Next Article in Journal
Novel Insights into the Therapeutic Potential of Curcumin and Derivatives
Previous Article in Journal
Multi-Omics Data Integration Reveals Key Variables Contributing to Subgingival Microbiome Dysbiosis-Induced Inflammatory Response in a Hyperglycemic Microenvironment
Previous Article in Special Issue
miR-142 Targets TIM-1 in Human Endothelial Cells: Potential Implications for Stroke, COVID-19, Zika, Ebola, Dengue, and Other Viral Infections
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Intestinal Enteroendocrine Cells: Present and Future Druggable Targets

1
Department of Internal Medicine, Division of Gastroenterology and Hepatology, University of New Mexico, Albuquerque, NM 87131, USA
2
Department of Medicine, Division of Gastroenterology and Hepatology, Johns Hopkins University School of Medicine, Baltimore, MD 21205, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(10), 8836; https://doi.org/10.3390/ijms24108836
Submission received: 18 March 2023 / Revised: 3 May 2023 / Accepted: 12 May 2023 / Published: 16 May 2023
(This article belongs to the Special Issue State-of-the-Art Molecular Pharmacology in USA)

Abstract

:
Enteroendocrine cells are specialized secretory lineage cells in the small and large intestines that secrete hormones and peptides in response to luminal contents. The various hormones and peptides can act upon neighboring cells and as part of the endocrine system, circulate systemically via immune cells and the enteric nervous system. Locally, enteroendocrine cells have a major role in gastrointestinal motility, nutrient sensing, and glucose metabolism. Targeting the intestinal enteroendocrine cells or mimicking hormone secretion has been an important field of study in obesity and other metabolic diseases. Studies on the importance of these cells in inflammatory and auto-immune diseases have only recently been reported. The rapid global increase in metabolic and inflammatory diseases suggests that increased understanding and novel therapies are needed. This review will focus on the association between enteroendocrine changes and metabolic and inflammatory disease progression and conclude with the future of enteroendocrine cells as potential druggable targets.

1. Introduction

The intestinal epithelia have a unique role in human health, as the interface between the luminal contents of the ‘outside’ and the systemic responses from the ‘inside’. Various intestinal epithelial cells (IECs) work in concert to maintain communication and proper function between the lumen and non-epithelial cells beneath the basolateral domain. Intestinal stem cells at the base of the crypts and the rapidly cycling transit amplifying cells are responsible for the rapid renewal of the intestinal epithelium and give rise to sub-populations of differentiated cells within the epithelial lining [1]. These cell types include enterocytes, the most prominent cell type of the intestinal epithelium that is responsible for nutrient and water absorption [2], various secretory cells such as goblet cells that secrete mucins [3], enteroendocrine cells that secrete hormones [4], and Paneth cells that release antimicrobial factors to protect nearby stem cells at the base of small intestinal crypts [5]. There are also chemosensory tuft cells [6] that play a key role in defense against helminths, and M cells that function in the uptake and transport of antigens from the lumen to specialized antigen-presenting cells in Peyer’s patches and lymphoid follicles [7].
With recent scientific advances, our appreciation of the diverse functions of IECs has dramatically expanded. Prime examples of this complexity are intestinal enteroendocrine cells (EECs) that comprise at least eight cellular subtypes. These subtypes are classified based on the hormones and/or peptides they produce and their localization along the crypt–villus axis [8]. Enteroendocrine cells, the focus of this review, are derived from the secretory lineage and distributed throughout the length of the small bowel and colon. These key sensory cells comprise approximately 1% of the epithelium but collectively form the largest endocrine system in humans [9]. Although many EEC subtypes exist [10], they are most commonly characterized as chromogranin A (CHGA)-positive cells [11]. These specialized cells function as sensors of luminal contents that then responsively secrete hormones and peptides from the basolateral surface. Short-chain fatty acids, other microbial metabolites, small peptides, amino acids, glucose, and fructose bind to specific apical receptors and can be transported through the EECs [12]. The secreted hormones or peptides can recruit immune cells and/or directly interact with spinal and vagal afferent neurons of the enteric nervous system [13]. Due to the systemic influence of these specialized cells and their secreted products, they have become an attractive therapeutic target against various metabolic, inflammatory, and pathophysiological diseases. This review will focus on current research investigating enteroendocrine cells as present and future druggable targets.

2. Classification of Enteroendocrine Cells

Based on their individual structure and position in the gastrointestinal epithelia, EECs are divided into two groups: “open type”, with a bottleneck shape and apical prolongation with microvilli facing the lumen, or “closed type”, located close to the basal membrane, not reaching the lumen of the gut and lacking microvilli [14,15]. Open EECs detect luminal contents directly through microvilli interaction with luminal contents whereas the closed EECs are activated indirectly either through neural or humoral pathways. Closed EECs can form basolateral projections called neuropods; they directly synapse with nerves of the enteric nervous system [13]. Both types accumulate their secretory products in cytoplasmic granules and release them by exocytosis at the basolateral membrane upon mechanical, chemical, or neural stimulation [16]. Although the molecular mechanisms that control secretion have not been fully characterized, there is a reliance on intracellular calcium [17,18] and cyclic AMP [19]. In the EEC subtype, enterochromaffin (EC), serotonin (5-HT) secretion is controlled by increased expression of the intracellular calcium receptor, ryanodine receptor 3 [18]. It is not yet known if this signaling pathway is ubiquitous to all EEC subtypes for hormone and peptide secretion or specific to enterochromaffin cells.
There are at least eight described EEC subtypes; however, our current understanding of epithelial plasticity suggests that the subtypes have the ability to switch and secrete different hormones and peptides in response to luminal changes [20]. The current classification of EEC subtypes, their function, and their location in the digestive tract are listed in Table 1.

3. Enteroendocrine Cells and Inflammatory Bowel Disease (IBD)

In addition to nutrient absorption in the gut, the epithelia act as a barrier to commensal and pathogenic bacteria. There is a delicate balance in the crosstalk between the gut microbiota and the immune cells that reside in the lamina propria below the barrier [46]. Barrier dysregulation disrupts the balance between tolerance and immunity and can lead to chronic dysfunctions, such as inflammatory bowel disease [47]. IBD has been closely associated with altered gut microbiota resulting from an excessive inflammatory response to commensals or to pathogenic microbes. EECs express chemosensory receptors such as Toll-like receptors (TLR1, TLR2, and TLR4) that respond to intestinal bacterial TLR ligands [48,49] and G-protein-coupled receptors (GPR40, GPR41, GPR43, GPR119, and GPR120) that respond to gut-microbiota-derived short-chain fatty acids (SCFAs) or long-chain fatty acids (LCFAs) [50,51]. In response to the binding of microbial metabolites on these apical receptors, EECs responsively produce specific peptides, hormones, and cytokines that can act on immune cells [52]. This suggests that EECs have an important, yet not fully understood role in gut inflammation. EECs may be able to relay intestinal dysbiosis into physiological adaptation [4,53,54,55]. Known changes to EEC subtypes and the changes in secreted hormone status in inflammatory diseases are summarized in Table 2.

3.1. Enteroendocrine Cells in Ulcerative Colitis (UC) and Crohn’s Disease (CD)

Inflammation in UC is localized to the colon but can occur throughout the small and large intestines in CD. Enteroendocrine hormones play a crucial role in the pathophysiology of inflammatory bowel diseases, both UC and CD; they can modulate intestinal epithelial barrier function in a transcellular and paracellular manner [63]. There is evidence that EECs produce pro-inflammatory interleukin (IL)-17C during CD and UC with a potential role in disease progression [64]. The EEC subtypes, L cells and EC cells increase in numbers in lymphocytic colitis [65], while secretion of peptide tyrosine tyrosine (peptide YY) from L cells has been shown to decrease in both UC and CD [60,66]. By contrast, peptide YY and glucagon-like peptide 1 (GLP-1) secretion, both from L cells, increase in terminal ileal CD [61,67]. The contrasting and oftentimes contradictory natures of these EEC hormones and peptides are likely due to their broad range of functions—ranging from anti-inflammatory and regenerative to the promotion of dysbiosis. Peptide YY and GLP-1 are produced and secreted from the same EEC subtype yet can sometimes have contrasting secretion patterns in CD. This may indicate variables not yet studied in the context of EECs, such as the severity of inflammation and stage of disease progression.
EEC peptides have shown evidence of anti-inflammatory properties, with studies in animal models of IBD indicating that glucagon-like peptide 2 (GLP-2), from L cells, can have a protective role in disease progression [68]. Likewise, an anti-inflammatory role of GLP-1 and GLP-1 agonists has been reported in humans and mice [69,70]. Clinical trials are underway for the use of GLP-2 analogs (teduglutide and apraglutide) for the treatment of CD and short bowel syndrome [71,72], with promising results thus far. Teduglutide has been approved and is currently in use. It is a recombinant GLP-2 analog with an amino acid mutation that allows it to have a longer half-life than endogenous GLP-2 [73] by preventing cleavage from dipeptidyl peptidase IV (DPPIV). By contrast, apraglutide contains four amino acid mutations, which allows it to be more resistant to cleavage by DPPIV, and thus has a longer half-life than teduglutide [74].
Further studies are required to provide direct evidence of the crosstalk between EECs and the immune system in disease progression and severity. The essential and varied roles EECs play in IBD pathogenesis are becoming increasingly important. Overall, EECs are interesting therapeutic targets for IBD treatment and a potential alternative to current immunosuppressive therapies, possibly utilizing currently available drugs for treating metabolic syndromes [63].

3.2. Enteroendocrine Cells and Celiac Disease

Celiac disease is a small intestinal autoimmune disease triggered in a subset of genetically predisposed individuals by consumption of the indigestible portion of gluten, or gliadins [75]. Celiac disease manifests as a malabsorptive enteropathy and is associated with significant morbidity secondary to malabsorption. With the current lack of therapeutics, celiac diagnosis is a significant burden on the individual. Disease progression has been attributed to the activation of CD4+ T helper cells [76], but a recent study into this disease using intestinal epithelial organoids has found that the intestinal epithelial layer is an underestimated hotspot of celiac pathology [77].
Alterations in EECs and respective incretins have been reported in celiac disease [78,79]. The peptide hormone levels of cholecystokinin (CCK, from I cells), glucose-dependent insulinotropic polypeptide (GIP, from K cells), and GLP-1 (from L cells) are reduced in the blood of celiac disease patients. These peptides are thought to be the basis of pancreatic dysfunction that occurs during celiac disease progression [80]. By contrast, ghrelin from X/A-like cells [56] and other chromogranin A (CHGA)+ cells [59] increase in the duodenum of celiac disease patients and correlate with inflammation [81]. Additionally, plasma levels of 5-HT were significantly increased in celiac disease patients following consumption of a gluten-rich diet [82]. Increases in serum GLP-2 [57], somatostatin [58], plasma neurotensin [83], and oxyntomodulin [84] have also been reported in these patients.
EC cells may play a direct role in the pathogenesis of refractory celiac disease, a rare and severe form of celiac disease that is characterized by the persistence of celiac disease symptoms during the initial (up to 12 months) phase of a gluten elimination diet [85]. An increase in EC cells is thought to promote inflammation by increasing interferon (IFN)-gamma production [59]. The number of EC cells doubles in the duodenum of celiac disease patients despite villus atrophy, and the expression of the rate-limiting enzyme for 5-HT synthesis, tryptophan hydroxylase (TPH1) was significantly higher as well [59].

4. Enteroendocrine Cells and Metabolic Diseases

Metabolic diseases encompass disorders in which the natural metabolism and metabolic processes are disrupted. Diabetes and obesity are the most common metabolic diseases that are not necessarily inherited. There is increasing evidence that enteroendocrine hormones have important roles in the pathogenesis of metabolic diseases, specifically diabetes and obesity. These diseases are rising globally and greatly disrupt the quality of life. A better understanding of metabolic diseases that may originate from the gut, as well as improved therapeutics and preventative measures are needed to control these diseases. Known changes to specific subtypes of EECs and their respective hormones in diabetes and obesity are summarized in Table 3.

4.1. Role of Enteroendocrine Cells in Diabetes

Diabetes is a common yet still complicated disease. Disease progression can be simplified as insufficient insulin production from the pancreas or insufficient cellular response to insulin [91]. Interestingly, bariatric surgery (Roux-en-Y gastric bypass, laparoscopic gastric banding) can be used to manage diabetes, control hyperglycemia, and reduce incidence of developing diabetes in obese individuals [92]. The intestinal endocrine system may have a large role in diabetes management and progression.
Enteroendocrine K and L cells play a key role in regulating appetite and glucose homeostasis. GLP-1 potentiates postprandial insulin secretion, delays gastric emptying, and inhibits glucagon secretion [93]. In the small intestine of diabetic subjects, there was a notable decrease in the number of CHGA+ cells, but an increase in the expression of peptide YY and GCG, both secreted by L cells, in the colon [90]. The variable responses from EECs between the small and large intestine during metabolic disease states have yet to be fully understood.
A recent study has emphasized the role of GLP-1 in the pathogenesis of type II diabetes (T2D) in obese patients [89]. Using RNA sequencing, they reported a decrease in GLP-1 cell lineage, GLP-1 maturation from proglucagon, and a significant reduction of GLP-1 positive cells in jejunal samples of obese subjects with T2D compared to obese individuals who do not have T2D. GLP-1 acts on the pancreas to control insulin secretion and GLP-1-based therapies are now FDA (Food and Drug Administration)-approved for the treatment of T2D [94,95], however, patients with multiple metabolic diseases are likely to have a complicated EEC response to the available therapeutics.

4.2. Enteroendocrine Cells and Obesity

Obesity, defined as excessive fat accumulation and body mass index over 30, is a major factor in the development of metabolic syndrome. Lifestyle and dietary changes are important considerations, yet the rate of obesity in adults and children continues to rise. A better understanding of the molecular changes in the gut that arise from obesity is needed to develop better and more efficient therapeutics.
EECs contain nutrient receptors such as free fatty acid receptors (FFARs), calcium-sensing receptors (CaSRs), and taste receptors (TR) on their luminal surface which bind nutrients and trigger hormone release [96,97,98]. These nutrient-responsive hormones act on cells and neuronal pathways to modulate food intake, digestion, energy balance, and body weight. The potential importance of EECs is notably demonstrated in insulin resistance, diabetes, and obesity, which have all been associated with altered levels of EEC hormones [90,99]. Ghrelin (X/A-like cells) stimulates food intake and hunger sensation and decreases energy expenditure; this ultimately promotes weight gain [99], and is indicative of a positive correlation between obesity and ghrelin secretion. By contrast, obesity that is related to insulin resistance inhibits ghrelin secretion. In this specific metabolic disease, X/A-like cells express insulin receptors and insulin can directly suppress ghrelin secretion [100].
Cholecystokinin (CCK) from I cells promotes satiety and reduces food intake by suppressing hunger [101]; however, the role of CCK in obesity remains controversial. CCK-producing cells sense free fatty acids using fatty acid transporters FFAR1, FFAR4, and CD36; calcium-sensing receptors CaSR; lysophosphatidic acid receptor 5, LPAR5; and taste receptors TAS1R1/TAS1R3 [102]. While some studies have shown a delayed CCK response in obese patients following consumption of high-fat diets [103], there is also evidence that CCK release is elevated in obese subjects compared to lean individuals [86] or not changed at all [104]. The discordance in these findings may be due to differences in meal composition, study duration, and microbiota. Further exploration of these differences will open up the possibility of CCK-targeted enteroendocrine therapeutics that can promote satiety and reduce the burden of obesity.
GLP-1 has been shown to have a role in regulating glucose, food intake, and promoting weight loss [105]. GLP-1 stimulates insulin release following glucose absorption and inhibits glucagon release. Obesity has been associated with low fasting GLP-1 levels and decreased postprandial GLP-1 response under conditions of dysregulated glycemic control [87]. However, there is increased expression and secretion of GLP-1 in obese patients with normal glycemia. This may be in part due to a compensatory mechanism to prevent the development of T2D [106]. Several GLP-1 analogs and receptor agonists that are effective in the treatment of T2D are being tested for efficacy in treating obesity as well [107].
One of the functions of peptide YY includes the inhibition of glucose-stimulated insulin release. Peptide YY is primarily stimulated by lipids, rather than proteins or glucose [102]. It regulates weight through the suppression of caloric consumption and increased fat oxidation [108]. Lower baseline levels of peptide YY are associated with obesity and an overall higher body mass index [109]. Chronic administration of peptide YY in murine models of diet-induced obesity led to decreased body weight, food intake, and overall improved insulin sensitivity [88]. Currently, peptide YY analogs are being explored in clinical trials for the management of obesity and T2D treatment [110].

5. The Role of Enteroendocrine Cells in Gut Injury

Enteroendocrine cells can be dynamically altered during injury to the gut and other organs. There is evidence that implicates EECs as facultative stem cells that promote plasticity and regeneration [111,112], but they can also contribute to increases in small intestinal neuroendocrine tumors [113]. However, other studies have shown that active, mature EECs are needed for the release of endocrine hormones necessary in halting injury or promoting healing and regeneration. To date, the regenerative role of EECs has been best characterized by ghrelin secretion from X/A-like cells. In a murine model of radiation-induced injury, ghrelin treatment promoted epithelial healing and attenuated inflammation via activation of Notch signaling [114]. Ghrelin has also been found to attenuate intestinal ischemia-induced injury in mice by activating the mammalian target of rapamycin (mTOR) signaling [115], as well as decrease sepsis-induced lung injury by blocking AKT, inducible nitric oxide synthase (iNOS), and nuclear factor kappa B (NF-kB) signaling in alveolar macrophages [116]. Overall, increased ghrelin secretion has an anti-inflammatory effect on injured tissue.
GLP-1 is secreted in response to the presence of lipopolysaccharides (LPS) following intestinal barrier injury and protects against inflammation in mice [49]. However, this functional mechanism of regeneration via GLP-1 is not understood. Similarly, peripheral 5-HT circulates systemically by platelets and acts as a mitogen in injured organs. In murine models of liver injury, 5-HT bound to 5-HT2A and 2B receptors and promotes liver regeneration after partial hepatectomy [117]. Mice lacking TPH1 needed for 5-HT synthesis were unable to fully regenerate their liver. Although the connection between platelet-derived 5-HT, liver regeneration, and intestinal EECs was not explored in this study, these findings suggest that EECs can undergo adaptive reprogramming during injury and provides insights for potential new regenerative therapeutics.

6. Enteroendocrine Cells in the Gut–Brain Axis

The secretory products from EECs can reach distant targets through release in the bloodstream or act directly on nerve endings close to the site of release. The first report of EECs having paracrine functions was by Larsson et al., which described basal cytoplasmic prolongation in somatostatin-containing D cells [118]. Similar cytoplasmic prolongation processes have been reported in L cells containing peptide YY in addition to D cells [13,119]. Recent advances have suggested EEC sensory transmission by showing direct connections between EECs and the nervous system through which EECs can directly communicate with the neurons innervating the gastrointestinal tract to initiate gut–brain communication [120,121]. In this process, EECs contribute to the brain–gut axis by exerting regulatory effects between the brain and the gut. Vagal afferent terminals innervate the wall of the gastrointestinal tract and are close to the mucosal epithelium [122]. In response to peptides or hormones secreted by EECs, receptors located along the vagal afferent fibers convey stimuli generated by EECs to the brainstem [123].
Recent advances have described a pathogenic role of the microbiota–gut–brain axis in neurological and psychiatric disorders including Parkinson’s disease (PD) and schizophrenia, in which EECs might participate [124,125]. PD is a debilitating neurodegenerative disease characterized by motor disturbances, including resting tremors, rigidity, and slow movements, as well as gastrointestinal symptoms, such as constipation and gastroparesis [126]. It was recently discovered that α-synuclein (a major component of the pathological hallmarks of PD) is expressed in the EECs of mouse and human intestines [127]. Furthermore, α-synuclein-containing EECs directly connect to α-synuclein-containing nerves, forming a neural circuit between the gut and the nervous system. It is possible that toxins or other environmental influences in the gut lumen could affect α-synuclein folding in EECs, thereby initiating a process by which misfolded α-synuclein could propagate from the gut to the brain. Similarly, EECs are predicted to be involved in another major neurological disorder, schizophrenia. Importantly, serotonergic dysfunction is thought to contribute to the pathophysiology of schizophrenia [128,129]. An elevated level of prefrontal 5-HT1A receptors and a reduction in prefrontal 5-HT2A receptors have been observed in schizophrenia in a meta-analysis of post-mortem studies [130]. Since the largest reserve of 5-HT is produced in enterochromaffin cells [131] (Table 1), studies are underway to determine if systemically circulating 5-HT can influence neuronal receptors.
The EEC–neuron connection has opened an exciting new perspective on the complexity of the bidirectional communication between the brain and the gut. Targeting EEC receptors on the luminal side can potentially be used to activate hormonal and neuronal pathways providing a novel approach to treating neurological diseases. Furthermore, drugs changing EEC functions may participate in the control of depression, anxiety, and visceral hypersensitivity. 5-HT, a primary signaling neurotransmitter in CNS, gets altered in psychotic disorders, including schizophrenia, thus constituting a potential target of second-generation antipsychotics. Evidence suggests that activation of 5-HT receptors, specifically 5-HT3 and 5-HT4, sends signals from the gut to the CNS, which activates multiple pathways including those involved in nociception [132]. Recently, a significant amount of attention has been given to EECs and their role in visceral pain perception [133,134]. Increased understanding of the mechanistic role that EECs play in visceral pain may lead to novel non-addictive therapeutics that target specific EEC receptors.

7. Current Therapeutics Targeting Enteroendocrine Hormones and Peptides

Short bowel syndrome (SBS) is a chronic malabsorption disorder in which the body is unable to absorb enough nutrients from food or drinks, generally from surgical removal of portions of the intestine [135]. GLP-2 analogs that target the GLP-2 receptor are currently used in the clinic to successfully treat SBS. These analogs are recombinant GLP-2 with amino acid mutations to prevent cleavage by DPPIV. GLP-2 administration in SBS patients greatly increases intestinal absorption of nutrients and water, reduces gastric emptying, and increases body weight and lean body mass [136]. These trophic effects of GLP-2 are likely due to stimulation of enterocyte proliferation and decreased enterocyte apoptosis [136,137], although further mechanistic studies are needed. Recent studies have conclusively found that semaglutide (a GLP-1 receptor agonist) mimics the action of GLP-1 by stimulating insulin release after a meal, prolonging satiety, and reducing glucose production by the liver [95]. Due to its mechanism of action on insulin release, it has been approved for the treatment of T2D. Additionally, GLP-1 receptor agonists are showing promising results in Phase 2 and 3 trials for the treatment of obesity [138,139]. One formulation of semaglutide, Wegovy, has recently been FDA-approved for treating obesity in adults and children over 12 years of age [140]. Although GLP-1 receptor agonists are useful for treating obesity, how will long-term use of these formulations affect the metabolic status of the intestinal epithelia and immune cells? As noted in Table 2 and Table 3, GLP-1 and 2 are often found with abnormal, increased secretion patterns in a variety of metabolic and inflammatory diseases. Therefore, further mechanistic studies on the systemic and long-term effects of GLP-1 receptor agonists are still needed. A potentially less physiologically disruptive approach to obesity treatment is supplementation with medium-chain fatty acids (MCFA) or their agonists. A recent study used MCFA agonists to activate MCFA colonic receptors, GPR84 and FFAR4, in obese adults [141]. These subjects experienced increased satiety and had higher levels of peptide YY but did not have significantly altered levels of GLP-1. The complexity of these hormones and peptides, particularly those secreted from the same EEC subtype, needs further exploration in this second wave of therapeutics for metabolic and inflammatory diseases.

8. Conclusions

Enteroendocrine cells are being recognized for their roles in various metabolic, inflammatory, and autoimmune diseases. However, the secreted hormones and peptides ofttimes have contradictory—beneficial and detrimental—roles in these diseases. More studies are needed to understand the exact mechanistic role of these specialized cells in driving disease states or influencing healing and regeneration. Future therapeutics may target the EECs in attempts to manipulate the microbiota or neurological pathways. The interaction between EECs, immune cells, and enteric nerves highlights their importance in the systemic spread of diseases that may originate in the gut. Thus, enteroendocrine cells and their hormones and peptides are promising therapeutic targets for diseases described in this review.

Author Contributions

Conceptualization—J.G.I.; writing, original draft preparation—R.A.; writing, review and editing—V.S. and J.G.I.; supervision—V.S. and J.G.I.; funding acquisition—V.S. and J.G.I. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by NIH grants K01DK106323 (JGI), R56ES034400 (JGI), Cystic Fibrosis Foundation grant CF-SINGH20G0 (VS), and American Cancer Society Institutional Research Grant IRG-14-187-19 (JGI).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

All authors have declared no conflict of interest.

References

  1. Takahashi, T.; Shiraishi, A. Stem Cell Signaling Pathways in the Small Intestine. Int. J. Mol. Sci. 2020, 21, 2032. [Google Scholar] [CrossRef]
  2. Miron, N.; Cristea, V. Enterocytes: Active cells in tolerance to food and microbial antigens in the gut. Clin. Exp. Immunol. 2012, 167, 405–412. [Google Scholar] [CrossRef] [PubMed]
  3. Kim, Y.S.; Ho, S.B. Intestinal goblet cells and mucins in health and disease: Recent insights and progress. Curr. Gastroenterol. Rep. 2010, 12, 319–330. [Google Scholar] [CrossRef] [PubMed]
  4. Gribble, F.M.; Reimann, F. Enteroendocrine Cells: Chemosensors in the Intestinal Epithelium. Annu. Rev. Physiol. 2016, 78, 277–299. [Google Scholar] [CrossRef] [PubMed]
  5. Gassler, N. Paneth cells in intestinal physiology and pathophysiology. World J. Gastrointest. Pathophysiol. 2017, 8, 150–160. [Google Scholar] [CrossRef]
  6. Hoover, B.; Baena, V.; Kaelberer, M.M.; Getaneh, F.; Chinchilla, S.; Bohorquez, D.V. The intestinal tuft cell nanostructure in 3D. Sci. Rep. 2017, 7, 1652. [Google Scholar] [CrossRef]
  7. Nicoletti, C. Unsolved mysteries of intestinal M cells. Gut 2000, 47, 735–739. [Google Scholar] [CrossRef] [PubMed]
  8. Basak, O.; Beumer, J.; Wiebrands, K.; Seno, H.; van Oudenaarden, A.; Clevers, H. Induced Quiescence of Lgr5+ Stem Cells in Intestinal Organoids Enables Differentiation of Hormone-Producing Enteroendocrine Cells. Cell Stem Cell 2017, 20, 177–190.e4. [Google Scholar] [CrossRef]
  9. Moran, G.W.; Leslie, F.C.; Levison, S.E.; Worthington, J.; McLaughlin, J.T. Enteroendocrine cells: Neglected players in gastrointestinal disorders? Therap. Adv. Gastroenterol. 2008, 1, 51–60. [Google Scholar] [CrossRef]
  10. Gunawardene, A.R.; Corfe, B.M.; Staton, C.A. Classification and functions of enteroendocrine cells of the lower gastrointestinal tract. Int. J. Exp. Pathol. 2011, 92, 219–231. [Google Scholar] [CrossRef]
  11. Kim, T.; Tao-Cheng, J.H.; Eiden, L.E.; Loh, Y.P. Chromogranin A, an “on/off” switch controlling dense-core secretory granule biogenesis. Cell 2001, 106, 499–509. [Google Scholar] [CrossRef] [PubMed]
  12. Duca, F.A.; Waise, T.M.Z.; Peppler, W.T.; Lam, T.K.T. The metabolic impact of small intestinal nutrient sensing. Nat. Commun. 2021, 12, 903. [Google Scholar] [CrossRef] [PubMed]
  13. Bohorquez, 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. 2011, 42, 3–13. [Google Scholar] [CrossRef] [PubMed]
  14. Bai, L.; Sivakumar, N.; Yu, S.; Mesgarzadeh, S.; Ding, T.; Ly, T.; Corpuz, T.V.; Grove, J.C.; Jarvie, B.C.; Knight, Z.A. Enteroendocrine cell types that drive food reward and aversion. Elife 2022, 11, e74964. [Google Scholar] [CrossRef]
  15. Hayashi, M.; Kaye, J.A.; Douglas, E.R.; Joshi, N.R.; Gribble, F.M.; Reimann, F.; Liberles, S.D. Enteroendocrine cell lineages that differentially control feeding and gut motility. Elife 2023, 12, e78512. [Google Scholar] [CrossRef] [PubMed]
  16. Ohara-Imaizumi, M.; Aoyagi, K.; Akimoto, Y.; Nakamichi, Y.; Nishiwaki, C.; Kawakami, H.; Nagamatsu, S. Imaging exocytosis of single glucagon-like peptide-1 containing granules in a murine enteroendocrine cell line with total internal reflection fluorescent microscopy. Biochem. Biophys. Res. Commun. 2009, 390, 16–20. [Google Scholar] [CrossRef]
  17. Bellono, N.W.; Bayrer, J.R.; Leitch, D.B.; Castro, J.; Zhang, C.; O’Donnell, T.A.; Brierley, S.M.; Ingraham, H.A.; Julius, D. Enterochromaffin Cells Are Gut Chemosensors that Couple to Sensory Neural Pathways. Cell 2017, 170, 185–198.e16. [Google Scholar] [CrossRef]
  18. Knutson, K.R.; Whiteman, S.T.; Alcaino, C.; Mercado-Perez, A.; Finholm, I.; Serlin, H.K.; Bellampalli, S.S.; Linden, D.R.; Farrugia, G.; Beyder, A. Intestinal enteroendocrine cells rely on ryanodine and IP(3) calcium store receptors for mechanotransduction. J. Physiol. 2023, 601, 287–305. [Google Scholar] [CrossRef]
  19. Simpson, A.K.; Ward, P.S.; Wong, K.Y.; Collord, G.J.; Habib, A.M.; Reimann, F.; Gribble, F.M. Cyclic AMP triggers glucagon-like peptide-1 secretion from the GLUTag enteroendocrine cell line. Diabetologia 2007, 50, 2181–2189. [Google Scholar] [CrossRef]
  20. Beumer, J.; Gehart, H.; Clevers, H. Enteroendocrine Dynamics—New Tools Reveal Hormonal Plasticity in the Gut. Endocr. Rev. 2020, 41, bnaa018. [Google Scholar] [CrossRef]
  21. Kanova, M.; Kohout, P. Serotonin-Its Synthesis and Roles in the Healthy and the Critically Ill. Int. J. Mol. Sci. 2021, 22, 4837. [Google Scholar] [CrossRef] [PubMed]
  22. Pithadia, A. 5-Hydroxytryptamine Receptor Subtypes and their Modulators with Therapeutic Potentials. J. Clin. Med. Res. 2009, 1, 72–80. [Google Scholar] [CrossRef]
  23. Shajib, M.S.; Khan, W.I. The role of serotonin and its receptors in activation of immune responses and inflammation. Acta Physiol. 2015, 213, 561–574. [Google Scholar] [CrossRef]
  24. Drucker, D.J. Mechanisms of Action and Therapeutic Application of Glucagon-like Peptide-1. Cell Metab. 2018, 27, 740–756. [Google Scholar] [CrossRef]
  25. Cinci, L.; Faussone-Pellegrini, M.S.; Rotondo, A.; Mule, F.; Vannucchi, M.G. GLP-2 receptor expression in excitatory and inhibitory enteric neurons and its role in mouse duodenum contractility. Neurogastroenterol. Motil. 2011, 23, e383–e392. [Google Scholar] [CrossRef]
  26. Guan, X.; Shi, X.; Li, X.; Chang, B.; Wang, Y.; Li, D.; Chan, L. GLP-2 receptor in POMC neurons suppresses feeding behavior and gastric motility. Am. J. Physiol. Endocrinol. Metab. 2012, 303, E853–E864. [Google Scholar] [CrossRef] [PubMed]
  27. Albarran-Zeckler, R.G.; Smith, R.G. The ghrelin receptors (GHS-R1a and GHS-R1b). Endocr. Dev. 2013, 25, 5–15. [Google Scholar] [CrossRef] [PubMed]
  28. Yu, R.; Li, Z.; Liu, S.; Huwatibieke, B.; Li, Y.; Yin, Y.; Zhang, W. Activation of mTORC1 signaling in gastric X/A-like cells induces spontaneous pancreatic fibrosis and derangement of glucose metabolism by reducing ghrelin production. EBioMedicine 2018, 36, 304–315. [Google Scholar] [CrossRef]
  29. Overton, H.A.; Fyfe, M.C.; Reynet, C. GPR119, a novel G protein-coupled receptor target for the treatment of type 2 diabetes and obesity. Br. J. Pharmacol. 2008, 153 (Suppl. 1), S76–S81. [Google Scholar] [CrossRef]
  30. Martin, B.; Lopez de Maturana, R.; Brenneman, R.; Walent, T.; Mattson, M.P.; Maudsley, S. Class II G protein-coupled receptors and their ligands in neuronal function and protection. Neuromol. Med. 2005, 7, 3–36. [Google Scholar] [CrossRef]
  31. Szewczyk, J.R.; Laudeman, C. CCK1R agonists: A promising target for the pharmacological treatment of obesity. Curr. Top. Med. Chem. 2003, 3, 837–854. [Google Scholar] [CrossRef]
  32. Tanaka, T.; Katsuma, S.; Adachi, T.; Koshimizu, T.A.; Hirasawa, A.; Tsujimoto, G. Free fatty acids induce cholecystokinin secretion through GPR120. Naunyn Schmiedeberg’s Arch. Pharmacol. 2008, 377, 523–527. [Google Scholar] [CrossRef]
  33. Theodoropoulou, M.; Stalla, G.K. Somatostatin receptors: From signaling to clinical practice. Front. Neuroendocrinol. 2013, 34, 228–252. [Google Scholar] [CrossRef]
  34. Harda, K.; Szabo, Z.; Juhasz, E.; Dezso, B.; Kiss, C.; Schally, A.V.; Halmos, G. Expression of Somatostatin Receptor Subtypes (SSTR-1-SSTR-5) in Pediatric Hematological and Oncological Disorders. Molecules 2020, 25, 5775. [Google Scholar] [CrossRef] [PubMed]
  35. Schmassmann, A.; Reubi, J.C. Cholecystokinin-B/gastrin receptors enhance wound healing in the rat gastric mucosa. J. Clin. Investig. 2000, 106, 1021–1029. [Google Scholar] [CrossRef] [PubMed]
  36. Larsson, L.I. Developmental biology of gastrin and somatostatin cells in the antropyloric mucosa of the stomach. Microsc. Res. Tech. 2000, 48, 272–281. [Google Scholar] [CrossRef]
  37. Jiang, X.; Liu, Y.; Zhang, X.Y.; Liu, X.; Liu, X.; Wu, X.; Jose, P.A.; Duan, S.; Xu, F.J.; Yang, Z. Intestinal Gastrin/CCKBR (Cholecystokinin B Receptor) Ameliorates Salt-Sensitive Hypertension by Inhibiting Intestinal Na(+)/H(+) Exchanger 3 Activity Through a PKC (Protein Kinase C)-Mediated NHERF1 and NHERF2 Pathway. Hypertension 2022, 79, 1668–1679. [Google Scholar] [CrossRef] [PubMed]
  38. Xiao, Y.; Yan, W.; Lu, Y.; Zhou, K.; Cai, W. Neurotensin contributes to pediatric intestinal failure-associated liver disease via regulating intestinal bile acids uptake. EBioMedicine 2018, 35, 133–141. [Google Scholar] [CrossRef]
  39. Li, J.; Song, J.; Yan, B.; Weiss, H.L.; Weiss, L.T.; Gao, T.; Evers, B.M. Neurotensin differentially regulates bile acid metabolism and intestinal FXR-bile acid transporter axis in response to nutrient abundance. FASEB J. 2021, 35, e21371. [Google Scholar] [CrossRef]
  40. Takeshita, E.; Matsuura, B.; Dong, M.; Miller, L.J.; Matsui, H.; Onji, M. Molecular characterization and distribution of motilin family receptors in the human gastrointestinal tract. J. Gastroenterol. 2006, 41, 223–230. [Google Scholar] [CrossRef]
  41. Miedzybrodzka, E.L.; Foreman, R.E.; Lu, V.B.; George, A.L.; Smith, C.A.; Larraufie, P.; Kay, R.G.; Goldspink, D.A.; Reimann, F.; Gribble, F.M. Stimulation of motilin secretion by bile, free fatty acids, and acidification in human duodenal organoids. Mol. Metab. 2021, 54, 101356. [Google Scholar] [CrossRef] [PubMed]
  42. Modvig, I.M.; Andersen, D.B.; Grunddal, K.V.; Kuhre, R.E.; Martinussen, C.; Christiansen, C.B.; Orskov, C.; Larraufie, P.; Kay, R.G.; Reimann, F.; et al. Secretin release after Roux-en-Y gastric bypass reveals a population of glucose-sensitive S cells in distal small intestine. Int. J. Obes. 2020, 44, 1859–1871. [Google Scholar] [CrossRef] [PubMed]
  43. Fukuhara, S.; Kobayashi, K.; Kusakizako, T.; Iida, W.; Kato, M.; Shihoya, W.; Nureki, O. Structure of the human secretin receptor coupled to an engineered heterotrimeric G protein. Biochem. Biophys. Res. Commun. 2020, 533, 861–866. [Google Scholar] [CrossRef]
  44. Coupe, M.; Rees, H.; Springer, C.J.; Bishop, A.E.; Morris, J.A.; Polak, J.M.; Calam, J. Gastric enterochromaffin-like (ECL) cells in hypergastrinaemic duodenal ulcer disease. Gut 1990, 31, 144–147. [Google Scholar] [CrossRef] [PubMed]
  45. Cui, G.; Waldum, H.L. Physiological and clinical significance of enterochromaffin-like cell activation in the regulation of gastric acid secretion. World J. Gastroenterol. 2007, 13, 493–496. [Google Scholar] [CrossRef]
  46. Garrett, W.S.; Gordon, J.I.; Glimcher, L.H. Homeostasis and Inflammation in the Intestine. Cell 2010, 140, 859–870. [Google Scholar] [CrossRef]
  47. Shin, W.; Kim, H.J. Intestinal barrier dysfunction orchestrates the onset of inflammatory host-microbiome cross-talk in a human gut inflammation-on-a-chip. Proc. Natl. Acad. Sci. USA 2018, 115, E10539–E10547. [Google Scholar] [CrossRef]
  48. Furness, J.B.; Rivera, L.R.; Cho, H.J.; Bravo, D.M.; Callaghan, B. The gut as a sensory organ. Nat. Rev. Gastroenterol. Hepatol. 2013, 10, 729–740. [Google Scholar] [CrossRef]
  49. Lebrun, L.J.; Lenaerts, K.; Kiers, D.; Pais de Barros, J.P.; Le Guern, N.; Plesnik, J.; Thomas, C.; Bourgeois, T.; Dejong, C.H.C.; Kox, M.; et al. Enteroendocrine L Cells Sense LPS after Gut Barrier Injury to Enhance GLP-1 Secretion. Cell Rep. 2017, 21, 1160–1168. [Google Scholar] [CrossRef]
  50. 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]
  51. Katsuma, S.; Hatae, N.; Yano, T.; Ruike, Y.; Kimura, M.; Hirasawa, A.; Tsujimoto, G. Free fatty acids inhibit serum deprivation-induced apoptosis through GPR120 in a murine enteroendocrine cell line STC-1. J. Biol. Chem. 2005, 280, 19507–19515. [Google Scholar] [CrossRef] [PubMed]
  52. Zhang, W.J.; Duan, J.Z.; Lei, N.; Xing, H.; Shao, Y.; Yang, G.B. Cellular bases for interactions between immunocytes and enteroendocrine cells in the intestinal mucosal barrier of rhesus macaques. Cell Tissue Res. 2012, 350, 135–141. [Google Scholar] [CrossRef] [PubMed]
  53. Begg, D.P.; Woods, S.C. The endocrinology of food intake. Nat. Rev. Endocrinol. 2013, 9, 584–597. [Google Scholar] [CrossRef] [PubMed]
  54. Worthington, J.J.; Reimann, F.; Gribble, F.M. Enteroendocrine cells-sensory sentinels of the intestinal environment and orchestrators of mucosal immunity. Mucosal Immunol. 2018, 11, 3–20. [Google Scholar] [CrossRef]
  55. Ye, L.; Bae, M.; Cassilly, C.D.; Jabba, S.V.; Thorpe, D.W.; Martin, A.M.; Lu, H.Y.; Wang, J.; Thompson, J.D.; Lickwar, C.R.; et al. Enteroendocrine cells sense bacterial tryptophan catabolites to activate enteric and vagal neuronal pathways. Cell Host Microbe 2021, 29, 179–196.e9. [Google Scholar] [CrossRef] [PubMed]
  56. Jarocka-Cyrta, E.; Kasacka, I.; Kaczmarski, M. The ghrelin-positive cells number is increased in duodenum in children with celiac disease. J. Endocrinol. Investig. 2010, 33, 165–170. [Google Scholar] [CrossRef]
  57. Caddy, G.R.; Ardill, J.E.; Fillmore, D.; Caldwell, C.M.; McKibben, B.M.; Gardiner, K.R.; Watson, P.R. Plasma concentrations of glucagon-like peptide-2 in adult patients with treated and untreated coeliac disease. Eur. J. Gastroenterol. Hepatol. 2006, 18, 195–202. [Google Scholar] [CrossRef]
  58. Fraquelli, M.; Bardella, M.T.; Peracchi, M.; Cesana, B.M.; Bianchi, P.A.; Conte, D. Gallbladder emptying and somatostatin and cholecystokinin plasma levels in celiac disease. Am. J. Gastroenterol. 1999, 94, 1866–1870. [Google Scholar] [CrossRef]
  59. Di Sabatino, A.; Giuffrida, P.; Vanoli, A.; Luinetti, O.; Manca, R.; Biancheri, P.; Bergamaschi, G.; Alvisi, C.; Pasini, A.; Salvatore, C.; et al. Increase in neuroendocrine cells in the duodenal mucosa of patients with refractory celiac disease. Am. J. Gastroenterol. 2014, 109, 258–269. [Google Scholar] [CrossRef]
  60. El-Salhy, M.; Danielsson, A.; Stenling, R.; Grimelius, L. Colonic endocrine cells in inflammatory bowel disease. J. Intern. Med. 1997, 242, 413–419. [Google Scholar] [CrossRef]
  61. Moran, G.W.; Pennock, J.; McLaughlin, J.T. Enteroendocrine cells in terminal ileal Crohn’s disease. J. Crohns Colitis 2012, 6, 871–880. [Google Scholar] [CrossRef]
  62. Keller, J.; Beglinger, C.; Holst, J.J.; Andresen, V.; Layer, P. Mechanisms of gastric emptying disturbances in chronic and acute inflammation of the distal gastrointestinal tract. Am. J. Physiol. Gastrointest. Liver Physiol. 2009, 297, G861–G868. [Google Scholar] [CrossRef]
  63. Yu, Y.; Yang, W.; Li, Y.; Cong, Y. Enteroendocrine Cells: Sensing Gut Microbiota and Regulating Inflammatory Bowel Diseases. Inflamm. Bowel Dis. 2020, 26, 11–20. [Google Scholar] [CrossRef]
  64. Friedrich, M.; Diegelmann, J.; Schauber, J.; Auernhammer, C.J.; Brand, S. Intestinal neuroendocrine cells and goblet cells are mediators of IL-17A-amplified epithelial IL-17C production in human inflammatory bowel disease. Mucosal Immunol. 2015, 8, 943–958. [Google Scholar] [CrossRef] [PubMed]
  65. El-Salhy, M.; Gundersen, D.; Hatlebakk, J.G.; Hausken, T. High densities of serotonin and peptide YY cells in the colon of patients with lymphocytic colitis. World J. Gastroenterol. 2012, 18, 6070–6075. [Google Scholar] [CrossRef] [PubMed]
  66. El-Salhy, M.; Mazzawi, T.; Gundersen, D.; Hatlebakk, J.G.; Hausken, T. The role of peptide YY in gastrointestinal diseases and disorders (review). Int. J. Mol. Med. 2013, 31, 275–282. [Google Scholar] [CrossRef] [PubMed]
  67. Khalaf, A.; Hoad, C.L.; Menys, A.; Nowak, A.; Radford, S.; Taylor, S.A.; Latief, K.; Lingaya, M.; Falcone, Y.; Singh, G.; et al. Gastrointestinal peptides and small-bowel hypomotility are possible causes for fasting and postprandial symptoms in active Crohn’s disease. Am. J. Clin. Nutr. 2020, 111, 131–140. [Google Scholar] [CrossRef] [PubMed]
  68. Qi, K.K.; Wu, J.; Wan, J.; Men, X.M.; Xu, Z.W. Purified PEGylated porcine glucagon-like peptide-2 reduces the severity of colonic injury in a murine model of experimental colitis. Peptides 2014, 52, 11–18. [Google Scholar] [CrossRef]
  69. Bendotti, G.; Montefusco, L.; Lunati, M.E.; Usuelli, V.; Pastore, I.; Lazzaroni, E.; Assi, E.; Seelam, A.J.; El Essawy, B.; Jang, J.; et al. The anti-inflammatory and immunological properties of GLP-1 Receptor Agonists. Pharmacol. Res. 2022, 182, 106320. [Google Scholar] [CrossRef]
  70. Lee, Y.S.; Jun, H.S. Anti-Inflammatory Effects of GLP-1-Based Therapies beyond Glucose Control. Mediat. Inflamm. 2016, 2016, 3094642. [Google Scholar] [CrossRef]
  71. Blonski, W.; Buchner, A.M.; Aberra, F.; Lichtenstein, G. Teduglutide in Crohn’s disease. Expert Opin. Biol. Ther. 2013, 13, 1207–1214. [Google Scholar] [CrossRef]
  72. Pizzoferrato, M.; Puca, P.; Ennas, S.; Cammarota, G.; Guidi, L. Glucagon-like peptide-2 analogues for Crohn’s disease patients with short bowel syndrome and intestinal failure. World J. Gastroenterol. 2022, 28, 6258–6270. [Google Scholar] [CrossRef]
  73. Marier, J.F.; Jomphe, C.; Peyret, T.; Wang, Y. Population pharmacokinetics and exposure-response analyses of teduglutide in adult and pediatric patients with short bowel syndrome. Clin. Transl. Sci. 2021, 14, 2497–2509. [Google Scholar] [CrossRef]
  74. Eliasson, J.; Hvistendahl, M.K.; Freund, N.; Bolognani, F.; Meyer, C.; Jeppesen, P.B. Apraglutide, a novel once-weekly glucagon-like peptide-2 analog, improves intestinal fluid and energy absorption in patients with short bowel syndrome: An open-label phase 1 and 2 metabolic balance trial. J. Parenter. Enteral. Nutr. 2022, 46, 1639–1649. [Google Scholar] [CrossRef] [PubMed]
  75. Caio, G.; Volta, U.; Sapone, A.; Leffler, D.A.; De Giorgio, R.; Catassi, C.; Fasano, A. Celiac disease: A comprehensive current review. BMC Med. 2019, 17, 142. [Google Scholar] [CrossRef]
  76. Han, A.; Newell, E.W.; Glanville, J.; Fernandez-Becker, N.; Khosla, C.; Chien, Y.H.; Davis, M.M. Dietary gluten triggers concomitant activation of CD4+ and CD8+ alphabeta T cells and gammadelta T cells in celiac disease. Proc. Natl. Acad. Sci. USA 2013, 110, 13073–13078. [Google Scholar] [CrossRef] [PubMed]
  77. Freire, R.; Ingano, L.; Serena, G.; Cetinbas, M.; Anselmo, A.; Sapone, A.; Sadreyev, R.I.; Fasano, A.; Senger, S. Human gut derived-organoids provide model to study gluten response and effects of microbiota-derived molecules in celiac disease. Sci. Rep. 2019, 9, 7029. [Google Scholar] [CrossRef] [PubMed]
  78. el-Salhy, M. The nature and implication of intestinal endocrine cell changes in coeliac disease. Histol. Histopathol. 1998, 13, 1069–1075. [Google Scholar] [CrossRef] [PubMed]
  79. Domschke, S.; Bloom, S.R.; Adrian, T.E.; Lux, G.; Bryant, M.G.; Domschke, W. Coeliac sprue: Abnormalities of the hormone profile of gastroduodenal mucosa. Scand. J. Gastroenterol. 1989, 167, 86–89. [Google Scholar] [CrossRef]
  80. Papastamataki, M.; Papassotiriou, I.; Bartzeliotou, A.; Vazeou, A.; Roma, E.; Chrousos, G.P.; Kanaka-Gantenbein, C. Incretins, amylin and other gut-brain axis hormones in children with coeliac disease. Eur. J. Clin. Investig. 2014, 44, 74–82. [Google Scholar] [CrossRef]
  81. Rocco, A.; Sarnelli, G.; Compare, D.; de Colibus, P.; Micheli, P.; Somma, P.; Marotti, B.; Cuomo, R.; Nardone, G. Tissue ghrelin level and gastric emptying rate in adult patients with celiac disease. Neurogastroenterol. Motil. 2008, 20, 884–890. [Google Scholar] [CrossRef] [PubMed]
  82. Sjolund, K.; Nobin, A. Increased levels of plasma 5-hydroxytryptamine in patients with coeliac disease. Scand. J. Gastroenterol. 1985, 20, 304–308. [Google Scholar] [CrossRef] [PubMed]
  83. Bardella, M.T.; Fraquelli, M.; Peracchi, M.; Cesana, B.M.; Bianchi, P.A.; Conte, D. Gastric emptying and plasma neurotensin levels in untreated celiac patients. Scand. J. Gastroenterol. 2000, 35, 269–273. [Google Scholar] [CrossRef] [PubMed]
  84. Le Quellec, A.; Kervran, A.; Blache, P.; Ciurana, A.J.; Bataille, D. [Oxyntomodulin, a new hormonal marker of intestinal malabsorption syndromes]. Rev. Med. Interne 1993, 14, 982. [Google Scholar] [CrossRef]
  85. Hujoel, I.A.; Murray, J.A. Refractory Celiac Disease. Curr. Gastroenterol. Rep. 2020, 22, 18. [Google Scholar] [CrossRef]
  86. French, S.J.; Murray, B.; Rumsey, R.D.; Sepple, C.P.; Read, N.W. Preliminary studies on the gastrointestinal responses to fatty meals in obese people. Int. J. Obes. Relat. Metab. Disord. 1993, 17, 295–300. [Google Scholar]
  87. Ahmed, R.H.; Huri, H.Z.; Muniandy, S.; Al-Hamodi, Z.; Al-Absi, B.; Alsalahi, A.; Razif, M.F. Altered circulating concentrations of active glucagon-like peptide (GLP-1) and dipeptidyl peptidase 4 (DPP4) in obese subjects and their association with insulin resistance. Clin. Biochem. 2017, 50, 746–749. [Google Scholar] [CrossRef]
  88. Vrang, N.; Madsen, A.N.; Tang-Christensen, M.; Hansen, G.; Larsen, P.J. PYY(3-36) reduces food intake and body weight and improves insulin sensitivity in rodent models of diet-induced obesity. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2006, 291, R367–R375. [Google Scholar] [CrossRef]
  89. Osinski, C.; Le Gleau, L.; Poitou, C.; de Toro-Martin, J.; Genser, L.; Fradet, M.; Soula, H.A.; Leturque, A.; Blugeon, C.; Jourdren, L.; et al. Type 2 diabetes is associated with impaired jejunal enteroendocrine GLP-1 cell lineage in human obesity. Int. J. Obes. 2021, 45, 170–183. [Google Scholar] [CrossRef]
  90. Jorsal, T.; Rhee, N.A.; Pedersen, J.; Wahlgren, C.D.; Mortensen, B.; Jepsen, S.L.; Jelsing, J.; Dalboge, 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]
  91. Sakran, N.; Graham, Y.; Pintar, T.; Yang, W.; Kassir, R.; Willigendael, E.M.; Singhal, R.; Kooreman, Z.E.; Ramnarain, D.; Mahawar, K.; et al. The many faces of diabetes. Is there a need for re-classification? A narrative review. BMC Endocr. Disord. 2022, 22, 9. [Google Scholar] [CrossRef] [PubMed]
  92. Booth, H.; Khan, O.; Prevost, T.; Reddy, M.; Dregan, A.; Charlton, J.; Ashworth, M.; Rudisill, C.; Littlejohns, P.; Gulliford, M.C. Incidence of type 2 diabetes after bariatric surgery: Population-based matched cohort study. Lancet Diabetes Endocrinol. 2014, 2, 963–968. [Google Scholar] [CrossRef]
  93. Muller, T.D.; Finan, B.; Bloom, S.R.; D’Alessio, D.; Drucker, D.J.; Flatt, P.R.; Fritsche, A.; Gribble, F.; Grill, H.J.; Habener, J.F.; et al. Glucagon-like peptide 1 (GLP-1). Mol. Metab. 2019, 30, 72–130. [Google Scholar] [CrossRef]
  94. Holst, J.J.; Madsbad, S. Mechanisms of surgical control of type 2 diabetes: GLP-1 is key factor. Surg. Obes. Relat. Dis. 2016, 12, 1236–1242. [Google Scholar] [CrossRef] [PubMed]
  95. Nauck, M.A.; Quast, D.R.; Wefers, J.; Meier, J.J. GLP-1 receptor agonists in the treatment of type 2 diabetes—State-of-the-art. Mol. Metab. 2021, 46, 101102. [Google Scholar] [CrossRef] [PubMed]
  96. Nohr, 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]
  97. 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] [PubMed]
  98. Rozengurt, N.; Wu, S.V.; Chen, M.C.; Huang, C.; Sternini, C.; Rozengurt, E. Colocalization of the alpha-subunit of gustducin with PYY and GLP-1 in L cells of human colon. Am. J. Physiol. Gastrointest. Liver Physiol. 2006, 291, G792–G802. [Google Scholar] [CrossRef]
  99. Poher, A.L.; Tschop, M.H.; Muller, T.D. Ghrelin regulation of glucose metabolism. Peptides 2018, 100, 236–242. [Google Scholar] [CrossRef]
  100. Gagnon, J.; Anini, Y. Insulin and norepinephrine regulate ghrelin secretion from a rat primary stomach cell culture. Endocrinology 2012, 153, 3646–3656. [Google Scholar] [CrossRef]
  101. Cawthon, C.R.; de La Serre, C.B. The critical role of CCK in the regulation of food intake and diet-induced obesity. Peptides 2021, 138, 170492. [Google Scholar] [CrossRef]
  102. Steinert, R.E.; Feinle-Bisset, C.; Asarian, L.; Horowitz, M.; Beglinger, C.; Geary, N. Ghrelin, CCK, GLP-1, and PYY(3-36): Secretory Controls and Physiological Roles in Eating and Glycemia in Health, Obesity, and after RYGB. Physiol. Rev. 2017, 97, 411–463. [Google Scholar] [CrossRef]
  103. Little, T.J.; Horowitz, M.; Feinle-Bisset, C. Modulation by high-fat diets of gastrointestinal function and hormones associated with the regulation of energy intake: Implications for the pathophysiology of obesity. Am. J. Clin. Nutr. 2007, 86, 531–541. [Google Scholar] [CrossRef]
  104. Brennan, I.M.; Luscombe-Marsh, N.D.; Seimon, R.V.; Otto, B.; Horowitz, M.; Wishart, J.M.; Feinle-Bisset, C. Effects of fat, protein, and carbohydrate and protein load on appetite, plasma cholecystokinin, peptide YY, and ghrelin, and energy intake in lean and obese men. Am. J. Physiol. Gastrointest. Liver Physiol. 2012, 303, G129–G140. [Google Scholar] [CrossRef]
  105. Bodnaruc, A.M.; Prud’homme, D.; Blanchet, R.; Giroux, I. Nutritional modulation of endogenous glucagon-like peptide-1 secretion: A review. Nutr. Metab. 2016, 13, 92. [Google Scholar] [CrossRef]
  106. Acosta, A.; Camilleri, M.; Shin, A.; Vazquez-Roque, M.I.; Iturrino, J.; Burton, D.; O’Neill, J.; Eckert, D.; Zinsmeister, A.R. Quantitative gastrointestinal and psychological traits associated with obesity and response to weight-loss therapy. Gastroenterology 2015, 148, 537–546.e4. [Google Scholar] [CrossRef] [PubMed]
  107. Htike, Z.Z.; Zaccardi, F.; Papamargaritis, D.; Webb, D.R.; Khunti, K.; Davies, M.J. Efficacy and safety of glucagon-like peptide-1 receptor agonists in type 2 diabetes: A systematic review and mixed-treatment comparison analysis. Diabetes Obes. Metab. 2017, 19, 524–536. [Google Scholar] [CrossRef] [PubMed]
  108. Karra, E.; Chandarana, K.; Batterham, R.L. The role of peptide YY in appetite regulation and obesity. J. Physiol. 2009, 587, 19–25. [Google Scholar] [CrossRef]
  109. Tschop, M.; Castaneda, T.R.; Joost, H.G.; Thone-Reineke, C.; Ortmann, S.; Klaus, S.; Hagan, M.M.; Chandler, P.C.; Oswald, K.D.; Benoit, S.C.; et al. Physiology: Does gut hormone PYY3-36 decrease food intake in rodents? Nature 2004, 430, 1 p following 165; discussion 2 p following 165. [Google Scholar] [CrossRef]
  110. Muller, T.D.; Bluher, M.; Tschop, M.H.; DiMarchi, R.D. Anti-obesity drug discovery: Advances and challenges. Nat. Rev. Drug Discov. 2022, 21, 201–223. [Google Scholar] [CrossRef] [PubMed]
  111. Sei, Y.; Lu, X.; Liou, A.; Zhao, X.; Wank, S.A. A stem cell marker-expressing subset of enteroendocrine cells resides at the crypt base in the small intestine. Am. J. Physiol. Gastrointest. Liver Physiol. 2011, 300, G345–G356. [Google Scholar] [CrossRef]
  112. Johnson, N.M.; Parham, L.R.; Na, J.; Monaghan, K.E.; Kolev, H.M.; Klochkova, A.; Kim, M.S.; Danan, C.H.; Cramer, Z.; Simon, L.A.; et al. Autophagic state prospectively identifies facultative stem cells in the intestinal epithelium. EMBO Rep. 2022, 23, e55209. [Google Scholar] [CrossRef] [PubMed]
  113. Sei, Y.; Feng, J.; Zhao, X.; Wank, S.A. Role of an active reserve stem cell subset of enteroendocrine cells in intestinal stem cell dynamics and the genesis of small intestinal neuroendocrine tumors. Am. J. Physiol. Gastrointest. Liver Physiol. 2020, 319, G494–G501. [Google Scholar] [CrossRef] [PubMed]
  114. Kwak, S.Y.; Shim, S.; Park, S.; Kim, H.; Lee, S.J.; Kim, M.J.; Jang, W.S.; Kim, Y.H.; Jang, H. Ghrelin reverts intestinal stem cell loss associated with radiation-induced enteropathy by activating Notch signaling. Phytomedicine 2021, 81, 153424. [Google Scholar] [CrossRef]
  115. Zhang, H.; Cui, Z.; Luo, G.; Zhang, J.; Ma, T.; Hu, N.; Cui, T. Ghrelin attenuates intestinal ischemia/reperfusion injury in mice by activating the mTOR signaling pathway. Int. J. Mol. Med. 2013, 32, 851–859. [Google Scholar] [CrossRef] [PubMed]
  116. Zheng, H.; Liang, W.; He, W.; Huang, C.; Chen, Q.; Yi, H.; Long, L.; Deng, Y.; Zeng, M. Ghrelin attenuates sepsis-induced acute lung injury by inhibiting the NF-kappaB, iNOS, and Akt signaling in alveolar macrophages. Am. J. Physiol. Lung Cell. Mol. Physiol. 2019, 317, L381–L391. [Google Scholar] [CrossRef] [PubMed]
  117. Lesurtel, M.; Graf, R.; Aleil, B.; Walther, D.J.; Tian, Y.; Jochum, W.; Gachet, C.; Bader, M.; Clavien, P.A. Platelet-derived serotonin mediates liver regeneration. Science 2006, 312, 104–107. [Google Scholar] [CrossRef]
  118. Larsson, L.I.; Goltermann, N.; de Magistris, L.; Rehfeld, J.F.; Schwartz, T.W. Somatostatin cell processes as pathways for paracrine secretion. Science 1979, 205, 1393–1395. [Google Scholar] [CrossRef]
  119. Bottcher, G.; Sjolund, K.; Ekblad, E.; Hakanson, R.; Schwartz, T.W.; Sundler, F. Coexistence of peptide YY and glicentin immunoreactivity in endocrine cells of the gut. Regul. Pept. 1984, 8, 261–266. [Google Scholar] [CrossRef]
  120. Dockray, G.J. Gastrointestinal hormones and the dialogue between gut and brain. J. Physiol. 2014, 592, 2927–2941. [Google Scholar] [CrossRef] [PubMed]
  121. Breer, H.; Eberle, J.; Frick, C.; Haid, D.; Widmayer, P. Gastrointestinal chemosensation: Chemosensory cells in the alimentary tract. Histochem. Cell Biol. 2012, 138, 13–24. [Google Scholar] [CrossRef] [PubMed]
  122. 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. 2011, 519, 644–660. [Google Scholar] [CrossRef]
  123. Liddle, R.A. Neuropods. Cell. Mol. Gastroenterol. Hepatol. 2019, 7, 739–747. [Google Scholar] [CrossRef]
  124. Generoso, J.S.; Giridharan, V.V.; Lee, J.; Macedo, D.; Barichello, T. The role of the microbiota-gut-brain axis in neuropsychiatric disorders. Braz. J. Psychiatr. 2021, 43, 293–305. [Google Scholar] [CrossRef] [PubMed]
  125. Liu, L.; Huh, J.R.; Shah, K. Microbiota and the gut-brain-axis: Implications for new therapeutic design in the CNS. EBioMedicine 2022, 77, 103908. [Google Scholar] [CrossRef] [PubMed]
  126. Liddle, R.A. Parkinson’s disease from the gut. Brain Res. 2018, 1693, 201–206. [Google Scholar] [CrossRef]
  127. Chandra, R.; Hiniker, A.; Kuo, Y.M.; Nussbaum, R.L.; Liddle, R.A. alpha-Synuclein in gut endocrine cells and its implications for Parkinson’s disease. JCI Insight 2017, 2, e92295. [Google Scholar] [CrossRef] [PubMed]
  128. Singh, S.P.; Singh, V.; Kar, N.; Chan, K. Efficacy of antidepressants in treating the negative symptoms of chronic schizophrenia: Meta-analysis. Br. J. Psychiatr. 2010, 197, 174–179. [Google Scholar] [CrossRef] [PubMed]
  129. Sharma, R.P.; Shapiro, L.E.; Kamath, S.K.; Soll, E.A.; Watanabe, M.D.; Davis, J.M. Acute dietary tryptophan depletion: Effects on schizophrenic positive and negative symptoms. Neuropsychobiology 1997, 35, 5–10. [Google Scholar] [CrossRef] [PubMed]
  130. Selvaraj, S.; Arnone, D.; Cappai, A.; Howes, O. Alterations in the serotonin system in schizophrenia: A systematic review and meta-analysis of postmortem and molecular imaging studies. Neurosci. Biobehav. Rev. 2014, 45, 233–245. [Google Scholar] [CrossRef]
  131. Mawe, G.M.; Hoffman, J.M. Serotonin signalling in the gut--functions, dysfunctions and therapeutic targets. Nat. Rev. Gastroenterol. Hepatol. 2013, 10, 473–486. [Google Scholar] [CrossRef]
  132. Greenwood-Van Meerveld, B.; Mohammadi, E.; Tyler, K.; Pietra, C.; Bee, L.A.; Dickenson, A. Synergistic effect of 5-hydroxytryptamine 3 and neurokinin 1 receptor antagonism in rodent models of somatic and visceral pain. J. Pharmacol. Exp. Ther. 2014, 351, 146–152. [Google Scholar] [CrossRef] [PubMed]
  133. Barton, J.R.; Londregran, A.K.; Alexander, T.D.; Entezari, A.A.; Bar-Ad, S.; Cheng, L.; Lepore, A.C.; Snook, A.E.; Covarrubias, M.; Waldman, S.A. Intestinal neuropod cell GUCY2C regulates visceral pain. J. Clin. Investig. 2023, 133, e165578. [Google Scholar] [CrossRef] [PubMed]
  134. Spencer, N.J.; Keating, D.J. Role of 5-HT in the enteric nervous system and enteroendocrine cells. Br. J. Pharmacol. 2022. [Google Scholar] [CrossRef] [PubMed]
  135. Martin, G.R.; Beck, P.L.; Sigalet, D.L. Gut hormones, and short bowel syndrome: The enigmatic role of glucagon-like peptide-2 in the regulation of intestinal adaptation. World J. Gastroenterol. 2006, 12, 4117–4129. [Google Scholar] [CrossRef]
  136. Jeppesen, P.B.; Hartmann, B.; Hansen, B.S.; Thulesen, J.; Holst, J.J.; Mortensen, P.B. Impaired meal stimulated glucagon-like peptide 2 response in ileal resected short bowel patients with intestinal failure. Gut 1999, 45, 559–563. [Google Scholar] [CrossRef]
  137. Drucker, D.J.; Erlich, P.; Asa, S.L.; Brubaker, P.L. Induction of intestinal epithelial proliferation by glucagon-like peptide 2. Proc. Natl. Acad. Sci. USA 1996, 93, 7911–7916. [Google Scholar] [CrossRef]
  138. O’Neil, P.M.; Birkenfeld, A.L.; McGowan, B.; Mosenzon, O.; Pedersen, S.D.; Wharton, S.; Carson, C.G.; Jepsen, C.H.; Kabisch, M.; Wilding, J.P.H. Efficacy and safety of semaglutide compared with liraglutide and placebo for weight loss in patients with obesity: A randomised, double-blind, placebo and active controlled, dose-ranging, phase 2 trial. Lancet 2018, 392, 637–649. [Google Scholar] [CrossRef]
  139. Wilding, J.P.H.; Batterham, R.L.; Calanna, S.; Davies, M.; Van Gaal, L.F.; Lingvay, I.; McGowan, B.M.; Rosenstock, J.; Tran, M.T.D.; Wadden, T.A.; et al. Once-Weekly Semaglutide in Adults with Overweight or Obesity. N. Engl. J. Med. 2021, 384, 989–1002. [Google Scholar] [CrossRef]
  140. Ghusn, W.; De la Rosa, A.; Sacoto, D.; Cifuentes, L.; Campos, A.; Feris, F.; Hurtado, M.D.; Acosta, A. Weight Loss Outcomes Associated With Semaglutide Treatment for Patients With Overweight or Obesity. JAMA Netw. Open 2022, 5, e2231982. [Google Scholar] [CrossRef]
  141. Peiris, M.; Aktar, R.; Reed, D.; Cibert-Goton, V.; Zdanaviciene, A.; Halder, W.; Robinow, A.; Corke, S.; Dogra, H.; Knowles, C.H.; et al. Decoy bypass for appetite suppression in obese adults: Role of synergistic nutrient sensing receptors GPR84 and FFAR4 on colonic endocrine cells. Gut 2022, 71, 928–937. [Google Scholar] [CrossRef] [PubMed]
Table 1. Known classifications of enteroendocrine cells.
Table 1. Known classifications of enteroendocrine cells.
SubtypeHormones or PeptidesReceptorsLocationMain FunctionReferences
EC cellsSerotonin (5-HT)FFARs 2,3; TRPA1; TLRs; toxin receptorsSmall intestine and colonGut motility and secretion, nausea, and vomiting[21,22,23]
L cellsGLP-1, GLP-2, peptide YYT2Rs; T1R2-T1R3; FFARs 1–3; GPR119, LPAR5, GPR120; CaSRIleum, colonAppetite control, insulin release, slows intestinal transit, and stimulates cell proliferation [24,25,26]
X/A-like cellsGhrelin, nesfatin-1T1R1-T1R3; T2RsStomachAppetite control and growth hormone release[27,28]
K cellsGIP, 5-HTGPR119, GPR120; FFAR1Proximal small intestineStimulate insulin release and regulates gut mucosal growth[29,30]
I cellsCholecystokinin, 5-HTT2Rs; FFA1; GPR120; CCK1R; CCK2 Proximal small intestineStimulate pancreatic enzyme secretion and activate bile release[31,32]
D cellsSomatostatinGPCRsStomach, small intestine Inhibit gastric acid secretion, motility, and intestinal absorption[33,34]
G cellsGastrinCCK2RStomach, duodenumRegulate gastric acid secretion and epithelial proliferation[35,36,37]
N cellsNeurotensinNTSR1, NTSR2Small intestineStimulate pancreatic and biliary secretions and suppress small-bowel gastric motility[38,39]
M cellsMotilinPAFR, TLR-4, α5β1-integrinSmall intestine, colon, rectumRegulate the migrating motor complex and increase insulin release[40,41]
S cellsSecretinSR, SecRStomach, small intestineRegulate pancreatic exocrine secretion, gastric acid secretion, and gastric motility[42,43]
ECL-like cellsHistamineH1R, H2R, H3RStomachStimulate gastric acid (HCl) and pepsinogen secretion[44,45]
Table 2. Summary of changes to enteroendocrine cells and secreted hormones or peptides in inflammatory diseases.
Table 2. Summary of changes to enteroendocrine cells and secreted hormones or peptides in inflammatory diseases.
PathophysiologySpecific EE Subtypes ChangedHormone or Peptide ChangedReference
Celiac diseaseX/A-like cells Ghrelin[56]
Celiac diseaseL cells GLP-2[57]
Celiac diseaseD cells SST[58]
Refractory Celiac disease (RCD)EC cells TPH1[59]
Crohn’s diseaseL cells PYY[60]
Terminal ileal Crohn’s diseaseL cells GLP-1, PYY[61]
Ulcerative colitisL cells GLP-1[62]
DiverticulitisL cells GLP-1[62]
Gut inflammation via LPSL cells GLP-1[49]
EE, enteroendocrine; EC, enterochromaffin; LPS, lipopolysaccharide; GLP, glucagon-like peptide; SST, somatostatin; TPH1, tryptophan hydroxylase 1; PYY, peptide tyrosine tyrosine. Red up arrows indicate an increase, blue down arrows indicate a decrease.
Table 3. Summary of changes to enteroendocrine cells and secreted hormones or peptides in metabolic diseases.
Table 3. Summary of changes to enteroendocrine cells and secreted hormones or peptides in metabolic diseases.
PathophysiologySpecific EE Subtypes ChangedHormone or Peptide ChangedReference
ObesityI cells CCK[86]
ObesityL cells GLP-1[87]
ObesityL cells PYY[88]
Obesity-induced Type II diabetesL cells GLP-1[89]
Type II diabetesK cells, L cells GIP, GCG, PYY[90]
EE, enteroendocrine; CCK, cholecystokinin, GLP, glucagon-like peptide; PYY, peptide tyrosine tyrosine; GIP, gastric inhibitory peptide; GCG, proglucagon. Red up arrows indicate an increase, blue down arrows indicate a decrease.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Atanga, R.; Singh, V.; In, J.G. Intestinal Enteroendocrine Cells: Present and Future Druggable Targets. Int. J. Mol. Sci. 2023, 24, 8836. https://doi.org/10.3390/ijms24108836

AMA Style

Atanga R, Singh V, In JG. Intestinal Enteroendocrine Cells: Present and Future Druggable Targets. International Journal of Molecular Sciences. 2023; 24(10):8836. https://doi.org/10.3390/ijms24108836

Chicago/Turabian Style

Atanga, Roger, Varsha Singh, and Julie G. In. 2023. "Intestinal Enteroendocrine Cells: Present and Future Druggable Targets" International Journal of Molecular Sciences 24, no. 10: 8836. https://doi.org/10.3390/ijms24108836

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