.; Bautch, V.L.; Solcia, E.; Polak, J.M. Development of neuroen-docrine tumors in the gastrointestinal tract of transgenic mice. Heterogeneity of hormone expression. Am. J. Pathol. 1990, 136, 1349–1363. [Google Scholar]
  98. Cheung, A.T.; Dayanandan, B.; Lewis, J.T.; Korbutt, G.S.; Rajotte, R.V.; Bryer-Ash, M.; Boylan, M.O.; Wolfe, M.M.; Kieffer, T.J. Glucose-dependent insulin release from genetically engineered K cells. Science 2000, 290, 1959–1962. [Google Scholar] [CrossRef] [PubMed]
  99. Kieffer, T.J.; Huang, Z.; McIntosh, C.H.; Buchan, A.M.; Brown, J.C.; Pederson, R.A. Gastric inhibitory polypeptide release from a tumor-derived cell line. Am. J. Physiol. Metab. 1995, 269, E316–E322. [Google Scholar] [CrossRef]
  100. Ramshur, E.B.; Rull, T.R.; Wice, B.M. Novel insulin/GIP co-producing cell lines provide unexpected insights into Gut K-cell function In Vivo. J. Cell. Physiol. 2002, 192, 339–350. [Google Scholar] [CrossRef]
  101. Aponte, G.W.; Taylor, I.L.; Soll, A.H. Primary culture of PYY cells from canine colon. Am. J. Physiol. Liver Physiol. 1988, 254, G829–G836. [Google Scholar] [CrossRef]
  102. Kieffer, T.J.; Buchan, A.M.; Barker, H.; Brown, J.C.; Pederson, R.A. Release of gastric inhibitory polypeptide from cultured canine endocrine cells. Am. J. Physiol. Metab. 1994, 267, E489–E496. [Google Scholar] [CrossRef] [PubMed]
  103. Brubaker, P.L.; Vranic, M. Fetal rat intestinal cells in monolayer culture: A new In Vitro System to study the glucagon-like immunoreactive peptides. Endocrinology 1987, 120, 1976–1985. [Google Scholar] [CrossRef] [PubMed]
  104. Reimann, F.; Habib, A.M.; Tolhurst, G.; Parker, H.E.; Rogers, G.J.; Gribble, F.M. Glucose sensing in L cells: A primary cell study. Cell Metab. 2008, 8, 532–539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Habib, A.M.; Richards, P.; Rogers, G.J.; Reimann, F.; Gribble, F.M. Co-localisation and secretion of glucagon-like peptide 1 and peptide YY from primary cultured human L cells. Diabetologia 2013, 56, 1413–1416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Parker, H.E.; Habib, A.M.; Rogers, G.J.; Gribble, F.M.; Reimann, F. Nutrient-dependent secretion of glucose-dependent insulinotropic polypeptide from primary murine K cells. Diabetologia 2009, 52, 289–298. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Sato, T.; Vries, R.G.; Snippert, H.J.; Van De Wetering, M.; Barker, N.; Stange, D.E.; Van Es, J.H.; Abo, A.; Kujala, P.; Peters, P.J.; et al. Single Lgr5 stem cells build crypt-villus structures In Vitro without a mesenchymal niche. Nature 2009, 459, 262–265. [Google Scholar] [CrossRef] [PubMed]
  108. Goldspink, D.A.; Lu, V.B.; Billing, L.J.; Larraufie, P.; Tolhurst, G.; Gribble, F.M.; Reimann, F. Mechanistic insights into the detection of free fatty and bile acids by ileal glucagon-like peptide-1 secreting cells. Mol. Metab. 2018, 7, 90–101. [Google Scholar] [CrossRef]
  109. Beumer, J.; Puschhof, J.; -Martinez, J.B.; Martínez-Silgado, A.; Elmentaite, R.; James, K.R.; Ross, A.; Hendriks, D.; Artegiani, B.; Busslinger, G.A.; et al. High-resolution mRNA and secretome atlas of human enteroendocrine cells. Cell 2020, 182, 1062–1064. [Google Scholar] [CrossRef]
  110. Goldspink, D.A.; Lu, V.B.; Miedzybrodzka, E.L.; Smith, C.A.; Foreman, R.E.; Billing, L.J.; Kay, R.G.; Reimann, F.; Gribble, F.M. Labeling and characterization of human GLP-1-secreting L-cells in primary ileal organoid culture. Cell Rep. 2020, 31, 107833. [Google Scholar] [CrossRef]
  111. Ussing, H.H.; Zerahn, K. Active transport of sodium as the source of electric current in the short-circuited isolated frog skin. Acta Physiol. Scand. 1951, 23, 110–127. [Google Scholar] [CrossRef]
  112. Brighton, C.A.; Rievaj, J.; Kuhre, R.E.; Glass, L.L.; Schoonjans, K.; Holst, J.J.; Gribble, F.M.; Reimann, F. Bile acids trigger GLP-1 release predominantly by accessing basolaterally located G protein—coupled bile acid receptors. Endocrinology 2015, 156, 3961–3970. [Google Scholar] [CrossRef] [Green Version]
  113. Joshi, S.; Tough, I.R.; Cox, H.M. Endogenous PYY and GLP-1 mediatel-glutamine responses in intestinal mucosa. Br. J. Pharm. 2013, 170, 1092–1101. [Google Scholar] [CrossRef] [PubMed]
  114. Wen, J.; Phillips, S.F.; Sarr, M.G.; Kost, L.J.; Holst, J.J. PYY and GLP-1 contribute to feedback inhibition from the canine ileum and colon. Am. J. Physiol. Liver Physiol. 1995, 269, G945–G952. [Google Scholar] [CrossRef]
  115. Ritzel, U.; Fromme, A.; Ottleben, M.; Leonhardt, U.; Ramadori, G. Release of glucagon-like peptide-1 (GLP-1) by carbohydrates in the perfused rat ileum. Acta Diabetol. 1997, 34, 18–21. [Google Scholar] [CrossRef]
  116. Cook, C.G.; Andrews, J.M.; Jones, K.L.; Wittert, G.A.; Chapman, I.M.; Morley, J.E.; Horowitz, M. Effects of small intestinal nutrient infusion on appetite and pyloric motility are modified by age. Am. J. Physiol. Content 1997, 273, R755–R761. [Google Scholar] [CrossRef]
  117. Lavin, J.H.; Wittert, G.A.; Andrews, J.; Yeap, B.; Wishart, J.M.; Morris, H.A.; Morley, J.E.; Horowitz, M.; Read, N.W. Interaction of insulin, glucagon-like peptide 1, gastric inhibitory polypeptide, and appetite in response to intraduodenal carbohydrate. Am. J. Clin. Nutr. 1998, 68, 591–598. [Google Scholar] [CrossRef]
  118. Elliott, R.M.; Morgan, L.M.; Tredger, J.A.; Deacon, S.; Wright, J.; Marks, V. Glucagon-like peptide-1(7–36) amide and glucose-dependent insulinotropic polypeptide secretion in response to nutrient ingestion in man: Acute post-prandial and 24-h secretion patterns. J. Endocrinol. 1993, 138, 159–166. [Google Scholar] [CrossRef]
  119. Herman, G.A.; Bergman, A.; Stevens, C.; Kotey, P.; Yi, B.; Zhao, P.; Dietrich, B.; Golor, G.; Schrodter, A.; Keymeulen, B.; et al. Effect of single oral doses of sitagliptin, a dipeptidyl peptidase-4 inhibitor, on incretin and plasma glucose levels after an oral glucose tolerance test in patients with Type 2 diabetes. J. Clin. Endocrinol. Metab. 2006, 91, 4612–4619. [Google Scholar] [CrossRef]
  120. Schirra, J.; Katschinski, M.; Weidmann, C.; Schäfer, T.; Wank, U.; Arnold, R.; Göke, B. Gastric emptying and release of incretin hormones after glucose ingestion in humans. J. Clin. Investig. 1996, 97, 92–103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Ferraris, R.P.; Yasharpour, S.; Lloyd, K.C.; Mirzayan, R.; Diamond, J.M. Luminal glucose concentrations in the gut under normal conditions. Am. J. Physiol. Liver Physiol. 1990, 259, G822–G837. [Google Scholar] [CrossRef]
  122. Stephen, A.M.; Haddad, A.C.; Phillips, S.F. Passage of carbohydrate into the colon. Direct measurements in humans. Gastroenterology 1983, 85, 589–595. [Google Scholar] [CrossRef]
  123. Sun, E.W.; De Fontgalland, D.; Rabbitt, P.; Hollington, P.; Sposato, L.; Due, S.L.; Wattchow, D.A.; Rayner, C.K.; Deane, A.M.; Young, R.L.; et al. Mechanisms controlling glucose-induced GLP-1 secretion in human small intestine. Diabetes 2017, 66, 2144–2149. [Google Scholar] [CrossRef] [Green Version]
  124. Nauck, M.A.; Siemsgluss, J.; Orskov, C.; Holst, J.J. Release of glucagon-like peptide 1 (GLP-1 [7–36 amide]), gastric inhibi-tory polypeptide (GIP) and insulin in response to oral glucose after upper and lower intestinal resections. Zeitschrift Gastroenterologie 1996, 34, 159–166. [Google Scholar]
  125. Yoshikawa, T.; Inoue, R.; Matsumoto, M.; Yajima, T.; Ushida, K.; Iwanaga, T. Comparative expression of hexose transporters (SGLT1, GLUT1, GLUT2 and GLUT5) throughout the mouse gastrointestinal tract. Histochem. Cell Biol. 2011, 135, 183–194. [Google Scholar] [CrossRef]
  126. Yoshida, A.; Takata, K.; Kasahara, T.; Aoyagi, T.; Saito, S.; Hirano, H. Immunohistochemical localization of Na(+)-dependent glucose transporter in the rat digestive tract. Histochem. J. 1995, 27, 420–426. [Google Scholar] [CrossRef]
  127. Hwang, E.-S.; Hirayama, B.A.; Wright, E.M. Distribution of the SGLT1 Na+glucose cotransporter and mRNA along the crypt-villus axis of rabbit small intestine. Biochem. Biophys. Res. Commun. 1991, 181, 1208–1217. [Google Scholar] [CrossRef]
  128. Gribble, F.M.; Williams, L.; Simpson, A.K.; Reimann, F. A novel glucose-sensing mechanism contributing to glucagon-like peptide-1 secretion from the GLUTag cell line. Diabetes 2003, 52, 1147–1154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Díez-Sampedro, A.; Lostao, M.; Wright, E.; Hirayama, B. Glycoside binding and translocation in Na+-dependent glucose cotransporters: Comparison of SGLT1 and SGLT3. J. Membr. Biol. 2000, 176, 111–117. [Google Scholar] [CrossRef]
  130. Landau, B.R.; Bernstein, L.; Wilson, T.H. Hexose transport by hamster intestine in vitro. Am. J. Physiol. Content 1962, 203, 237–240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Csáky, T.Z.; Glenn, J.E. Urinary recovery of 3-methylglucose administered to rats. Am. J. Physiol. Content 1956, 188, 159–162. [Google Scholar] [CrossRef]
  132. Sykes, S.; Morgan, L.M.; English, J.; Marks, V. Evidence for preferential stimulation of gastric inhibitory polypeptide secretion in the rat by actively transported carbohydrates and their analogues. J. Endocrinol. 1980, 85, 201–207. [Google Scholar] [CrossRef]
  133. Gorboulev, V.; Schürmann, A.; Vallon, V.; Kipp, H.; Jaschke, A.; Klessen, D.; Friedrich, A.; Scherneck, S.; Rieg, T.; Cunard, R.; et al. Na+-D-glucose cotransporter SGLT1 is pivotal for intestinal glucose absorption and glucose-dependent incretin secretion. Diabetes 2011, 61, 187–196. [Google Scholar] [CrossRef] [Green Version]
  134. Powell, D.R.; Smith, M.; Greer, J.; Harris, A.; Zhao, S.; Dacosta, C.; Mseeh, F.; Shadoan, M.K.; Sands, A.; Zambrowicz, B.; et al. LX4211 increases serum glucagon-like peptide 1 and peptide YY levels by reducing sodium/glucose cotransporter 1 (SGLT1)–mediated absorption of intestinal glucose. J. Pharm. Exp. 2013, 345, 250–259. [Google Scholar] [CrossRef]
  135. Díez-Sampedro, A.; Hirayama, B.A.; Osswald, C.; Gorboulev, V.; Baumgarten, K.; Volk, C.; Wright, E.M.; Koepsell, H. A glucose sensor hiding in a family of transporters. Proc. Natl. Acad. Sci. USA 2003, 100, 11753–11758. [Google Scholar] [CrossRef] [Green Version]
  136. Kellett, G.L.; Helliwell, P.A. The diffusive component of intestinal glucose absorption is mediated by the glucose-induced recruitment of GLUT2 to the brush-border membrane. Biochem. J. 2000, 350, 155–162. [Google Scholar] [CrossRef]
  137. Affleck, J.A.; Helliwell, P.A.; Kellett, G.L. Immunocytochemical detection of GLUT2 at the rat intestinal brush-border membrane. J. Histochem. Cytochem. 2003, 51, 1567–1574. [Google Scholar] [CrossRef] [Green Version]
  138. Scow, J.S.; Iqbal, C.W.; Jones, T.W.; Qandeel, H.G.; Zheng, Y.; Duenes, J.A.; Nagao, M.; Madhavan, S.; Sarr, M.G.; Madhaven, S.; et al. Absence of evidence of translocation of GLUT2 to the apical membrane of enterocytes in everted intestinal sleeves. J. Surg. Res. 2011, 167, 56–61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Röder, P.V.; Geillinger, K.E.; Zietek, T.S.; Thorens, B.; Koepsell, H.; Daniel, H. The role of SGLT1 and GLUT2 in intestinal glucose transport and sensing. PLoS ONE 2014, 9, e89977. [Google Scholar] [CrossRef]
  140. Parker, H.E.; Adriaenssens, A.; Rogers, G.; Richards, P.; Koepsell, H.; Reimann, F.; Gribble, F.M. Predominant role of active versus facilitative glucose transport for glucagon-like peptide-1 secretion. Diabetologia 2012, 55, 2445–2455. [Google Scholar] [CrossRef] [Green Version]
  141. Jetton, T.L.; Liang, Y.; Pettepher, C.C.; Zimmerman, E.C.; Cox, F.G.; Horvath, K.; Matschinsky, F.M.; Magnuson, M.A. Analysis of upstream glucokinase promoter activity in transgenic mice and identification of glucokinase in rare neuroen-docrine cells in the brain and gut. J. Biol. Chem. 1994, 269, 3641–3654. [Google Scholar] [CrossRef]
  142. Murphy, R.; Tura, A.; Clark, P.M.; Holst, J.J.; Mari, A.; Hattersley, A.T. Glucokinase, the pancreatic glucose sensor, is not the gut glucose sensor. Diabetologia 2008, 52, 154–159. [Google Scholar] [CrossRef] [Green Version]
  143. Rorsman, P. The pancreatic beta-cell as a fuel sensor: An electrophysiologist’s viewpoint. Diabetologia 1997, 40, 487–495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Nielsen, L.B.; Ploug, K.B.; Swift, P.; Ørskov, C.; Jansen-Olesen, I.; Chiarelli, F.; Holst, J.J.; Hougaard, P.; Pörksen, S.; Holl, R.; et al. Co-localisation of the Kir6.2/SUR1 channel complex with glucagon-like peptide-1 and glucose-dependent insulinotrophic polypeptide expression in human ileal cells and implications for glycaemic control in new onset type 1 diabetes. Eur. J. Endocrinol. 2007, 156, 663–671. [Google Scholar] [CrossRef]
  145. Reimann, F.; Gribble, F.M. Glucose-sensing in glucagon-like peptide-1-secreting cells. Diabetes 2002, 51, 2757–2763. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Kuhre, R.E.; Frost, C.R.; Svendsen, B.; Holst, J.J. Molecular mechanisms of glucose-stimulated GLP-1 secretion from perfused rat small intestine. Diabetes 2014, 64, 370–382. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Mangel, A.W.; Prpic, V.; Snow, N.D.; Basavappa, S.; Hurst, L.J.; Sharara, A.I.; Liddle, R.A. Regulation of cholecystokinin secretion by ATP-sensitive potassium channels. Am. J. Physiol. Liver Physiol. 1994, 267, G595–G600. [Google Scholar] [CrossRef] [PubMed]
  148. Tsukiyama, K.; Yamada, Y.; Miyawaki, K.; Hamasaki, A.; Nagashima, K.; Hosokawa, M.; Fujimoto, S.; Takahashi, A.; Toyoda, K.; Toyokuni, S.; et al. Gastric inhibitory polypeptide is the major insulinotropic factor in K(ATP) null mice. Eur. J. Endocrinol. 2004, 151, 407–412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Stephens, J.; Bodvarsdottir, T.; Wareham, K.; Prior, S.; Bracken, R.; Lowe, G.; Rumley, A.; Dunseath, G.; Luzio, S.; Deacon, C.; et al. Effects of short-term therapy with glibenclamide and repaglinide on incretin hormones and oxidative damage associated with postprandial hyperglycaemia in people with type 2 diabetes mellitus. Diabetes Res. Clin. Pract. 2011, 94, 199–206. [Google Scholar] [CrossRef] [PubMed]
  150. El-Ouaghlidi, A.; Rehring, E.; Holst, J.J.; Schweizer, A.; Foley, J.; Holmes, D.; Nauck, M.A. The dipeptidyl peptidase 4 inhibitor vildagliptin does not accentuate glibenclamide-induced hypoglycemia but reduces glucose-induced glucagon-like peptide 1 and gastric inhibitory polypeptide secretion. J. Clin. Endocrinol. Metab. 2007, 92, 4165–4171. [Google Scholar] [CrossRef]
  151. Moran, T.H.; McHugh, P.R. Distinctions among three sugars in their effects on gastric emptying and satiety. Am. J. Physiol. Integr. Comp. Physiol. 1981, 241, R25–R30. [Google Scholar] [CrossRef] [PubMed]
  152. Rayner, C.K.; Park, H.S.; Wishart, J.M.; Kong, M.-F.; Doran, S.M.; Horowitz, M. Effects of intraduodenal glucose and fructose on antropyloric motility and appetite in healthy humans. Am. J. Physiol. Integr. Comp. Physiol. 2000, 278, R360–R366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Kong, M.-F.; Chapman, I.; Goble, E.; Wishart, J.; Wittert, G.; Morris, H.; Horowitz, M. Effects of oral fructose and glucose on plasma GLP-1 and appetite in normal subjects. Peptides 1999, 20, 545–551. [Google Scholar] [CrossRef]
  154. Shima, K.; Suda, T.; Nishimoto, K.; Yoshimoto, S. Relationship between molecular structures of sugars and their ability to stimulate the release of glucagon-like peptide-1 from canine ileal loops. Eur. J. Endocrinol. 1990, 123, 464–470. [Google Scholar] [CrossRef]
  155. Burant, C.; Takeda, J.; Brot-Laroche, E.; Bell, G.; Davidson, N. Fructose transporter in human spermatozoa and small intestine is GLUT5. J. Biol. Chem. 1992, 267, 14523–14526. [Google Scholar] [CrossRef]
  156. Rand, E.B.; De Paoli, A.M.; Davidson, N.O.; Bell, G.I.; Burant, C.F. Sequence, tissue distribution, and functional characterization of the rat fructose transporter Glutam. J. Physiol. Liver Physiol. 1993, 264, G1169–G1176. [Google Scholar] [CrossRef]
  157. Seino, Y.; Ogata, H.; Maekawa, R.; Izumoto, T.; Iida, A.; Harada, N.; Miki, T.; Seino, S.; Inagaki, N.; Tsunekawa, S.; et al. Fructose induces glucose-dependent insulinotropic polypeptide, glucagon-like peptide-1 and insulin secretion: Role of adenosine triphosphate-sensitive K+channels. J. Diabetes Investig. 2015, 6, 522–526. [Google Scholar] [CrossRef]
  158. Wong, G.T.; Gannon, K.S.; Margolskee, R.F. Transduction of bitter and sweet taste by gustducin. Nat. Cell Biol. 1996, 381, 796–800. [Google Scholar] [CrossRef] [PubMed]
  159. Nelson, G.; Chandrashekar, J.; Hoon, M.A.; Feng, L.; Zhao, G.; Ryba, N.J.P.; Zuker, C.S. An amino-acid taste receptor. Nat. Cell Biol. 2002, 416, 199–202. [Google Scholar] [CrossRef]
  160. Rozengurt, N.; Wu, S.V.; Chen, M.C.; Huang, C.; Sternini, C.; Rozengurt, E. Colocalization of the α-subunit of gustducin with PYY and GLP-1 in L cells of human colon. Am. J. Physiol. Liver Physiol. 2006, 291, G792–G802. [Google Scholar] [CrossRef]
  161. Kokrashvili, Z.; Mosinger, B.; Margolskee, R.F. Taste signaling elements expressed in gut enteroendocrine cells regulate nutrient-responsive secretion of gut hormones. Am. J. Clin. Nutr. 2009, 90, 822S–825S. [Google Scholar] [CrossRef]
  162. Jang, H.-J.; Kokrashvili, Z.; Theodorakis, M.J.; Carlson, O.D.; Kim, B.-J.; Zhou, J.; Kim, H.H.; Xu, X.; Chan, S.L.; Juhaszova, M.; et al. Gut-expressed gustducin and taste receptors regulate secretion of glucagon-like peptide-1. Proc. Natl. Acad. Sci. USA 2007, 104, 15069–15074. [Google Scholar] [CrossRef] [Green Version]
  163. Fujita, Y.; Wideman, R.D.; Speck, M.; Asadi, A.; King, D.S.; Webber, T.D.; Haneda, M.; Kieffer, T.J. Incretin release from gut is acutely enhanced by sugar but not by sweeteners In Vivo. Am. J. Physiol. Metab. 2009, 296, E473–E479. [Google Scholar] [CrossRef] [PubMed]
  164. Bezençon, C.; Le Coutre, J.; Damak, S. Taste-signaling proteins are coexpressed in solitary intestinal epithelial cells. Chem. Senses 2007, 32, 41–49. [Google Scholar] [CrossRef] [Green Version]
  165. Margolskee, R.F.; Dyer, J.; Kokrashvili, Z.; Salmon, K.S.H.; Ilegems, E.; Daly, K.; Maillet, E.L.; Ninomiya, Y.; Mosinger, B.; Shirazi-Beechey, S.P. T1R3 and gustducin in gut sense sugars to regulate expression of Na+-glucose cotransporter. Proc. Natl. Acad. Sci. USA 2007, 104, 15075–15080. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Ma, J.; Chang, J.; Checklin, H.L.; Young, R.L.; Jones, K.L.; Horowitz, M.; Rayner, C.K. Effect of the artificial sweetener, sucralose, on small intestinal glucose absorption in healthy human subjects. Br. J. Nutr. 2010, 104, 803–806. [Google Scholar] [CrossRef] [Green Version]
  167. Saltiel, M.Y.; Kuhre, R.E.; Christiansen, C.B.; Eliasen, R.; Conde-Frieboes, K.W.; Rosenkilde, M.M.; Holst, J.J. Sweet taste receptor activation in the gut is of limited importance for glucose-stimulated GLP-1 and GIP secretion. Nutrients 2017, 9, 418. [Google Scholar] [CrossRef]
  168. Thomas, F.; Sinar, D.; Mazzaferri, E.; Cataland, S.; Mekhjian, H.; Caldwell, J.; Fromkes, J. Selective release of gastric inhibitory polypeptide by intraduodenal amino acid perfusion in man. Gastroenterology 1978, 74, 1261–1265. [Google Scholar] [CrossRef]
  169. Cordier-Bussat, M.; Bernard, C.; Levenez, F.; Klages, N.; Laser-Ritz, B.; Philippe, J.; Chayvialle, J.A.; Cuber, J.C. Peptones stimulate both the secretion of the incretin hormone glucagon-like peptide 1 and the transcription of the proglucagon gene. Diabetes 1998, 47, 1038–1045. [Google Scholar] [CrossRef]
  170. Reimer, R.A. Meat hydrolysate and essential amino acid-induced glucagon-like peptide-1 secretion, in the human NCI-H716 enteroendocrine cell line, is regulated by extracellular signal-regulated kinase1/2 and p38 mitogen-activated protein kinases. J. Endocrinol. 2006, 191, 159–170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Diakogiannaki, E.; Pais, R.; Tolhurst, G.; Parker, H.E.; Horscroft, J.; Rauscher, B.; Zietek, T.; Daniel, H.; Gribble, F.M.; Reimann, F.; et al. Oligopeptides stimulate glucagon-like peptide-1 secretion in mice through proton-coupled uptake and the calcium-sensing receptor. Diabetologia 2013, 56, 2688–2696. [Google Scholar] [CrossRef] [Green Version]
  172. Matsumura, K.; Miki, T.; Jhomori, T.; Gonoi, T.; Seino, S. Possible role of PEPT1 in gastrointestinal hormone secretion. Biochem. Biophys. Res. Commun. 2005, 336, 1028–1032. [Google Scholar] [CrossRef] [PubMed]
  173. Jiang, Y.; Rose, A.J.; Sijmonsma, T.P.; Bröer, A.; Pfenninger, A.; Herzig, S.; Schmoll, D.; Bröer, S. Mice lacking neutral amino acid transporter B0AT1 (Slc6a19) have elevated levels of FGF21 and GLP-1 and improved glycaemic control. Mol. Metab. 2015, 4, 406–417. [Google Scholar] [CrossRef]
  174. Pais, R.; Gribble, F.M.; Reimann, F. Signalling pathways involved in the detection of peptones by murine small intestinal enteroendocrine L-cells. Peptides 2016, 77, 9–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Mace, O.J.; Schindler, M.; Patel, S. The regulation of K- and L-cell activity by GLUT2 and the calcium-sensing receptor CasR in rat small intestine. J. Physiol. 2012, 590, 2917–2936. [Google Scholar] [CrossRef]
  176. Liou, A.P.; Sei, Y.; Zhao, X.; Feng, J.; Lu, X.; Thomas, C.; Pechhold, S.; Raybould, H.E.; Wank, S.A. The extracellular calcium-sensing receptor is required for cholecystokinin secretion in response to l-phenylalanine in acutely isolated intestinal I cells. Am. J. Physiol. Liver Physiol. 2011, 300, G538–G546. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Rey, O.; Young, S.H.; Jacamo, R.; Moyer, M.P.; Rozengurt, E. Extracellular calcium sensing receptor stimulation in human colonic epithelial cells induces intracellular calcium oscillations and proliferation inhibition. J. Cell. Physiol. 2010, 225, 73–83. [Google Scholar] [CrossRef] [Green Version]
  178. Modvig, I.M.; Kuhre, R.E.; Holst, J.J. Peptone-mediated glucagon-like peptide-1 secretion depends on intestinal absorption and activation of basolaterally located Calcium-sensing receptors. Physiol. Rep. 2019, 7, e14056. [Google Scholar] [CrossRef]
  179. Choi, S.; Lee, M.; Shiu, A.L.; Yo, S.J.; Halldén, G.; Aponte, G.W. GPR93 activation by protein hydrolysate induces CCK transcription and secretion in STC-1 cells. Am. J. Physiol. Liver Physiol. 2007, 292, G1366–G1375. [Google Scholar] [CrossRef] [PubMed]
  180. Lee, C.-W.; Rivera, R.; Gardell, S.; Dubin, A.E.; Chun, J. GPR92 as a New G12/13- and Gq-coupled lysophosphatidic acid receptor that increases cAMP, LPA5. J. Biol. Chem. 2006, 281, 23589–23597. [Google Scholar] [CrossRef] [Green Version]
  181. Rudenko, O.; Shang, J.; Munk, A.; Ekberg, J.P.; Petersen, N.; Engelstoft, M.S.; Egerod, K.L.; Hjorth, S.A.; Wu, M.; Feng, Y.; et al. The aromatic amino acid sensor GPR142 controls metabolism through balanced regulation of pancreatic and gut hormones. Mol. Metab. 2019, 19, 49–64. [Google Scholar] [CrossRef]
  182. Lin, H.V.; Efanov, A.M.; Fang, X.; Beavers, L.S.; Wang, X.; Wang, J.; Valcarcel, I.C.G.; Ma, T. GPR142 controls tryptophan-induced insulin and incretin hormone secretion to improve glucose metabolism. PLoS ONE 2016, 11, e0157298. [Google Scholar] [CrossRef]
  183. Li, X.; Staszewski, L.; Xu, H.; Durick, K.; Zoller, M.; Adler, E. Human receptors for sweet and umami taste. Proc. Natl. Acad. Sci. USA 2002, 99, 4692–4696. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Wang, J.-H.; Inoue, T.; Higashiyama, M.; Guth, P.H.; Engel, E.; Kaunitz, J.D.; Akiba, Y. Umami receptor activation increases duodenal bicarbonate secretion via glucagon-like peptide-2 release in rats. J. Pharm. Exp. 2011, 339, 464–473. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Blonde, G.D.; Travers, S.P.; Spector, A.C. Taste sensitivity to a mixture of monosodium glutamate and inosine 5′-monophosphate by mice lacking both subunits of the T1R1 + T1R3 amino acid receptor. Am. J. Physiol. Integr. Comp. Physiol. 2018, 314, R802–R810. [Google Scholar] [CrossRef]
  186. Lin, W.; Kinnamon, S.C. Physiological evidence for ionotropic and metabotropic glutamate receptors in rat taste cells. J. Neurophysiol. 1999, 82, 2061–2069. [Google Scholar] [CrossRef] [PubMed]
  187. Bigiani, A.; De Lay, R.J.; Chaudhari, N.; Kinnamon, S.C.; Roper, S.D. Responses to glutamate in rat taste cells. J. Neurophysiol. 1997, 77, 3048–3059. [Google Scholar] [CrossRef]
  188. Akiba, Y.; Watanabe, C.; Mizumori, M.; Kaunitz, J.D. Luminal l-glutamate enhances duodenal mucosal defense mechanisms via multiple glutamate receptors in rats. Am. J. Physiol. Liver Physiol. 2009, 297, G781–G791. [Google Scholar] [CrossRef] [Green Version]
  189. Wellendorph, P.; Hansen, K.B.; Balsgaard, A.; Greenwood, J.R.; Egebjerg, J.; Bräuner-Osborne, H. Deorphanization of GPRC6A: A promiscuous l-α-Amino acid receptor with preference for basic amino acids. Mol. Pharm. 2004, 67, 589–597. [Google Scholar] [CrossRef]
  190. Pi, M.; Quarles, L.D. Multiligand specificity and wide tissue expression of GPRC6A reveals new endocrine networks. Endocrinology 2012, 153, 2062–2069. [Google Scholar] [CrossRef] [Green Version]
  191. Mizokami, A.; Yasutake, Y.; Higashi, S.; Kawakubo-Yasukochi, T.; Chishaki, S.; Takahashi, I.; Takeuchi, H.; Hirata, M. Oral administration of osteocalcin improves glucose utilization by stimulating glucagon-like peptide-1 secretion. Bone 2014, 69, 68–79. [Google Scholar] [CrossRef]
  192. Wang, H.; Murthy, K.S.; Grider, J.R. Expression patterns of l-amino acid receptors in the murine STC-1 enteroendocrine cell line. Cell Tissue Res. 2019, 378, 471–483. [Google Scholar] [CrossRef]
  193. Oya, M.; Kitaguchi, T.; Pais, R.; Reimann, F.; Gribble, F.; Tsuboi, T. The G protein-coupled receptor family C group 6 subtype A (GPRC6A) receptor is involved in amino acid-induced glucagon-like peptide-1 secretion from GLUTag cells. J. Biol. Chem. 2013, 288, 4513–4521. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Carney, B.I.; Jones, K.L.; Horowitz, M.; Sun, W.M.; Hebbard, G.; Edelbroek, M.A.L. Stereospecific effects of tryptophan on gastric emptying and hunger in humans. J. Gastroenterol. Hepatol. 1994, 9, 557–563. [Google Scholar] [CrossRef] [PubMed]
  195. Steinert, R.E.; Luscombe-Marsh, N.D.; Little, T.J.; Standfield, S.; Otto, B.; Horowitz, M.; Feinle-Bisset, C. Effects of intraduodenal infusion of L-tryptophan on ad libitum eating, antropyloroduodenal motility, glycemia, insulinemia, and gut peptide secretion in healthy men. J. Clin. Endocrinol. Metab. 2014, 99, 3275–3284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. McVeay, C.; Fitzgerald, P.C.E.; Ullrich, S.S.; Steinert, R.E.; Horowitz, M.; Feinle-Bisset, C. Effects of intraduodenal administration of lauric acid and L-tryptophan, alone and combined, on gut hormones, pyloric pressures, and energy intake in healthy men. Am. J. Clin. Nutr. 2019, 109, 1335–1343. [Google Scholar] [CrossRef] [PubMed]
  197. Edholm, T.; Degerblad, M.; Grybäck, P.; Hilsted, L.; Holst, J.J.; Jacobsson, H.; Efendic, S.; Schmidt, P.T.; Hellström, P.M. Differential incretin effects of GIP and GLP-1 on gastric emptying, appetite, and insulin-glucose homeostasis. Neurogastroenterol. Motil. 2010, 22, 1191. [Google Scholar] [CrossRef] [PubMed]
  198. Reimann, F.; Williams, L.; Xavier, G.D.S.; Rutter, G.A.; Gribble, F.M. Glutamine potently stimulates glucagon-like peptide-1 secretion from GLUTag cells. Diabetologia 2004, 47, 1592–1601. [Google Scholar] [CrossRef] [Green Version]
  199. Greenfield, J.R.; Farooqi, I.S.; Keogh, J.M.; Henning, E.; Habib, A.M.; Blackwood, A.; Reimann, F.; Holst, J.J.; Gribble, F.M. Oral glutamine increases circulating glucagon-like peptide 1, glucagon, and insulin concentrations in lean, obese, and Type 2 diabetic subjects. Am. J. Clin. Nutr. 2008, 89, 106–113. [Google Scholar] [CrossRef]
  200. Chang, J.; Wu, T.; Greenfield, J.R.; Samocha-Bonet, D.; Horowitz, M.; Rayner, C.K. Effects of intraduodenal glutamine on incretin hormone and insulin release, the glycemic response to an intraduodenal glucose infusion, and antropyloroduodenal motility in health and Type 2 diabetes. Diabetes Care 2013, 36, 2262–2265. [Google Scholar] [CrossRef] [Green Version]
  201. Tolhurst, G.; Zheng, Y.; Parker, H.E.; Habib, A.M.; Reimann, F.; Gribble, F.M. Glutamine triggers and potentiates glucagon-like peptide-1 secretion by raising cytosolic Ca2+ and cAMP. Endocrinology 2011, 152, 405–413. [Google Scholar] [CrossRef] [Green Version]
  202. Mamillapalli, R.; Wysolmerski, J. The calcium-sensing receptor couples to Gαs and regulates PTHrP and ACTH secretion in pituitary cells. J. Endocrinol. 2009, 204, 287–297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Geoghegan, J.G.; A Cheng, C.; Lawson, C.; Pappas, T.N. The effect of caloric load and nutrient composition on induction of small intestinal satiety in dogs. Physiol. Behav. 1997, 62, 39–42. [Google Scholar] [CrossRef]
  204. Chapman, I.M.; A Goble, E.; A Wittert, G.; Horowitz, M. Effects of small-intestinal fat and carbohydrate infusions on appetite and food intake in obese and nonobese men. Am. J. Clin. Nutr. 1999, 69, 6–12. [Google Scholar] [CrossRef] [Green Version]
  205. Herrmann, C.; Göke, R.; Richter, G.; Fehmann, H.-C.; Arnold, R.; Göke, B. Glucagon-like peptide-1 and glucose-dependent insulin-releasing polypeptide plasma levels in response to nutrients. Digestion 1995, 56, 117–126. [Google Scholar] [CrossRef] [PubMed]
  206. Thomsen, C.; Rasmussen, O.; Lousen, T.; Holst, J.J.; Fenselau, S.; Schrezenmeir, J.; Hermansen, K. Differential effects of saturated and monounsaturated fatty acids on postprandial lipemia and incretin responses in healthy subjects. Am. J. Clin. Nutr. 1999, 69, 1135–1143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Rocca, A.S.; Brubaker, P.L. Stereospecific effects of fatty acids on proglucagon-derived peptide secretion in fetal rat intestinal cultures. Endocrinology 1995, 136, 5593–5599. [Google Scholar] [CrossRef]
  208. Kwasowski, P.; Flatt, P.R.; Bailey, C.J.; Marks, V. Effects of fatty acid chain length and saturation on gastric inhibitory polypeptide release in obese hyperglycaemic (ob/ob) mice. Biosci. Rep. 1985, 5, 701–705. [Google Scholar] [CrossRef]
  209. Feltrin, K.L.; Little, T.J.; Meyer, J.H.; Horowitz, M.; Smout, A.J.P.M.; Wishart, J.; Pilichiewicz, A.N.; Rades, T.; Chapman, I.M.; Feinle-Bisset, C. Effects of intraduodenal fatty acids on appetite, antropyloroduodenal motility, and plasma CCK and GLP-1 in humans vary with their chain length. Am. J. Physiol. Integr. Comp. Physiol. 2004, 287, R524–R533. [Google Scholar] [CrossRef] [Green Version]
  210. Maggio, C.A.; Koopmans, H.S. Food intake after intragastric meals of short-, medium-, or long-chain triglyceride. Physiol. Behav. 1982, 28, 921–926. [Google Scholar] [CrossRef]
  211. Pilichiewicz, A.; O’Donovan, D.; Feinle, C.; Lei, Y.; Wishart, J.M.; Bryant, L.; Meyer, J.H.; Horowitz, M.; Jones, K.L. Effect of lipase inhibition on gastric emptying of, and the glycemic and incretin responses to, an oil/aqueous drink in Type 2 diabetes mellitus. J. Clin. Endocrinol. Metab. 2003, 88, 3829–3834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Enç, F.Y.; Öneş, T.; Akın, H.L.; DeDe, F.; Turoğlu, H.T.; Ülfer, G.; Bekiroğlu, N.; Haklar, G.; Rehfeld, J.F.; Holst, J.J.; et al. Orlistat accelerates gastric emptying and attenuates GIP release in healthy subjects. Am. J. Physiol. Liver Physiol. 2009, 296, G482–G489. [Google Scholar] [CrossRef]
  213. Ellrichmann, M.; Kapelle, M.; Ritter, P.R.; Holst, J.J.; Herzig, K.-H.; Schmidt, W.E.; Schmitz, F.; Meier, J.J. Orlistat inhibition of intestinal lipase acutely increases appetite and attenuates postprandial glucagon-like peptide-1-(7–36)-Amide-1, cholecystokinin, and peptide YY concentrations. J. Clin. Endocrinol. Metab. 2008, 93, 3995–3998. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Beglinger, S.; Drewe, J.; Schirra, J.; Göke, B.; D’Amato, M.; Beglinger, C. Role of fat hydrolysis in regulating glucagon-like peptide-1 secretion. J. Clin. Endocrinol. Metab. 2010, 95, 879–886. [Google Scholar] [CrossRef]
  215. Rajalahti, T.; Lin, C.; Mjøs, S.A.; Kvalheim, O.M. Serum fatty acid and lipoprotein subclass concentrations and their associations in prepubertal healthy Norwegian children. Metabolomics 2016, 12, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Psichas, A.; Larraufie, P.F.; Goldspink, D.A.; Gribble, F.M.; Reimann, F. Chylomicrons stimulate incretin secretion in mouse and human cells. Diabetologia 2017, 60, 2475–2485. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Fatma, S.; Yakubov, R.; Anwar, K.; Hussain, M.M. Pluronic L81 enhances triacylglycerol accumulation in the cytosol and inhibits chylomicron secretion. J. Lipid Res. 2006, 47, 2422–2432. [Google Scholar] [CrossRef] [Green Version]
  218. Shimotoyodome, A.; Fukuoka, D.; Suzuki, J.; Fujii, Y.; Mizuno, T.; Meguro, S.; Tokimitsu, I.; Hase, T. Coingestion of acylglycerols differentially affects glucose-induced insulin secretion via glucose-dependent insulinotropic polypeptide in C57BL/6J mice. Endocrinology 2009, 150, 2118–2126. [Google Scholar] [CrossRef] [Green Version]
  219. Lu, W.J.; Yang, Q.; Yang, L.; Lee, D.; D’Alessio, D.; Tso, P. Chylomicron formation and secretion is required for lipid-stimulated release of incretins GLP-1 and GIP. Lipids 2012, 47, 571–580. [Google Scholar] [CrossRef] [Green Version]
  220. Okawa, M.; Fujii, K.; Ohbuchi, K.; Okumoto, M.; Aragane, K.; Sato, H.; Tamai, Y.; Seo, T.; Itoh, Y.; Yoshimoto, R. Role of MGAT2 and DGAT1 in the release of gut peptides after triglyceride ingestion. Biochem. Biophys. Res. Commun. 2009, 390, 377–381. [Google Scholar] [CrossRef]
  221. Liu, J.; McLaren, D.G.; Chen, D.; Kan, Y.; Stout, S.J.; Shen, X.; Murphy, B.A.; Forrest, G.; Karanam, B.; Sonatore, L.; et al. Potential mechanism of enhanced postprandial glucagon-like peptide-1 release following treatment with a diacylglycerol acyltransferase 1 inhibitor. Pharm. Res. Perspect. 2015, 3, e00193. [Google Scholar] [CrossRef] [Green Version]
  222. Sclafani, A.; Ackroff, K.; Schwartz, G.J. Selective effects of vagal deafferentation and celiac-superior mesenteric ganglionectomy on the reinforcing and satiating action of intestinal nutrients. Physiol. Behav. 2003, 78, 285–294. [Google Scholar] [CrossRef]
  223. Tamura, C.S.; Ritter, R.C. Intestinal capsaicin transiently attenuates suppression of sham feeding by oleate. Am. J. Physiol. Integr. Comp. Physiol. 1994, 267, R561–R568. [Google Scholar] [CrossRef]
  224. Greenberg, D.; Smith, G.P.; Gibbs, J. Intraduodenal infusions of fats elicit satiety in sham-feeding rats. Am. J. Physiol. Integr. Comp. Physiol. 1990, 259, R110–R118. [Google Scholar] [CrossRef] [PubMed]
  225. Kotarsky, K.; Nilsson, N.E.; Flodgren, E.; Owman, C.; Olde, B. A human cell surface receptor activated by free fatty acids and thiazolidinedione drugs. Biochem. Biophys. Res. Commun. 2003, 301, 406–410. [Google Scholar] [CrossRef]
  226. Briscoe, C.P.; Tadayyon, M.; Andrews, J.L.; Benson, W.G.; Chambers, J.K.; Eilert, M.M.; Ellis, C.; Elshourbagy, N.A.; Goetz, A.S.; Minnick, D.T.; et al. The orphan G protein-coupled receptor GPR40 is activated by medium and long chain fatty acids. J. Biol. Chem. 2003, 278, 11303–11311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Edfalk, S.; Steneberg, P.; Edlund, H. Gpr40 is expressed in enteroendocrine cells and mediates free fatty acid stimulation of incretin secretion. Diabetes 2008, 57, 2280–2287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Hauge, M.; Vestmar, M.A.; Husted, A.S.; Ekberg, J.P.; Wright, M.J.; Di Salvo, J.; Weinglass, A.B.; Engelstoft, M.S.; Madsen, A.N.; Lückmann, M.; et al. GPR40 (FFAR1)—Combined Gs and Gq signaling In Vitro is associated with robust incretin secretagogue action Ex Vivo and In Vivo. Mol. Metab. 2015, 4, 3–14. [Google Scholar] [CrossRef] [PubMed]
  229. Qian, J.; Gu, Y.; Wu, C.; Yu, F.; Chen, Y.; Zhu, J.; Yao, X.; Bei, C.; Zhu, Q. Agonist-induced activation of human FFA1 receptor signals to extracellular signal-regulated kinase 1 and 2 through Gq- and Gi-coupled signaling cascades. Cell. Mol. Biol. Lett. 2017, 22, 1–12. [Google Scholar] [CrossRef] [Green Version]
  230. Yamada, H.; Yoshida, M.; Ito, K.; Dezaki, K.; Yada, T.; Ishikawa, S.-E.; Kakei, M. Potentiation of glucose-stimulated insulin secretion by the GPR40–PLC–TRPC pathway in pancreatic β-Cells. Sci. Rep. 2016, 6, 25912. [Google Scholar] [CrossRef]
  231. Hirasawa, A.; Tsumaya, K.; Awaji, T.; Katsuma, S.; Adachi, T.; Yamada, M.; Sugimoto, Y.; Miyazaki, S.; Tsujimoto, G. Free fatty acids regulate gut incretin glucagon-like peptide-1 secretion through GPR120. Nat. Med. 2005, 11, 90–94. [Google Scholar] [CrossRef]
  232. Tanaka, T.; Yano, T.; Adachi, T.; Koshimizu, T.-A.; Hirasawa, A.; Tsujimoto, G. Cloning and characterization of the rat free fatty acid receptor GPR120: In Vivo effect of the natural ligand on GLP-1 secretion and proliferation of pancreatic β cells. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2008, 377, 515–522. [Google Scholar] [CrossRef] [PubMed]
  233. Iwasaki, K.; Harada, N.; Sasaki, K.; Yamane, S.; Iida, K.; Suzuki, K.; Hamasaki, A.; Nasteska, D.; Shibue, K.; Joo, E.; et al. Free fatty acid receptor GPR120 is highly expressed in enteroendocrine K cells of the upper small intestine and has a critical role in GIP secretion after fat ingestion. Endocrinology 2015, 156, 837–846. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Ekberg, J.H.; Hauge, M.; Kristensen, L.V.; Madsen, A.N.; Engelstoft, M.S.; Husted, A.-S.; Sichlau, R.; Egerod, K.L.; Timshel, P.; Kowalski, T.J.; et al. GPR119, a major enteroendocrine sensor of dietary triglyceride metabolites coacting in synergy with FFA1 (GPR40). Endocrinology 2016, 157, 4561–4569. [Google Scholar] [CrossRef] [PubMed]
  235. Sankoda, A.; Harada, N.; Iwasaki, K.; Yamane, S.; Murata, Y.; Shibue, K.; Thewjitcharoen, Y.; Suzuki, K.; Harada, T.; Kanemaru, Y.; et al. Long-chain free fatty acid receptor GPR120 mediates oil-induced GIP secretion through CCK in male mice. Endocrinology 2017, 158, 1172–1180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Engelstoft, M.S.; Park, W.-M.; Sakata, I.; Kristensen, L.V.; Husted, A.S.; Osborne-Lawrence, S.; Piper, P.K.; Walker, A.K.; Pedersen, M.H.; Nøhr, M.K.; et al. Seven transmembrane G protein-coupled receptor repertoire of gastric ghrelin cells. Mol. Metab. 2013, 2, 376–392. [Google Scholar] [CrossRef]
  237. Stone, V.M.; Dhayal, S.; Brocklehurst, K.J.; Lenaghan, C.; Winzell, M.S.; Hammar, M.; Xu, X.; Smith, D.M.; Morgan, N.G. GPR120 (FFAR4) is preferentially expressed in pancreatic delta cells and regulates somatostatin secretion from murine islets of Langerhans. Diabetologia 2014, 57, 1182–1191. [Google Scholar] [CrossRef] [Green Version]
  238. Iakoubov, R.; Izzo, A.; Yeung, A.; Whiteside, C.I.; Brubaker, P.L. Protein kinase Cζ is required for oleic acid-induced secretion of glucagon-like peptide-1 by intestinal endocrine L cells. Endocrinology 2007, 148, 1089–1098. [Google Scholar] [CrossRef] [Green Version]
  239. Shah, B.P.; Liu, P.; Yu, T.; Hansen, D.R.; Gilbertson, T.A. TRPM5 is critical for linoleic acid-induced CCK secretion from the enteroendocrine cell line, STC-1. Am. J. Physiol. Physiol. 2012, 302, C210–C219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  240. Lauffer, L.M.; Iakoubov, R.; Brubaker, P.L. GPR119 is essential for oleoylethanolamide-induced glucagon-like peptide-1 secretion from the intestinal enteroendocrine L-cell. Diabetes 2009, 58, 1058–1066. [Google Scholar] [CrossRef] [Green Version]
  241. Schwartz, G.J.; Fu, J.; Astarita, G.; Li, X.; Gaetani, S.; Campolongo, P.; Cuomo, V.; Piomelli, D. The lipid messenger OEA links dietary fat intake to satiety. Cell Metab. 2008, 8, 281–288. [Google Scholar] [CrossRef] [Green Version]
  242. Hansen, K.B.; Rosenkilde, M.M.; Knop, F.K.; Wellner, N.; Diep, T.A.; Rehfeld, J.F.; Andersen, U.B.; Holst, J.J.; Hansen, H.S. 2-oleoyl glycerol is a GPR119 agonist and signals GLP-1 release in humans. J. Clin. Endocrinol. Metab. 2011, 96, E1409–E1417. [Google Scholar] [CrossRef] [Green Version]
  243. Chu, Z.-L.; Carroll, C.; Chen, R.; Alfonso, J.; Gutierrez, V.; He, H.; Lucman, A.; Xing, C.; Sebring, K.; Zhou, J.; et al. N-oleoyldopamine enhances glucose homeostasis through the activation of GPR119. Mol. Endocrinology 2010, 24, 161–170. [Google Scholar] [CrossRef] [Green Version]
  244. Soga, T.; Ohishi, T.; Matsui, T.; Saito, T.; Matsumoto, M.; Takasaki, J.; Matsumoto, S.-I.; Kamohara, M.; Hiyama, H.; Yoshida, S.; et al. Lysophosphatidylcholine enhances glucose-dependent insulin secretion via an orphan G-protein-coupled receptor. Biochem. Biophys. Res. Commun. 2005, 326, 744–751. [Google Scholar] [CrossRef]
  245. Lan, H.; Vassileva, G.; Corona, A.; Liu, L.; Baker, H.; Golovko, A.; Abbondanzo, S.J.; Hu, W.; Yang, S.; Ning, Y.; et al. GPR119 is required for physiological regulation of glucagon-like peptide-1 secretion but not for metabolic homeostasis. J. Endocrinol. 2009, 201, 219–230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Fu, J.; Oveisi, F.; Gaetani, S.; Lin, E.; Piomelli, D. Oleoylethanolamide, an endogenous PPAR-α agonist, lowers body weight and hyperlipidemia in obese rats. Neuropharmacology 2005, 48, 1147–1153. [Google Scholar] [CrossRef]
  247. Fu, J.; Gaetani, S.; Oveisi, F.; Verme, J.L.; Serrano, A.; De Fonseca, F.R.; Rosengarth, A.; Luecke, H.; Di Giacomo, B.; Tarzia, G.; et al. Oleylethanolamide regulates feeding and body weight through activation of the nuclear receptor PPAR-α. Nat. Cell Biol. 2003, 425, 90–93. [Google Scholar] [CrossRef] [PubMed]
  248. Chu, Z.-L.; Carroll, C.; Alfonso, J.; Gutierrez, V.; He, H.; Lucman, A.; Pedraza, M.; Mondala, H.; Gao, H.; Bagnol, D.; et al. A role for intestinal endocrine cell-expressed G protein-coupled receptor 119 in glycemic control by enhancing glucagon-like peptide-1 and glucose-dependent insulinotropic peptide release. Endocrinology 2008, 149, 2038–2047. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  249. De Fonseca, F.R.; Navarro, M.; Gómez, R.; Escuredo, L.; Nava, F.; Fu, J.; Murillo-Rodríguez, E.; Giuffrida, A.; LoVerme, J.; Gaetani, S.; et al. An anorexic lipid mediator regulated by feeding. Nature 2001, 414, 209–212. [Google Scholar] [CrossRef] [Green Version]
  250. Fu, J.; Astarita, G.; Gaetani, S.; Kim, J.; Cravatt, B.F.; Mackie, K.; Piomelli, D. Food intake regulates oleoylethanolamide formation and degradation in the proximal small intestine. J. Biol. Chem. 2007, 282, 1518–1528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  251. Overton, H.A.; Babbs, A.J.; Doel, S.M.; Fyfe, M.C.; Gardner, L.S.; Griffin, G.; Jackson, H.C.; Procter, M.J.; Rasamison, C.M.; Tang-Christensen, M.; et al. Deorphanization of a G protein-coupled receptor for oleoylethanolamide and its use in the discovery of small-molecule hypophagic agents. Cell Metab. 2006, 3, 167–175. [Google Scholar] [CrossRef] [Green Version]
  252. Moss, C.E.; Glass, L.L.; Diakogiannaki, E.; Pais, R.; Lenaghan, C.; Smith, D.M.; Wedin, M.; Bohlooly, M.-Y.; Gribble, F.M.; Reimann, F. Lipid derivatives activate GPR119 and trigger GLP-1 secretion in primary murine L-cells. Peptides 2016, 77, 16–20. [Google Scholar] [CrossRef]
  253. Katz, L.B.; Gambale, J.J.; Rothenberg, P.L.; Vanapalli, S.R.; Vaccaro, N.; Xi, L.; Sarich, T.C.; Stein, P.P. Effects of JNJ-38431055, a novel GPR119 receptor agonist, in randomized, double-blind, placebo-controlled studies in subjects with Type 2 diabetes. Diabetes Obes. Metab. 2012, 14, 709–716. [Google Scholar] [CrossRef] [PubMed]
  254. Christensen, L.W.; Kuhre, R.E.; Janus, C.; Svendsen, B.; Holst, J.J. Vascular, but not luminal, activation of FFAR1 (GPR40) stimulates GLP-1 secretion from isolated perfused rat small intestine. Physiol. Rep. 2015, 3, e12551. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Tough, I.R.; Forbes, S.; Herzog, H.; Jones, R.M.; Schwartz, T.W.; Cox, H.M. Bidirectional GPR119 agonism requires peptide YY and glucose for activity in mouse and human colon mucosa. Endocrinology 2018, 159, 1704–1717. [Google Scholar] [CrossRef]
  256. Simons, P.J.; Kummer, J.A.; Luiken, J.J.; Boon, L. Apical CD36 immunolocalization in human and porcine taste buds from circumvallate and foliate papillae. Acta Histochem. 2011, 113, 839–843. [Google Scholar] [CrossRef] [PubMed]
  257. Fukuwatari, T.; Kawada, T.; Tsuruta, M.; Hiraoka, T.; Iwanaga, T.; Sugimoto, E.; Fushiki, T. Expression of the putative membrane fatty acid transporter (FAT) in taste buds of the circumvallate papillae in rats. FEBS Lett. 1997, 414, 461–464. [Google Scholar] [CrossRef] [PubMed]
  258. Chen, M.; Yang, Y.; Braunstein, E.; Georgeson, K.E.; Harmon, C.M. Gut expression and regulation of FAT/CD36: Possible role in fatty acid transport in rat enterocytes. Am. J. Physiol. Metab. 2001, 281, E916–E923. [Google Scholar] [CrossRef] [Green Version]
  259. Laugerette, F.; Passilly-Degrace, P.; Patris, B.; Niot, I.; Febbraio, M.; Montmayeur, J.-P.; Besnard, P. CD36 involvement in orosensory detection of dietary lipids, spontaneous fat preference, and digestive secretions. J. Clin. Investig. 2005, 115, 3177–3184. [Google Scholar] [CrossRef] [Green Version]
  260. Drover, V.A.; Nguyen, D.V.; Bastie, C.C.; Darlington, Y.F.; Abumrad, N.A.; Pessin, J.E.; London, E.; Sahoo, D.; Phillips, M.C. CD36 mediates both cellular uptake of very long chain fatty acids and their intestinal absorption in mice. J. Biol. Chem. 2008, 283, 13108–13115. [Google Scholar] [CrossRef] [Green Version]
  261. Nauli, A.M.; Nassir, F.; Zheng, S.; Yang, Q.; Lo, C.; Von Lehmden, S.B.; Lee, D.; Jandacek, R.J.; Abumrad, N.A.; Tso, P.; et al. CD36 is important for chylomicron formation and secretion and may mediate cholesterol uptake in the proximal intestine. Gastroenterology 2006, 131, 1197–1207. [Google Scholar] [CrossRef] [Green Version]
  262. Nassir, F.; Wilson, B.; Han, X.; Gross, R.W.; Abumrad, N.A. CD36 is important for fatty acid and cholesterol uptake by the proximal but not distal intestine. J. Biol. Chem. 2007, 282, 19493–19501. [Google Scholar] [CrossRef] [Green Version]
  263. Martin, C.; Passilly-Degrace, P.; Gaillard, D.; Merlin, J.-F.; Chevrot, M.; Besnard, P. The lipid-sensor candidates CD36 and GPR120 are differentially regulated by dietary lipids in mouse taste buds: Impact on spontaneous fat preference. PLoS ONE 2011, 6, e24014. [Google Scholar] [CrossRef] [Green Version]
  264. Maruyama, T.; Miyamoto, Y.; Nakamura, T.; Tamai, Y.; Okada, H.; Sugiyama, E.; Nakamura, T.; Itadani, H.; Tanaka, K. Identification of membrane-type receptor for bile acids (M-BAR). Biochem. Biophys. Res. Commun. 2002, 298, 714–719. [Google Scholar] [CrossRef]
  265. Kawamata, Y.; Fujii, R.; Hosoya, M.; Harada, M.; Yoshida, H.; Miwa, M.; Fukusumi, S.; Habata, Y.; Itoh, T.; Shintani, Y.; et al. A G protein-coupled receptor responsive to bile acids. J. Biol. Chem. 2003, 278, 9435–9440. [Google Scholar] [CrossRef] [Green Version]
  266. Makishima, M.; Okamoto, A.Y.; Repa, J.J.; Tu, H.; Learned, R.M.; Luk, A.; Hull, M.V.; Lustig, K.D.; Mangelsdorf, D.J.; Shan, B.; et al. Identification of a nuclear receptor for bile acids. Science 1999, 284, 1362–1365. [Google Scholar] [CrossRef]
  267. Namba, M.; Matsuyama, T.; Nonaka, K.; Tarui, S. Effect of intraluminal bile or bile acids on release of gut glucagon-like immunoreactive materials in the dog. Horm. Metab. Res. 1983, 15, 82–84. [Google Scholar] [CrossRef]
  268. Namba, M.; Matsuyama, T.; Itoh, H.; Imai, Y.; Horie, H.; Tarui, S. Inhibition of pentagastrin-stimulated gastric acid secretion by intraileal administration of bile and elevation of plasma concentrations of gut glucagon-like immunoreactivity in anesthetized dogs. Regul. Pept. 1986, 15, 121–128. [Google Scholar] [CrossRef]
  269. Thomas, C.; Gioiello, A.; Noriega, L.; Strehle, A.; Oury, J.; Rizzo, G.; Macchiarulo, A.; Yamamoto, H.; Mataki, C.; Pruzanski, M.; et al. TGR5-mediated bile acid sensing controls glucose homeostasis. Cell Metab. 2009, 10, 167–177. [Google Scholar] [CrossRef] [Green Version]
  270. Parker, H.E.; Wallis, K.; Le Roux, C.W.; Wong, K.Y.; Reimann, F.; Gribble, F.M. Molecular mechanisms underlying bile acid-stimulated glucagon-like peptide-1 secretion. Br. J. Pharm. 2011, 165, 414–423. [Google Scholar] [CrossRef] [Green Version]
  271. Katsuma, S.; Hirasawa, A.; Tsujimoto, G. Bile acids promote glucagon-like peptide-1 secretion through TGR5 in a murine enteroendocrine cell line STC-1. Biochem. Biophys. Res. Commun. 2005, 329, 386–390. [Google Scholar] [CrossRef]
  272. Wu, T.; Bound, M.J.; Standfield, S.D.; Gedulin, B.; Jones, K.L.; Horowitz, M.; Rayner, C.K. Effects of rectal administration of taurocholic acid on glucagon-like peptide-1 and peptide YY secretion in healthy humans. Diabetes Obes. Metab. 2012, 15, 474–477. [Google Scholar] [CrossRef]
  273. Adrian, T.E.; Ballantyne, G.H.; Longo, W.E.; Bilchik, A.J.; Graham, S.; Basson, M.D.; Tierney, R.P.; Modlin, I.M. Deoxycholate is an important releaser of peptide YY and enteroglucagon from the human colon. Gut 1993, 34, 1219–1224. [Google Scholar] [CrossRef] [Green Version]
  274. Kuhre, R.E.; Albrechtsen, N.J.W.; Larsen, O.; Jepsen, S.L.; Balk-Møller, E.; Andersen, D.B.; Deacon, C.F.; Schoonjans, K.; Reimann, F.; Gribble, F.M.; et al. Bile acids are important direct and indirect regulators of the secretion of appetite- and metabolism-regulating hormones from the gut and pancreas. Mol. Metab. 2018, 11, 84–95. [Google Scholar] [CrossRef]
  275. Burhol, P.G.; Lygren, I.; Waldum, H.; Jorde, R. The effect of duodenal infusion of bile on plasma VIP, GIP, and secretin and on duodenal bicarbonate secretion. Scand. J. Gastroenterol. 1980, 15, 1007–1011. [Google Scholar] [CrossRef] [PubMed]
  276. Lund, M.L.; Sorrentino, G.; Egerod, K.L.; Kroone, C.; Mortensen, B.; Knop, F.K.; Reimann, F.; Gribble, F.M.; Drucker, D.J.; De Koning, E.J.; et al. L-cell differentiation is induced by bile acids through GPBAR1 and paracrine GLP-1 and serotonin signaling. Diabetes 2020, 69, 614–623. [Google Scholar] [CrossRef]
  277. Christiansen, C.B.; Trammell, S.A.J.; Albrechtsen, N.J.W.; Schoonjans, K.; Albrechtsen, R.; Gillum, M.P.; Kuhre, R.E.; Holst, J.J. Bile acids drive colonic secretion of glucagon-like-peptide 1 and peptide-YY in rodents. Am. J. Physiol. Liver Physiol. 2019, 316, G574–G584. [Google Scholar] [CrossRef]
  278. Balakrishnan, A.; Wring, S.A.; Polli, J.E. Interaction of native bile acids with human apical sodium-dependent bile acid transporter (hASBT): Influence of steroidal hydroxylation pattern and C-24 conjugation. Pharm. Res. 2006, 23, 1451–1459. [Google Scholar] [CrossRef] [Green Version]
  279. Trabelsi, M.-S.; Daoudi, M.; Prawitt, J.; Ducastel, S.; Touche, V.; Sayin, S.I.; Perino, A.; Brighton, C.A.; Sebti, Y.; Kluza, J.; et al. Farnesoid X receptor inhibits glucagon-like peptide-1 production by enteroendocrine L cells. Nat. Commun. 2015, 6, 7629. [Google Scholar] [CrossRef] [Green Version]
  280. Poole, D.P.; Godfrey, C.; Cattaruzza, F.; Cottrell, G.S.; Kirkland, J.G.; Pelayo, J.C.; Bunnett, N.W.; Corvera, C.U. Expression and function of the bile acid receptor GpBAR1 (TGR5) in the murine enteric nervous system. Neurogastroenterol. Motil. 2010, 22, 814–e228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  281. Le Poul, E.; Loison, C.; Struyf, S.; Springael, J.-Y.; Lannoy, V.; Decobecq, M.-E.; Brezillon, S.; Dupriez, V.; Vassart, G.; Van Damme, J.; et al. Functional characterization of human receptors for short chain fatty acids and their role in polymorphonuclear cell activation. J. Biol. Chem. 2003, 278, 25481–25489. [Google Scholar] [CrossRef] [Green Version]
  282. Brown, A.J.; Goldsworthy, S.M.; Barnes, A.A.; Eilert, M.M.; Tcheang, L.; Daniels, D.; Muir, A.I.; Wigglesworth, M.J.; Kinghorn, I.; Fraser, N.J.; et al. The orphan G protein-coupled receptors GPR41 and GPR43 are activated by propionate and other short chain carboxylic acids. J. Biol. Chem. 2003, 278, 11312–11319. [Google Scholar] [CrossRef] [Green Version]
  283. Karaki, S.-I.; Mitsui, R.; Hayashi, H.; Kato, I.; Sugiya, H.; Iwanaga, T.; Furness, J.B.; Kuwahara, A. Short-chain fatty acid receptor, GPR43, is expressed by enteroendocrine cells and mucosal mast cells in rat intestine. Cell Tissue Res. 2006, 324, 353–360. [Google Scholar] [CrossRef] [PubMed]
  284. Karaki, S.-I.; Tazoe, H.; Hayashi, H.; Kashiwabara, H.; Tooyama, K.; Suzuki, Y.; Kuwahara, A. Expression of the short-chain fatty acid receptor, GPR43, in the human colon. J. Mol. Histol. 2008, 39, 135–142. [Google Scholar] [CrossRef] [PubMed]
  285. Fukumoto, S.; Tatewaki, M.; Yamada, T.; Fujimiya, M.; Mantyh, C.; Voss, M.; Eubanks, S.; Harris, M.; Pappas, T.N.; Takahashi, T.; et al. Short-chain fatty acids stimulate colonic transit via intraluminal 5-HT release in rats. Am. J. Physiol. Integr. Comp. Physiol. 2003, 284, R1269–R1276. [Google Scholar] [CrossRef] [Green Version]
  286. Nøhr, M.K.; Pedersen, M.H.; Gille, A.; Egerod, K.L.; Engelstoft, M.S.; Husted, A.S.; Sichlau, R.M.; Grunddal, K.V.; Poulsen, S.S.; Han, S.; et al. GPR41/FFAR3 and GPR43/FFAR2 as cosensors for short-chain fatty acids in enteroendocrine cells vs FFAR3 in enteric neurons and FFAR2 in enteric leukocytes. Endocrinology 2013, 154, 3552–3564. [Google Scholar] [CrossRef] [PubMed]
  287. Tolhurst, G.; Heffron, H.; Lam, Y.S.; Parker, H.E.; Habib, A.M.; Diakogiannaki, E.; Cameron, J.; Grosse, J.; Reimann, F.; Gribble, F.M.; et al. Short-chain fatty acids stimulate glucagon-like peptide-1 secretion via the g-protein-coupled receptor FFAR2. Diabetes 2011, 61, 364–371. [Google Scholar] [CrossRef] [Green Version]
  288. Nilsson, N.E.; Kotarsky, K.; Owman, C.; Olde, B. Identification of a free fatty acid receptor, FFA2R, expressed on leukocytes and activated by short-chain fatty acids. Biochem. Biophys. Res. Commun. 2003, 303, 1047–1052. [Google Scholar] [CrossRef]
  289. Tazoe, H.; Otomo, Y.; Karaki, S.-I.; Kato, I.; Fukami, Y.; Terasaki, M.; Kuwahara, A. Expression of short-chain fatty acid receptor GPR41 in the human colon. Biomed. Res. 2009, 30, 149–156. [Google Scholar] [CrossRef] [Green Version]
  290. Nøhr, M.; Egerod, K.; Christiansen, S.; Gille, A.; Offermanns, S.; Schwartz, T.; Møller, M. Expression of the short chain fatty acid receptor GPR41/FFAR3 in autonomic and somatic sensory ganglia. Neuroscience 2015, 290, 126–137. [Google Scholar] [CrossRef]
  291. Won, Y.-J.; Lu, V.B.; Puhl, H.L.; Ikeda, S.R. β-Hydroxybutyrate modulates N-type calcium channels in rat sympathetic neurons by acting as an agonist for the G-protein-coupled receptor FFA3. J. Neurosci. 2013, 33, 19314–19325. [Google Scholar] [CrossRef]
  292. Psichas, A.; Sleeth, M.L.; Murphy, K.G.; Brooks, L.J.; A Bewick, G.; Hanyaloglu, A.C.; A Ghatei, M.; Bloom, S.R.; Frost, G. The short chain fatty acid propionate stimulates GLP-1 and PYY secretion via free fatty acid receptor 2 in rodents. Int. J. Obes. 2015, 39, 424–429. [Google Scholar] [CrossRef] [Green Version]
  293. Forbes, S.; Stafford, S.; Coope, G.; Heffron, H.; Real, K.; Newman, R.; Davenport, R.; Barnes, M.; Grosse, J.; Cox, H.; et al. Selective FFA2 agonism appears to act via intestinal PYY to reduce transit and food intake but does not improve glucose tolerance in mouse models. Diabetes 2015, 64, 3763–3771. [Google Scholar] [CrossRef] [Green Version]
  294. Bolognini, D.; Moss, C.E.; Nilsson, K.; Petersson, A.U.; Donnelly, I.; Sergeev, E.; König, G.M.; Kostenis, E.; Kurowska-Stolarska, M.; Miller, A.; et al. A novel allosteric activator of free fatty acid 2 receptor displays unique gi-functional bias. J. Biol. Chem. 2016, 291, 18915–18931. [Google Scholar] [CrossRef] [Green Version]
  295. Christiansen, C.B.; Gabe, M.B.N.; Svendsen, B.; Dragsted, L.O.; Rosenkilde, M.M.; Holst, J.J. The impact of short-chain fatty acids on GLP-1 and PYY secretion from the isolated perfused rat colon. Am. J. Physiol. Liver Physiol. 2018, 315, G53–G65. [Google Scholar] [CrossRef] [Green Version]
  296. Müller, M.; Hernández, M.A.G.; Goossens, G.H.; Reijnders, D.; Holst, J.J.; Jocken, J.W.E.; Van Eijk, H.; Canfora, E.E.; Blaak, E.E. Circulating but not faecal short-chain fatty acids are related to insulin sensitivity, lipolysis and GLP-1 concentrations in humans. Sci. Rep. 2019, 9, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  297. Høverstad, T.; Bjørneklett, A.; Midtvedt, T.; Fausa, O.; Bøhmer, T. Short-chain fatty acids in the proximal gastrointestinal tract of healthy subjects. Scand J. Gastroenterol. 1984, 19, 1053–1058. [Google Scholar]
  298. Lee, E.-Y.; Zhang, X.; Miyamoto, J.; Kimura, I.; Taknaka, T.; Furusawa, K.; Jomori, T.; Fujimoto, K.; Uematsu, S.; Miki, T.; et al. Gut carbohydrate inhibits GIP secretion via a microbiota/SCFA/FFAR3 pathway. J. Endocrinol. 2018, 239, 267–276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  299. Zhou, J.; Martin, R.J.; Tulley, R.T.; Raggio, A.M.; McCutcheon, K.L.; Shen, L.; Danna, S.C.; Tripathy, S.; Hegsted, M.; Keenan, M.J.; et al. Dietary resistant starch upregulates total GLP-1 and PYY in a sustained day-long manner through fermentation in rodents. Am. J. Physiol. Metab. 2008, 295, E1160–E1166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  300. Larraufie, P.; Martin-Gallausiaux, C.; Lapaque, N.; Dore, J.; Gribble, F.M.; Reimann, F.; Blottiere, H.M. SCFAs strongly stimulate PYY production in human enteroendocrine cells. Sci. Rep. 2018, 8, 1–9. [Google Scholar] [CrossRef] [PubMed]
  301. Petersen, N.; Reimann, F.; Bartfeld, S.; Farin, H.F.; Ringnalda, F.C.; Vries, R.G.J.; Brink, S.V.D.; Clevers, H.; Gribble, F.M.; De Koning, E.J.P.; et al. Generation of L cells in mouse and human small intestine organoids. Diabetes 2014, 63, 410–420. [Google Scholar] [CrossRef] [Green Version]
  302. Arora, T.; Akrami, R.; Pais, R.; Bergqvist, L.; Johansson, B.R.; Schwartz, T.W.; Reimann, F.; Gribble, F.M.; Bäckhed, F. Microbial regulation of the L cell transcriptome. Sci. Rep. 2018, 8, 1207. [Google Scholar] [CrossRef]
  303. Wichmann, A.; Allahyar, A.; Greiner, T.U.; Plovier, H.; Östergren, G.L.; Larsson, T.; Drucker, D.J.; Delzenne, N.M.; Cani, P.; Bäckhed, F.; et al. Microbial modulation of energy availability in the colon regulates intestinal transit. Cell Host Microbe 2013, 14, 582–590. [Google Scholar] [CrossRef] [Green Version]
  304. De Fonseca, F.R.; Navarro, M.; Alvarez, E.; Roncero, I.; Chowen, J.A.; Maestre, O.; Gómez, R.; Muñoz, R.M.; Eng, J.; Blázquez, E.; et al. Peripheral versus central effects of glucagon-like peptide-1 receptor agonists on satiety and body weight loss in Zucker obese rats. Metabolism 2000, 49, 709–717. [Google Scholar] [CrossRef] [PubMed]
  305. Näslund, E.; Barkeling, B.; King, N.; Gutniak, M.; Blundell, J.; Holst, J.; Rössner, S.; Hellström, P. Energy intake and appetite are suppressed by glucagon-like peptide-1 (GLP-1) in obese men. Int. J. Obes. 1999, 23, 304–311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  306. Gutzwiller, J.-P.; Göke, B.; Drewe, J.; Hildebrand, P.; Ketterer, S.; Handschin, D.; Winterhalder, R.; Conen, D.; Beglinger, C. Glucagon-like peptide-1: A potent regulator of food intake in humans. Gut 1999, 44, 81–86. [Google Scholar] [CrossRef]
  307. Flint, A.; Raben, A.; Astrup, A.; Holst, J.J. Glucagon-like peptide 1 promotes satiety and suppresses energy intake in humans. J. Clin. Investig. 1998, 101, 515–520. [Google Scholar] [CrossRef] [PubMed]
  308. Kim, D.; MacConell, L.; Zhuang, N.; Kothare, P.A.; Trautmann, M.; Fineman, M.; Taylor, K. Effects of once-weekly dosing of a long-acting release formulation of exenatide on glucose control and body weight in subjects with Type 2 diabetes. Diabetes Care 2007, 30, 1487–1493. [Google Scholar] [CrossRef] [Green Version]
  309. Holst, J.J. Incretin hormones and the satiation signal. Int. J. Obes. 2013, 37, 1161–1168. [Google Scholar] [CrossRef] [Green Version]
  310. Astrup, A.; Rössner, S.; Van Gaal, L.; Rissanen, A.; Niskanen, L.; Al Hakim, M.; Madsen, J.; Rasmussen, M.F.; Lean, M.E. Effects of liraglutide in the treatment of obesity: A randomised, double-blind, placebo-controlled study. Lancet 2009, 374, 1606–1616. [Google Scholar] [CrossRef]
  311. Knudsen, L.B.; Lau, J. The discovery and development of liraglutide and semaglutide. Front. Endocrinol. 2019, 10, 155. [Google Scholar] [CrossRef] [Green Version]
  312. Finan, B.; Ma, T.; Ottaway, N.; Müller, T.D.; Habegger, K.M.; Heppner, K.M.; Kirchner, H.; Holland, J.; Hembree, J.; Raver, C.; et al. Unimolecular dual incretins maximize metabolic benefits in rodents, monkeys, and humans. Sci. Transl. Med. 2013, 5, 209. [Google Scholar] [CrossRef] [Green Version]
  313. Bergmann, N.C.; Lund, A.; Gasbjerg, L.S.; Meessen, E.C.E.; Andersen, M.M.; Bergmann, S.; Hartmann, B.; Holst, J.J.; Jessen, L.; Christensen, M.B.; et al. Effects of combined GIP and GLP-1 infusion on energy intake, appetite and energy expenditure in overweight/obese individuals: A randomised, crossover study. Diabetologia 2019, 62, 665–675. [Google Scholar] [CrossRef] [Green Version]
  314. Miyawaki, K.; Yamada, Y.; Ban, N.; Ihara, Y.; Tsukiyama, K.; Zhou, H.; Fujimoto, S.; Oku, A.; Tsuda, K.; Toyokuni, S.; et al. Inhibition of gastric inhibitory polypeptide signaling prevents obesity. Nat. Med. 2002, 8, 738–742. [Google Scholar] [CrossRef]
  315. Althage, M.C.; Ford, E.L.; Wang, S.; Tso, P.; Polonsky, K.S.; Wice, B.M. Targeted ablation of glucose-dependent insulinotropic polypeptide-producing cells in transgenic mice reduces obesity and insulin resistance induced by a high fat diet. J. Biol. Chem. 2008, 283, 18365–18376. [Google Scholar] [CrossRef] [Green Version]
  316. Pathak, V.; Gault, V.A.; Flatt, P.R.; Irwin, N. Antagonism of gastric inhibitory polypeptide (GIP) by palmitoylation of GIP analogues with N- and C-terminal modifications improves obesity and metabolic control in high fat fed mice. Mol. Cell. Endocrinol. 2015, 401, 120–129. [Google Scholar] [CrossRef] [PubMed]
  317. Nakamura, T.; Tanimoto, H.; Mizuno, Y.; Okamoto, M.; Takeuchi, M.; Tsubamoto, Y.; Noda, H. Gastric inhibitory polypeptide receptor antagonist, SKL-14959, suppressed body weight gain on diet-induced obesity mice. Obes. Sci. Pr. 2018, 4, 194–203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  318. Kim, S.-J.; Nian, C.; Karunakaran, S.; Clee, S.M.; Isales, C.M.; McIntosh, C.H.S. GIP-overexpressing mice demonstrate reduced diet-induced obesity and steatosis, and improved glucose homeostasis. PLoS ONE 2012, 7, e40156. [Google Scholar] [CrossRef]
  319. Geary, N.; Smith, G.P. Pancreatic glucagon and postprandial satiety in the rat. Physiol. Behav. 1982, 28, 313–322. [Google Scholar] [CrossRef]
  320. Schwartz, M.W.; Figlewicz, D.P.; Woods, S.C.; Porte, D.; Baskin, D.G. Insulin, neuropeptide Y, and food intake. Ann. N. Y. Acad. Sci. 1993, 692, 60–71. [Google Scholar] [CrossRef]
  321. Pamir, N.; Lynn, F.C.; Buchan, A.M.J.; Ehses, J.; Hinke, S.A.; Pospisilik, J.A.; Miyawaki, K.; Yamada, Y.; Seino, Y.; McIntosh, C.H.S.; et al. Glucose-dependent insulinotropic polypeptide receptor null mice exhibit compensatory changes in the enteroinsular axis. Am. J. Physiol. Metab. 2003, 284, E931–E939. [Google Scholar] [CrossRef] [Green Version]
  322. Sparre-Ulrich, A.H.; Hansen, L.S.; Svendsen, B.; Christensen, M.S.; Knop, F.K.; Hartmann, B.; Holst, J.J.; Rosenkilde, M.M. Species-specific action of (Pro3)GIP—A full agonist at human GIP receptors, but a partial agonist and competitive antagonist at rat and mouse GIP receptors. Br. J. Pharm. 2015, 173, 27–38. [Google Scholar] [CrossRef] [Green Version]
  323. Gault, V.A.; O’Harte, F.P.; Harriott, P.; Flatt, P.R. Characterization of the cellular and metabolic effects of a novel enzyme-resistant antagonist of glucose-dependent insulinotropic polypeptide. Biochem. Biophys. Res. Commun. 2002, 290, 1420–1426. [Google Scholar] [CrossRef] [PubMed]
  324. Mroz, P.A.; Finan, B.; Gelfanov, V.; Yang, B.; Tschöp, M.H.; DiMarchi, R.D.; Perez-Tilve, D. Optimized GIP analogs promote body weight lowering in mice through GIPR agonism not antagonism. Mol. Metab. 2019, 20, 51–62. [Google Scholar] [CrossRef] [PubMed]
  325. Nørregaard, P.K.; Deryabina, M.A.; Shelton, P.T.; Fog, J.U.; Daugaard, J.R.; Eriksson, P.; Larsen, L.F.; Jessen, L. A novel GIP analogue, ZP 4165, enhances glucagon-like peptide-1-induced body weight loss and improves glycaemic control in rodents. Diabetes Obes. Metab. 2018, 20, 60–68. [Google Scholar] [CrossRef]
  326. Gault, V.A.; Kerr, B.D.; Harriott, P.; Flatt, P.R. Administration of an acylated GLP-1 and GIP preparation provides added beneficial glucose-lowering and insulinotropic actions over single incretins in mice with Type 2 diabetes and obesity. Clin. Sci. 2011, 121, 107–117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  327. Portron, A.; Jadidi, S.; Sarkar, N.; Di Marchi, R.; Schmitt, C. Pharmacodynamics, pharmacokinetics, safety and tolerability of the novel dual glucose-dependent insulinotropic polypeptide/glucagon-like peptide-1 agonist RG7697 after single subcutaneous administration in healthy subjects. Diabetes Obes. Metab. 2017, 19, 1446–1453. [Google Scholar] [CrossRef]
  328. Schmitt, C.; Portron, A.; Jadidi, S.; Sarkar, N.; Di Marchi, R. Pharmacodynamics, pharmacokinetics and safety of multiple ascending doses of the novel dual glucose-dependent insulinotropic polypeptide/glucagon-like peptide-1 agonist RG7697 in people with Type 2 diabetes mellitus. Diabetes Obes. Metab. 2017, 19, 1436–1445. [Google Scholar] [CrossRef]
  329. Frias, J.P.; Bastyr, E.J.; Vignati, L.; Tschöp, M.H.; Schmitt, C.; Owen, K.; Christensen, R.H.; DiMarchi, R.D. The sustained effects of a dual GIP/GLP-1 receptor agonist, NNC0090–2746, in patients with Type 2 diabetes. Cell Metab. 2017, 26, 343–352.e2. [Google Scholar] [CrossRef]
  330. Frias, J.P.; Nauck, M.A.; Van, J.; Kutner, M.E.; Cui, X.; Benson, C.; Urva, S.; Gimeno, R.E.; Milicevic, Z.; Robins, D.; et al. Efficacy and safety of LY3298176, a novel dual GIP and GLP-1 receptor agonist, in patients with Type 2 diabetes: A randomised, placebo-controlled and active comparator-controlled phase 2 trial. Lancet 2018, 392, 2180–2193. [Google Scholar] [CrossRef]
  331. Coskun, T.; Sloop, K.W.; Loghin, C.; Alsina-Fernandez, J.; Urva, S.; Bokvist, K.B.; Cui, X.; Briere, D.A.; Cabrera, O.; Roell, W.C.; et al. LY3298176, a novel dual GIP and GLP-1 receptor agonist for the treatment of Type 2 diabetes mellitus: From discovery to clinical proof of concept. Mol. Metab. 2018, 18, 3–14. [Google Scholar] [CrossRef]
  332. Welch, I.; Saunders, K.; Read, N. Effect of ileal and intravenous infusions of fat emulsions on feeding and satiety in human volunteers. Gastroenterology 1985, 89, 1293–1297. [Google Scholar] [CrossRef]
  333. Lavin, J.H.; Wittert, G.; Sun, W.M.; Horowitz, M.; Morley, J.E.; Read, N.W. Appetite regulation by carbohydrate: Role of blood glucose and gastrointestinal hormones. Am. J. Physiol. Metab. 1996, 271, E209–E214. [Google Scholar] [CrossRef]
  334. Naslund, E.; Grybäck, P.; Bäckman, L.; Jacobsson, H.; Holst, J.J.; Theodorsson, H.E.; Hellstrom, P.M.; Juul, J. Distal small bowel hormones (Correlation with fasting antroduodenal motility and gastric emptying). Dig. Dis. Sci. 1998, 43, 945–952. [Google Scholar] [CrossRef]
  335. Imeryüz, N.; Yeğen, B.Ç.; Bozkurt, A.; Coşkun, T.; Villanueva-Peñacarrillo, M.L.; Ulusoy, N.B. Glucagon-like peptide-1 inhibits gastric emptying via vagal afferent-mediated central mechanisms. Am. J. Physiol. Liver Physiol. 1997, 273, G920–G927. [Google Scholar] [CrossRef]
  336. Schirra, J.; Nicolaus, M.; Woerle, H.J.; Struckmeier, C.; Katschinski, M.; Göke, B. GLP-1 regulates gastroduodenal motility involving cholinergic pathways. Neurogastroenterol. Motil. 2009, 21, 609–e22. [Google Scholar] [CrossRef]
  337. Richards, P.; Parker, H.E.; Adriaenssens, A.E.; Hodgson, J.M.; Cork, S.C.; Trapp, S.; Gribble, F.M.; Reimann, F. Identification and characterization of GLP-1 receptor-expressing cells using a new transgenic mouse model. Diabetes 2014, 63, 1224–1233. [Google Scholar] [CrossRef] [Green Version]
  338. Fitzgerald, P.C.E.; Manoliu, B.; Herbillon, B.; Steinert, R.E.; Horowitz, M.; Feinle-Bisset, C. Effects of L-phenylalanine on energy intake and glycaemia—impacts on appetite perceptions, gastrointestinal hormones and gastric emptying in healthy males. Nutrients 2020, 12, 1788. [Google Scholar] [CrossRef]
  339. Layer, P.; Holst, J.J.; Grandt, D.; Goebell, H. Ileal release of glucagon-like peptide-1 (GLP-1). Dig. Dis. Sci. 1995, 40, 1074–1082. [Google Scholar] [CrossRef]
  340. Wettergren, A.; Wojdemann, M.; Meisner, S.; Stadil, F.; Holst, J.J. The inhibitory effect of glucagon-like peptide-1 (GLP-1) 7–36 amide on gastric acid secretion in humans depends on an intact vagal innervation. Gut 1997, 40, 597–601. [Google Scholar] [CrossRef] [PubMed]
  341. Villar, H.V.; Fender, H.R.; Rayford, P.L.; Bloom, S.R.; Ramus, N.I.; Thompson, J.C. Suppression of gastrin release and gastric secretion by gastric inhibitory polypeptide (GIP) and vasoactive intestinal polypeptide (VIP). Ann. Surg. 1976, 184, 97–102. [Google Scholar] [CrossRef]
  342. Tolessa, T.; Gutniak, M.; Holst, J.J.; Efendic, S.; Hellström, P.M. Inhibitory effect of glucagon-like peptide-1 on small bowel motility. Fasting but not fed motility inhibited via nitric oxide independently of insulin and somatostatin. J. Clin. Investig. 1998, 102, 764–774. [Google Scholar] [CrossRef] [Green Version]
  343. Halim, A.; Degerblad, M.; Sundbom, M.; Karlbom, U.; Holst, J.J.; Webb, D.-L.; Hellström, P.M. Glucagon-like peptide-1 inhibits prandial gastrointestinal motility through myenteric neuronal mechanisms in humans. J. Clin. Endocrinol. Metab. 2017, 103, 575–585. [Google Scholar] [CrossRef] [Green Version]
  344. Ogawa, E.; Hosokawa, M.; Harada, N.; Yamane, S.; Hamasaki, A.; Toyoda, K.; Fujimoto, S.; Fujita, Y.; Fukuda, K.; Tsukiyama, K.; et al. The effect of gastric inhibitory polypeptide on intestinal glucose absorption and intestinal motility in mice. Biochem. Biophys. Res. Commun. 2011, 404, 115–120. [Google Scholar] [CrossRef] [Green Version]
  345. Edholm, T.; Cejvan, K.; Efendic, S.; Schmidt, P.T.; Hellström, P.M.; Abdel-Halim, S.M. The incretin hormones GIP and GLP-1 in diabetic rats: Effects on insulin secretion and small bowel motility. Neurogastroenterol. Motil. 2009, 21, 313–321. [Google Scholar] [CrossRef]
  346. Powley, T.L.; Phillips, R.J. Gastric satiation is volumetric, intestinal satiation is nutritive. Physiol. Behav. 2004, 82, 69–74. [Google Scholar] [CrossRef] [PubMed]
  347. Kupari, J.; Häring, M.; Agirre, E.; Castelo-Branco, G.; Ernfors, P. An atlas of vagal sensory neurons and their molecular specialization. Cell Rep. 2019, 27, 2508–2523.e4. [Google Scholar] [CrossRef] [Green Version]
  348. Egerod, K.L.; Petersen, N.; Timshel, P.N.; Rekling, J.C.; Wang, Y.; Liu, Q.; Schwartz, T.W.; Gautron, L. Profiling of G protein-coupled receptors in vagal afferents reveals novel gut-to-brain sensing mechanisms. Mol. Metab. 2018, 12, 62–75. [Google Scholar] [CrossRef] [PubMed]
  349. Bucinskaite, V.; Tolessa, T.; Pedersen, J.; Rydqvist, B.; Zerihun, L.; Holst, J.J.; Hellström, P.M. Receptor-mediated activation of gastric vagal afferents by glucagon-like peptide-1 in the rat. Neurogastroenterol. Motil. 2009, 21, e78. [Google Scholar] [CrossRef] [PubMed]
  350. Cork, S.C.; Richards, J.E.; Holt, M.K.; Gribble, F.M.; Reimann, F.; Trapp, S. Distribution and characterisation of Glucagon-like peptide-1 receptor expressing cells in the mouse brain. Mol. Metab. 2015, 4, 718–731. [Google Scholar] [CrossRef] [Green Version]
  351. Jensen, C.B.; Pyke, C.; Rasch, M.G.; Dahl, A.B.; Knudsen, L.B.; Secher, A. Characterization of the glucagonlike peptide-1 receptor in male mouse brain using a novel antibody and In Situ hybridization. Endocrinology 2017, 159, 665–675. [Google Scholar] [CrossRef] [Green Version]
  352. Ast, J.; Arvaniti, A.; Fine, N.H.F.; Nasteska, D.; Ashford, F.B.; Stamataki, Z.; Koszegi, Z.; Bacon, A.; Jones, B.J.; Lucey, M.A.; et al. Super-resolution microscopy compatible fluorescent probes reveal endogenous glucagon-like peptide-1 receptor distribution and dynamics. Nat. Commun. 2020, 11, 1–18. [Google Scholar] [CrossRef] [Green Version]
  353. Gabery, S.; Salinas, C.G.; Paulsen, S.J.; Ahnfelt-Rønne, J.; Alanentalo, T.; Baquero, A.F.; Buckley, S.T.; Farkas, E.; Fekete, C.; Frederiksen, K.S.; et al. Semaglutide lowers body weight in rodents via distributed neural pathways. JCI Insight 2020, 5. [Google Scholar] [CrossRef] [Green Version]
  354. Yox, D.P.; Stokesberry, H.; Ritter, R.C. Fourth ventricular capsaicin attenuates suppression of sham feeding induced by intestinal nutrients. Am. J. Physiol. Integr. Comp. Physiol. 1991, 260, R681–R687. [Google Scholar] [CrossRef]
  355. Yox, D.P.; Ritter, R.C. Capsaicin attenuates suppression of sham feeding induced by intestinal nutrients. Am. J. Physiol. Integr. Comp. Physiol. 1988, 255, R569–R574. [Google Scholar] [CrossRef] [PubMed]
  356. Larsen, P.; Tang-Christensen, M.; Holst, J.J.; Ørskov, C. Distribution of glucagon-like peptide-1 and other preproglucagon-derived peptides in the rat hypothalamus and brainstem. Neuroscience 1997, 77, 257–270. [Google Scholar] [CrossRef]
  357. Hisadome, K.; Reimann, F.; Gribble, F.M.; Trapp, S. Leptin directly depolarizes preproglucagon neurons in the nucleus tractus solitarius. Diabetes 2010, 59, 1890–1898. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  358. Brierley, D.I.; Holt, M.K.; Singh, A.; de Araujo, A.; Vergara, M.; Afaghani, M.H.; Lee, S.J.; Scott, K.; Langhans, W.; Krause, E.; et al. Central and peripheral GLP-1 systems independently and additively suppress eating. bioRxiv 2020. [Google Scholar] [CrossRef]
  359. Powley, T.L.; Spaulding, R.A.; Haglof, S.A. Vagal afferent innervation of the proximal gastrointestinal tract mucosa: Chemoreceptor and mechanoreceptor architecture. J. Comp. Neurol. 2010, 519, 644–660. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  360. Kaelberer, M.M.; Buchanan, K.L.; Klein, M.E.; Barth, B.B.; Montoya, M.M.; Shen, X.; Bohórquez, D.V. A gut-brain neural circuit for nutrient sensory transduction. Science 2018, 361, 5236. [Google Scholar] [CrossRef] [Green Version]
  361. Lu, V.B.; Rievaj, J.; O’Flaherty, E.A.; Smith, C.A.; Pais, R.; Pattison, L.A.; Tolhurst, G.; Leiter, A.B.; Bulmer, D.C.; Gribble, F.M.; et al. Adenosine triphosphate is co-secreted with glucagon-like peptide-1 to modulate intestinal enterocytes and afferent neurons. Nat. Commun. 2019, 10, 1029. [Google Scholar] [CrossRef]
  362. Williams, D.L.; Baskin, D.G.; Schwartz, M.W. Evidence that intestinal glucagon-like peptide-1 plays a physiological role in satiety. Endocrinology 2008, 150, 1680–1687. [Google Scholar] [CrossRef] [PubMed]
  363. Turton, M.D.; O’Shea, D.; Gunn, I.; Beak, S.A.; Edwards, C.M.B.; Meeran, K.; Choi, S.J.; Taylor, G.M.; Heath, M.M.; Lambert, P.D.; et al. A role for glucagon-like peptide-1 in the central regulation of feeding. Nat. Cell Biol. 1996, 379, 69–72. [Google Scholar] [CrossRef]
  364. Tang-Christensen, M.; Larsen, P.J.; Göke, R.; Fink-Jensen, A.; Jessop, D.S.; Möller, M.; Sheikh, S.P. Central administration of GLP-1-(7–36) amide inhibits food and water intake in rats. Am. J. Physiol. Integr. Comp. Physiol. 1996, 271, R848–R856. [Google Scholar] [CrossRef]
  365. Schick, R.R.; Zimmermann, J.P.; Walde, T.V.; Schusdziarra, V. Glucagon-like peptide 1-(7–36) amide acts at lateral and medial hypothalamic sites to suppress feeding in rats. Am. J. Physiol. Integr. Comp. Physiol. 2003, 284, R1427–R1435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  366. Baggio, L.L.; Huang, Q.; Brown, T.J.; Drucker, D.J. Oxyntomodulin and glucagon-like peptide-1 differentially regulate murine food intake and energy expenditure. Gastroenterology 2004, 127, 546–558. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  367. Namkoong, C.; Kim, M.S.; Jang, B.-T.; Lee, Y.H.; Cho, Y.-M.; Choi, H.J. Central administration of GLP-1 and GIP decreases feeding in mice. Biochem. Biophys. Res. Commun. 2017, 490, 247–252. [Google Scholar] [CrossRef] [PubMed]
  368. Adriaenssens, A.E.; Biggs, E.K.; Darwish, T.; Tadross, J.; Sukthankar, T.; Girish, M.; Polex-Wolf, J.; Lam, B.Y.; Zvetkova, I.; Pan, W.; et al. Glucose-dependent insulinotropic polypeptide receptor-expressing cells in the hypothalamus regulate food intake. Cell Metab. 2019, 30, 987–996.e6. [Google Scholar] [CrossRef] [Green Version]
  369. Holt, M.K.; Richards, J.E.; Cook, D.R.; Brierley, D.I.; Williams, D.L.; Reimann, F.; Gribble, F.M.; Trapp, S. Preproglucagon neurons in the nucleus of the solitary tract are the main source of brain GLP-1, mediate stress-induced hypophagia, and limit unusually large intakes of food. Diabetes 2019, 68, 21–33. [Google Scholar] [CrossRef] [Green Version]
  370. Mietlicki-Baase, E.G.; Ortinski, P.I.; Rupprecht, L.E.; Olivos, D.R.; Alhadeff, A.L.; Pierce, R.C.; Hayes, M.R. The food intake-suppressive effects of glucagon-like peptide-1 receptor signaling in the ventral tegmental area are mediated by AMPA/kainate receptors. Am. J. Physiol. Metab. 2013, 305, E1367–E1374. [Google Scholar] [CrossRef] [Green Version]
  371. Dickson, S.L.; Shirazi, R.H.; Hansson, C.; Bergquist, F.; Nissbrandt, H.; Skibicka, K.P. The Glucagon-like peptide 1 (GLP-1) analogue, exendin-4, decreases the rewarding value of food: A new role for mesolimbic GLP-1 receptors. J. Neurosci. 2012, 32, 4812–4820. [Google Scholar] [CrossRef]
  372. Dossat, A.M.; Diaz, R.; Gallo, L.; Panagos, A.; Kay, K.; Williams, D.L. Nucleus accumbens GLP-1 receptors influence meal size and palatability. Am. J. Physiol. Metab. 2013, 304, E1314. [Google Scholar] [CrossRef] [Green Version]
  373. Kentish, S.; Li, H.; Philp, L.K.; O’Donnell, T.A.; Isaacs, N.J.; Young, R.L.; Wittert, G.A.; Blackshaw, L.A.; Page, A.J. Diet-induced adaptation of vagal afferent function. J. Physiol. 2011, 590, 209–221. [Google Scholar] [CrossRef]
  374. Jepsen, S.L.; Grunddal, K.V.; Albrechtsen, N.J.W.; Engelstoft, M.S.; Gabe, M.B.N.; Jensen, E.P.; Ørskov, C.; Poulsen, S.S.; Rosenkilde, M.M.; Pedersen, J.; et al. Paracrine crosstalk between intestinal L- and D-cells controls secretion of glucagon-like peptide-1 in mice. Am. J. Physiol. Metab. 2019, 317, E1081–E1093. [Google Scholar] [CrossRef]
  375. Adriaenssens, A.; Lam, B.Y.H.; Billing, L.; Skeffington, K.; Sewing, S.; Reimann, F.; Gribble, F. A transcriptome-led exploration of molecular mechanisms regulating somatostatin-producing D-Cells in the gastric epithelium. Endocrinology 2015, 156, 3924–3936. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Overview of macronutrient digestion. (A) Schematic representation of the distribution of glucose-dependent insulinotropic polypeptide (GIP) expressing K-cells and glucagon-like peptide 1 (GLP-1) expressing L-cells along the longitudinal intestinal axis. (B) Schematic representation of major sites of macronutrient digestion and absorption for carbohydrates, proteins and fats along the longitudinal intestinal axis. The breakdown of macronutrients denoted above primary location of nutrient absorption. Whereas absorbed monosaccharides and amino acids are exported from the intestinal epithelium as such, the majority of free fatty acids and monoglycerides are stepwise re-synthesised within the epithelium by MGAT2 and DGAT1 into triglycerides, which together with other lipophilic substances are secreted as chylomicrons. The production of lipid metabolites, such as OEA and 2-monoacylglycerides, which are synthesized following absorption of dietary fats, is represented in light green. Conjugated bile acids released following fat detection in the proximal small intestine is deconjugated in the distal intestine by colonic gut bacteria, as indicated by a dark green bar. Few macronutrients escape absorption in the small intestine, but bacterial fermentation of “indigestible fibres” provides SCFA as another nutritional source in the large intestine. Abbreviations: DGAT1, diacylglyceride-acyltransferase-1; GIP, glucose-dependent insulinotropic polypeptide; GLP-1, glucagon-like peptide 1; MGAT2, monoacylglyceride-acyltransferase-2; OEA, oleoylethanolamide; SCFAs, short-chain fatty acids.
Figure 1. Overview of macronutrient digestion. (A) Schematic representation of the distribution of glucose-dependent insulinotropic polypeptide (GIP) expressing K-cells and glucagon-like peptide 1 (GLP-1) expressing L-cells along the longitudinal intestinal axis. (B) Schematic representation of major sites of macronutrient digestion and absorption for carbohydrates, proteins and fats along the longitudinal intestinal axis. The breakdown of macronutrients denoted above primary location of nutrient absorption. Whereas absorbed monosaccharides and amino acids are exported from the intestinal epithelium as such, the majority of free fatty acids and monoglycerides are stepwise re-synthesised within the epithelium by MGAT2 and DGAT1 into triglycerides, which together with other lipophilic substances are secreted as chylomicrons. The production of lipid metabolites, such as OEA and 2-monoacylglycerides, which are synthesized following absorption of dietary fats, is represented in light green. Conjugated bile acids released following fat detection in the proximal small intestine is deconjugated in the distal intestine by colonic gut bacteria, as indicated by a dark green bar. Few macronutrients escape absorption in the small intestine, but bacterial fermentation of “indigestible fibres” provides SCFA as another nutritional source in the large intestine. Abbreviations: DGAT1, diacylglyceride-acyltransferase-1; GIP, glucose-dependent insulinotropic polypeptide; GLP-1, glucagon-like peptide 1; MGAT2, monoacylglyceride-acyltransferase-2; OEA, oleoylethanolamide; SCFAs, short-chain fatty acids.
Nutrients 13 00883 g001
Figure 2. Carbohydrate-sensing mechanisms in incretin hormone secreting EECs. Schematic of EEC with apical process (top) and basolateral surface (bottom) populated with secretory vesicles containing incretin hormones. Enterocytes shown in grey beside EEC. Glucose sensing by incretin-secreting EECs is critically dependent on sodium coupled uptake via SGLT1. Other sensors have been described, but are controversial (labelled with a question mark) or appear to be of limited physiological relevance. When SGLT1 is inhibited, GLP-1 secretion is elevated at later time points, possibly downstream of fermentation to short-chain fatty acids, which target G-protein coupled fatty acid receptors FFA2 and FFA3. These have been depicted on the basolateral membrane in analogy to the location of FFA1 (see text for details). Transport mechanisms of SCFAs across the intestinal epithelium illustrated in enterocytes on the right. Abbreviations: ATP, adenosine triphosphate; FFA2 or 3, Free fatty acid receptors; GCK, glucokinase; GLUT2, glucose transporter 2; GLUT5, glucose transporter 5; IP3, inositol triphosphate; KATP, ATP sensitive potassium channel; MCT1 or 4/5, monocarboxylate transporters; PLCβ, phospholipase C beta; SCFAs, short-chain fatty acids; SGLT1, sodium-glucose linked transporter 1; SMCT1, sodium-coupled monocarboxylate transporter 1; STR, sweet taste receptor; TrpM5, transient receptor potential cation channel subfamily M member 5; ΔΨ, membrane depolarization.
Figure 2. Carbohydrate-sensing mechanisms in incretin hormone secreting EECs. Schematic of EEC with apical process (top) and basolateral surface (bottom) populated with secretory vesicles containing incretin hormones. Enterocytes shown in grey beside EEC. Glucose sensing by incretin-secreting EECs is critically dependent on sodium coupled uptake via SGLT1. Other sensors have been described, but are controversial (labelled with a question mark) or appear to be of limited physiological relevance. When SGLT1 is inhibited, GLP-1 secretion is elevated at later time points, possibly downstream of fermentation to short-chain fatty acids, which target G-protein coupled fatty acid receptors FFA2 and FFA3. These have been depicted on the basolateral membrane in analogy to the location of FFA1 (see text for details). Transport mechanisms of SCFAs across the intestinal epithelium illustrated in enterocytes on the right. Abbreviations: ATP, adenosine triphosphate; FFA2 or 3, Free fatty acid receptors; GCK, glucokinase; GLUT2, glucose transporter 2; GLUT5, glucose transporter 5; IP3, inositol triphosphate; KATP, ATP sensitive potassium channel; MCT1 or 4/5, monocarboxylate transporters; PLCβ, phospholipase C beta; SCFAs, short-chain fatty acids; SGLT1, sodium-glucose linked transporter 1; SMCT1, sodium-coupled monocarboxylate transporter 1; STR, sweet taste receptor; TrpM5, transient receptor potential cation channel subfamily M member 5; ΔΨ, membrane depolarization.
Nutrients 13 00883 g002
Figure 3. Protein-sensing mechanisms in incretin hormone secreting EECs. Schematic as in Fig2. Good evidence exists for a role of PEPT1, CaSR and GPR142 in incretin secretion, however, the exact locations of receptors and transporters are incompletely defined. Controversial contributors are marked with a question mark (see text for details). Abbreviations: AA, amino acid; B(0)AT-1, sodium-dependent neutral amino acid transporter; CaSR, calcium-sensing receptor; GPR142, G-protein coupled-receptor 142; GPRC6A, G protein-coupled receptor family C group 6 member A; IP3, inositol triphosphate; LPAR5, lysophosphatidic acid receptor 5; PEPT1, H+/peptide co-transporter; PLCβ, phospholipase C beta; TrpM5, transient receptor potential cation channel subfamily M member 5; UTR, umami taste receptor; ΔΨ, membrane depolarization.
Figure 3. Protein-sensing mechanisms in incretin hormone secreting EECs. Schematic as in Fig2. Good evidence exists for a role of PEPT1, CaSR and GPR142 in incretin secretion, however, the exact locations of receptors and transporters are incompletely defined. Controversial contributors are marked with a question mark (see text for details). Abbreviations: AA, amino acid; B(0)AT-1, sodium-dependent neutral amino acid transporter; CaSR, calcium-sensing receptor; GPR142, G-protein coupled-receptor 142; GPRC6A, G protein-coupled receptor family C group 6 member A; IP3, inositol triphosphate; LPAR5, lysophosphatidic acid receptor 5; PEPT1, H+/peptide co-transporter; PLCβ, phospholipase C beta; TrpM5, transient receptor potential cation channel subfamily M member 5; UTR, umami taste receptor; ΔΨ, membrane depolarization.
Nutrients 13 00883 g003
Figure 4. Fat-sensing mechanisms in incretin hormone secreting EECs. Lipid sensing appears dominated by activation of G-protein coupled receptors. Contrary to what had been thought originally, these do not directly sample the luminal contents, but are shielded behind lipid absorption by enterocytes, with the possible exception of the mono-acyl-glyceride sensor GPR119. Absorption of micelles and formation of chylomicrons through enterocytes is illustrated on the left. Bile acids, which assist in the emulsification and breakdown of fatty acids, also stimulate incretin hormone secretion and the mechanism of bile acid-sensing in EECs is illustrated in dark green on right side of EEC. Transport route for conjugated bile acids in the small intestine via ASBT, illustrated in enterocyte on the right. See text for details. Abbreviations: 2-OG, 2-oleoylglycerol; AC, adenylyl cyclase; ASBT, sodium-dependent bile acid transporter; cAMP, cyclic adenosine monophosphate; CD36, cluster of differentiation also known as fatty acid translocase; FA; fatty acid; FFA1, free fatty acid receptor 1; FFA4, free fatty acid receptor 4; GPBAR1, G-protein coupled bile acid receptor 1; GPR119, G-protein coupled receptor 119; IP3, inositol triphosphate; LCFA, long-chain fatty acid; MG; monoglycerides; OEA, oleoylethanolamide; PLCβ, phospholipase C beta; TrpC3, transient receptor potential cation channel subfamily C member 3; ΔΨ, membrane depolarization.
Figure 4. Fat-sensing mechanisms in incretin hormone secreting EECs. Lipid sensing appears dominated by activation of G-protein coupled receptors. Contrary to what had been thought originally, these do not directly sample the luminal contents, but are shielded behind lipid absorption by enterocytes, with the possible exception of the mono-acyl-glyceride sensor GPR119. Absorption of micelles and formation of chylomicrons through enterocytes is illustrated on the left. Bile acids, which assist in the emulsification and breakdown of fatty acids, also stimulate incretin hormone secretion and the mechanism of bile acid-sensing in EECs is illustrated in dark green on right side of EEC. Transport route for conjugated bile acids in the small intestine via ASBT, illustrated in enterocyte on the right. See text for details. Abbreviations: 2-OG, 2-oleoylglycerol; AC, adenylyl cyclase; ASBT, sodium-dependent bile acid transporter; cAMP, cyclic adenosine monophosphate; CD36, cluster of differentiation also known as fatty acid translocase; FA; fatty acid; FFA1, free fatty acid receptor 1; FFA4, free fatty acid receptor 4; GPBAR1, G-protein coupled bile acid receptor 1; GPR119, G-protein coupled receptor 119; IP3, inositol triphosphate; LCFA, long-chain fatty acid; MG; monoglycerides; OEA, oleoylethanolamide; PLCβ, phospholipase C beta; TrpC3, transient receptor potential cation channel subfamily C member 3; ΔΨ, membrane depolarization.
Nutrients 13 00883 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lu, V.B.; Gribble, F.M.; Reimann, F. Nutrient-Induced Cellular Mechanisms of Gut Hormone Secretion. Nutrients 2021, 13, 883. https://doi.org/10.3390/nu13030883

AMA Style

Lu VB, Gribble FM, Reimann F. Nutrient-Induced Cellular Mechanisms of Gut Hormone Secretion. Nutrients. 2021; 13(3):883. https://doi.org/10.3390/nu13030883

Chicago/Turabian Style

Lu, Van B., Fiona M. Gribble, and Frank Reimann. 2021. "Nutrient-Induced Cellular Mechanisms of Gut Hormone Secretion" Nutrients 13, no. 3: 883. https://doi.org/10.3390/nu13030883

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop