Systematic Reviews Open Access
Copyright ©The Author(s) 2015. Published by Baishideng Publishing Group Inc. All rights reserved.
World J Gastrointest Pharmacol Ther. May 6, 2015; 6(2): 22-27
Published online May 6, 2015. doi: 10.4292/wjgpt.v6.i2.22
Clinical relevance of intestinal peptide uptake
Hugh James Freeman, Department of Medicine, University of British Columbia, Vancouver V6T 1W5, Canada
Author contributions: Freeman HJ contributed all to this paper.
Conflict-of-interest: No conflict of interest declared.
Data sharing: No data sharing.
Open-Access: This article is an open-access article which was selected by an in-house editor and fully peer-reviewed by external reviewers. It is distributed in accordance with the Creative Commons Attribution Non Commercial (CC BY-NC 4.0) license, which permits others to distribute, remix, adapt, build upon this work non-commercially, and license their derivative works on different terms, provided the original work is properly cited and the use is non-commercial. See: http://creativecommons.org/licenses/by-nc/4.0/
Correspondence to: Dr. Hugh James Freeman, MD, CM, FRCPC, FACP, Department of Medicine, University of British Columbia, 2211 Wesbrook Mall, Vancouver V6T 1W5, Canada. hugfree@shaw.ca
Telephone: +1-604-8227216 Fax: +1-604-8227236
Received: October 9, 2014
Peer-review started: October 9, 2014
First decision: November 14, 2014
Revised: November 22, 2014
Accepted: March 30, 2015
Article in press: April 2, 2015
Published online: May 6, 2015

Abstract

AIM: To determine available information on an independent peptide transporter 1 (PepT1) and its potential relevance to treatment, this evaluation was completed.

METHODS: Fully published English language literature articles sourced through PubMed related to protein digestion and absorption, specifically human peptide and amino acid transport, were accessed and reviewed. Papers from 1970 to the present, with particular emphasis on the past decade, were examined. In addition, abstracted information translated to English in PubMed was also included. Finally, studies and reviews relevant to nutrient or drug uptake, particularly in human intestine were included for evaluation. This work represents a summary of all of these studies with particular reference to peptide transporter mediated assimilation of nutrients and pharmacologically active medications.

RESULTS: Assimilation of dietary protein in humans involves gastric and pancreatic enzyme hydrolysis to luminal oligopeptides and free amino acids. During the ensuing intestinal phase, these hydrolytic products are transported into the epithelial cell and, eventually, the portal vein. A critical component of this process is the uptake of intact di-peptides and tri-peptides by an independent PepT1. A number of “peptide-mimetic” pharmaceutical agents may also be transported through this carrier, important for uptake of different antibiotics, antiviral agents and angiotensin-converting enzyme inhibitors. In addition, specific peptide products of intestinal bacteria may also be transported by PepT1, with initiation and persistence of an immune response including increased cytokine production and associated intestinal inflammatory changes. Interestingly, these inflammatory changes may also be attenuated with orally-administered anti-inflammatory tripeptides administered as site-specific nanoparticles and taken up by this PepT1 transport protein.

CONCLUSION: Further evaluation of the role of this transporter in treatment of intestinal disorders, including inflammatory bowel disease is needed.

Key Words: Dietary peptides, Peptide transport, Peptide transporter 1, Intestinal inflammation, Drug absorption, Bacterial peptides

Core tip: Intestinal uptake of intact di-peptides and tri-peptides occurs by an independent epithelial transport process for protein assimilation. This carrier may also be used to absorb specific drugs and bacterial peptide products that may result in inflammatory disease.



INTRODUCTION

Protein digestion and absorption in humans depends on initial enzymatic hydrolysis in the stomach and proximal small intestine. The hydrolytic products include oligopeptides and amino acids that ultimately undergo small intestinal uptake into the portal vein. A critical step in this overall uptake process involves a transmembrane protein [peptide transporter 1 (PepT1)], located in the brush border that can transport nutrient peptides into the enterocyte[1-3]. In addition, studies have also demonstrated that PepT1 is able to transport some pharmaceutical agents along with bacterial by-products from the intestinal lumen that may trigger an ongoing and persistent inflammatory intestinal mucosal response.

MATERIALS AND METHODS

Fully published English language literature articles sourced through PubMed related to protein digestion and absorption, specifically human peptide and amino acid transport, were accessed and reviewed. Papers from 1970 to the present, with particular emphasis on the past decade, were examined. In addition, abstracted information translated to English in PubMed was also included. Finally, studies and reviews relevant to nutrient or drug uptake, particularly in human intestine were included for evaluation. This work represents a summary of all of these studies with particular reference to peptide transporter mediated assimilation of nutrients and pharmacologically active medications.

RESULTS

Assimilation of dietary protein in humans involves gastric and pancreatic enzyme hydrolysis to luminal oligopeptides and free amino acids. During the ensuing intestinal phase, these hydrolytic products are transported into the epithelial cell and, eventually, the portal vein. A critical component of this process is the uptake of intact di-peptides and tri-peptides by an independent PepT1. A number of “peptide-mimetic” pharmaceutical agents may also be transported through this carrier, important for uptake of different antibiotics, antiviral agents and angiotensin-converting enzyme inhibitors. In addition, specific peptide products of intestinal bacteria may also be transported by PepT1, with initiation and persistence of an immune response including increased cytokine production and associated intestinal inflammatory changes. Interestingly, these inflammatory changes may also be attenuated with orally-administered anti-inflammatory tripeptides administered as site-specific nanoparticles and taken up by this PepT1 transport protein.

DISCUSSION
Gastric and pancreatic phases

Critical nutrients derived from digested protein are absorbed in the intestinal tract, specifically amino acids and peptides, during health as well as during disease. Normally, gastric and pancreatic enzymes initiate hydrolysis of dietary and other luminal proteins from endogenous sources. As a result of this initial hydrolytic phase, an array of free amino acids and different oligopeptides of variable length appear in the small intestinal lumen. Information on human protein digestion and absorption has been previously reviewed and updated[1-3].

Intestinal phase

Protein digestion studied in human volunteers using long intestinal tubes showed that infused bovine serum albumin appeared to be completely hydrolyzed before the distal ileum[4]. A host of brush border microvillus membrane transport proteins are located in the intestinal epithelial cell resulting in the uptake of specific substrates into the enterocyte. These transporters are specialized membrane proteins that can recognize, bind and translocate a specific substrate or multiple different substrates across the brush border membrane into the epithelial cell. In addition, other transport proteins involved in this process have been detected and characterized to a limited extent on the basolateral membrane. Most free amino acids that present on the luminal or apical surface of the epithelial cell are transported by both brush border and basolateral membranes into the portal venous blood. A number of brush border membrane amino acid carriers, linked to different ions, have been defined that result in transport of basic, neutral and anionic amino acids. For oligopeptides, however, different cellular transport routes are evident.

Peptide uptake

For both di-peptides and tri-peptides, a separate membrane protein, PepT1, is present that appears to have very broad substrate capacity and, theoretically, it is believed, could transport all possible di-peptides and tri-peptides into the epithelial cell[5]. For 20 different amino acids, a total of 400 di-peptides and 8000 tri-peptides have been enumerated. For those peptides that consist of 4 or more amino acids, brush border enzymes must first hydrolyze each of these to free amino acids, di-peptides and tri-peptides. Then, substrate uptake into the epithelial cell follows. Once inside the epithelial cell, cytoplasmic enzymes hydrolyze these di-peptides and tri-peptides further into free amino acids for transport into the portal venous blood. Most oligopeptidases are aminopeptidases, acting to remove an amino acid residue from the amino-terminus of the peptide. Peptide chain length determines the location of hydrolysis with longer peptides hydrolyzed at the brush border and di-peptides and tri-peptides mainly in the cytoplasm[6]. A number of other brush border and cytoplasmic peptidases are present. In particular, proline-containing oligopeptides are poorly hydrolyzed by most peptidases, yet are very important for assimilation of many normal dietary proteins with a high proline content (e.g., gliadin). Proline-specific dipeptidases are also located in the brush border membrane and cytoplasmic portion of the cell and these serve to hydrolyze most proline-containing peptides (e.g., dipeptidyl aminopeptidase IV)[7]. Particularly important was the early observation that amino acids infused into human intestine in peptide form are more readily absorbed than if infused into the intestinal lumen as free amino acids[8]. Some peptides, particularly in other non-human mammalian species, are incompletely hydrolyzed and may be transported out of enterocytes into the circulation, likely by a novel peptide transporter located in the baso-lateral membrane of the epithelial cell[9,10]. Other routes of uptake into the enterocyte have been hypothesized to exist[5]. For example, so-called “cell penetrating peptides” may carry peptides into the cell, either by direct penetration through the apical membrane or associated with endocytosis. Finally, enhanced permeability of the tight junctions between epithelial cells may result in increased paracellular uptake.

PepT1

The peptide transporters are part of a proton-coupled oligopeptide transporter superfamily, or peptide transporter family[10,11]. PepT1 (or SLC15) has several transmembrane domains and acts as a cotransporter with hydrogen ions (H+ion)[12]. After uptake of di-peptides or tri-peptides along with H+ion into the enterocyte, H+ion is then removed from the cell through the sodium-hydrogen (Na+ion/H+ion) exchanger on the brush border membrane in exchange for Na+ion. The Na+ion is then moved out of the cell by a Na+ion/K+ion ATPase pump on the basolateral membrane where 3 Na+ion are transported out of the cell and 2 K+ion are transported into the cell causing the epithelial intra-cellular electrochemical gradient to normalize.

Tissue and cellular distribution studies have also located this carrier protein in intestinal and renal brush border membranes along with lysosomal membranes. Interestingly, most PepT1 activity is located in the proximal small intestine (specifically, duodenum and jejunum), but some activity exists in other intestinal sites, including the ileum and colon. As little dietary protein actually normally reaches the distal portions of the intestine, some investigators have suggested that endogenous proteins might serve as proteolytic substrates for intestinal microflora, particularly in the colon. In addition, a transcription factor, CDX2, that appears to play important roles in proliferation, differentiation and maturation of epithelial cells, has been shown to specifically regulate this enterocyte brush border membrane transporter, PepT1[13].

The transporter has been cloned from several mammalian species, including humans, with a size estimated to be about 708 amino acids[14]. Of particular clinical importance, PepT1 may accept other non-nutrients for uptake, including pharmaceutical agents that have similar structural characteristics and actually mimic peptide substrates. These “peptide-mimetic” therapeutic agents include some antibiotics like cephalosporins and penicillins, some anti-viral agents (e.g., acyclovir, ganciclovir) and inhibitors of angiotensin-converting enzyme. Each may undergo uptake across the intestinal epithelial cell through the PepT1 transporter[15-17]. Important molecular insights into proton coupled peptide transporters have resulted from evaluation of crystal structures of bacterial transporters combined with some biochemical studies of transport, including use of genetically modified animals have recently been reviewed[18,19].

Peptide transporter regulation

A number of factors may serve to regulate PepT1, including altered dietary intake[20,21]. For example, increased expression of PepT1 may be caused by an increased quantity of dietary protein, as well as the specific amino acid composition of the dietary protein. Behavioral changes may also affect expression of the transporter. In particular, a diurnal rhythm in PepT1 expression may occur due to feeding behavior, increasing at night in some mammalian species that tend to be nocturnal feeders[22,23], a pattern abolished by fasting or imposed daytime feeding[24]. Increased expression of PepT1 during food deprivation or starvation may also occur, particularly with mucosal changes and reduced intestinal surface area associated with long-term parenteral feeding. Developmental factors also play a role in alteration of transporter expression, especially at the time following birth with suckling of a high protein milk diet and then the post-weaning phase with a shift from milk to solid food[25].

Peptide transporter in disease

PepT1 expression persists with intestinal disease, even with severe mucosal damage. Normally, PepT1 is expressed to only a limited extent in the colon compared to the small intestine[26]. In the short bowel syndrome, PepT1 expression is increased in the colon, possibly serving to conserve amino acids[27]. Similar changes have been reported in the colon of patients with inflammatory bowel diseases[28]. As a result of PepT1 up-regulation associated with the inflammatory process, dipeptides and tripeptides from bacteria in the colonic lumen may be transported by PepT1 into epithelial cells. Some of these bacterial peptides include N-formylmethionyl-leucyl-phenylalanine, a tripeptide from Escherichia coli, muramyl dipeptide (MDP), found in the cell walls of gram negative and gram positive bacteria, and L-Ala-(γ)-D-Glu-meso-diaminopimelic acid (Tri-DAP), a cell wall byproduct of gram negative bacteria. After uptake into the intestinal cell, the NF-κB pathway is activated while downstream pro-inflammatory cytokine and chemokine production are enhanced. Some bacterial peptides probably also gain access by a paracellular pathway. Added studies demonstrate that Tri-DAP transport is mediated by PepT1 expressed intestinal epithelial cells, but not in cells that did not express PepT1[29]. In the lamina propria of intestinal mucosa, these bacterial peptides may then be taken up by macrophages causing up-regulation of major histocompatibility class I molecules and increased cytokine and chemokine production, further contributing to the inflammatory process[30]. Others have reported that colonic expression of the PepT1 was down-regulated during intestinal inflammation[31]. Interestingly, PepT1 may play an important interactive role with receptors of the innate immune system to eliminate pathogens[32]. For example, ligands specific for members of the nucleotide-binding oligomerization domain (NOD) family of receptors, specifically NOD 1 and NOD 2 present in cytoplasm, may be transported by PepT1. Activation of NOD leads to NF-κB activation. A linkage with NOD mutations and risk of Crohn’s disease has been reported[33-35]. In addition, recent studies suggest that PepT1 polymorphisms may be associated with development of inflammatory bowel disease in some Scandanavian populations[36]. Recently, cellular and molecular mechanism underlying NOD 2 risk-associated polymorphisms in Crohn’s disease have been reviewed[37]. Finally, a specific tri-peptide (Lys-Pro-Val) or KPV has anti-inflammatory activities that may be transported by PepT1 and cause inhibition of NF-κB activation[38]. KPV encapsulated in polysaccharide for release primarily in the colon may reduce inflammatory mucosal changes[39]. Similarly, a PepT1-transportable soy tripeptide VPY reduced intestinal inflammation, suggesting its use as a possible treatment for inflammatory bowel disease[40].

Recently, it has been hypothesized that different microbial genomes within the intestinal tract may play a role in the immune response in several inflammatory diseases[41]. Further clinically-relevant studies focused on the human intestinal microbiome and the potential role of PepT1 are needed.

COMMENTS
Background

Assimilation of dietary protein in humans involves gastric and pancreatic enzyme hydrolysis to luminal oligopeptides and free amino acids. During the ensuing intestinal phase, these hydrolytic products are transported into the epithelial cell and, eventually, the portal vein.

Research frontiers

A critical component of this process is the uptake of intact di-peptides and tri-peptides by an independent peptide transporter transmembrane protein, peptide transporter 1 (PepT1). In recent years, a number of “peptide-mimetic” pharmaceutical agents may also be transported through this carrier, important for uptake of different antibiotics, antiviral agents and angiotensin-converting enzyme inhibitors. In addition, specific peptide products of intestinal bacteria may also be transported by PepT1, with initiation and persistence of an immune response including increased cytokine production and associated intestinal inflammatory changes. Interestingly, these inflammatory changes may also be attenuated with orally-administered anti-inflammatory tri-peptides administered as site-specific nanoparticles and taken up by this PepT1 transport protein.

Innovations and breakthroughs

Further evaluation of the role of this transmembrane transport protein, PepT1, in transport of pharmaceutical agents is needed. This may provide novel approaches to treatment, particularly for intestinal disorders. In particular, use of agents that employ this peptide transporter to permit access into intestinal cells may have a special role in inflammatory bowel disease treatment.

Applications

From a practical perspective, use of agents that particularly localize to the intestinal mucosal cells might have an important role in localization of treatment regimens rather than use of current systemically-applied pharmaceutical or biological agents.

Terminology

The PepT1 transporter is a special transmembrane intestinal transport protein located in the microvillus membrane. Its role as a nutrient transporter, specifically for di-peptides and tri-peptides is well established. However, in recent years, its role in uptake of several pharmaceutical agents has become apparent, including its potential relevance for management of inflammatory intestinal disorders.

Peer-review

Conceptually, the peer reviewers have raised the important linkage of this peptide transporter and the modern “metagenomic revolution” that should further our understanding of intestinal, particularly inflammatory, disorders. Added studies are also needed that explore these intestinal uptake processes and role of this PepT1 transporter in the developing human intestinal tract, particularly in fetal and neonatal settings.

Footnotes

P- Reviewer: Actis GC, Catania VA S- Editor: Tian YL L- Editor: A E- Editor: Lu YJ

References
1.  Freeman HJ, Kim YS, Sleisenger MH. Protein digestion and absorption in man. Normal mechanisms and protein-energy malnutrition. Am J Med. 1979;67:1030-1036.  [PubMed]  [DOI]  [Cited in This Article: ]
2.  Thomson AB, Chopra A, Clandinin MT, Freeman H. Recent advances in small bowel diseases: Part I. World J Gastroenterol. 2012;18:3336-3352.  [PubMed]  [DOI]  [Cited in This Article: ]
3.  Thomson AB, Chopra A, Clandinin MT, Freeman H. Recent advances in small bowel diseases: Part II. World J Gastroenterol. 2012;18:3353-3374.  [PubMed]  [DOI]  [Cited in This Article: ]
4.  Chung YC, Kim YS, Shadchehr A, Garrido A, Macgregor IL, Sleisenger MH. Protein digestion and absorption in human small intestine. Gastroenterology. 1979;76:1415-1421.  [PubMed]  [DOI]  [Cited in This Article: ]
5.  Gilbert ER, Wong EA, Webb KE. Board-invited review: Peptide absorption and utilization: Implications for animal nutrition and health. J Anim Sci. 2008;86:2135-2155.  [PubMed]  [DOI]  [Cited in This Article: ]
6.  Kim YS, Birtwhistle W, Kim YW. Peptide hydrolases in the bruch border and soluble fractions of small intestinal mucosa of rat and man. J Clin Invest. 1972;51:1419-1430.  [PubMed]  [DOI]  [Cited in This Article: ]
7.  Morita A, Chung YC, Freeman HJ, Erickson RH, Sleisenger MH, Kim YS. Intestinal assimilation of a proline-containing tetrapeptide. Role of a brush border membrane postproline dipeptidyl aminopeptidase IV. J Clin Invest. 1983;72:610-616.  [PubMed]  [DOI]  [Cited in This Article: ]
8.  Adibi SA, Morse EL, Masilamani SS, Amin PM. Evidence for two different modes of tripeptide disappearance in human intestine. Uptake by peptide carrier systems and hydrolysis by peptide hydrolases. J Clin Invest. 1975;56:1355-1363.  [PubMed]  [DOI]  [Cited in This Article: ]
9.  Terada T, Sawada K, Saito H, Hashimoto Y, Inui K. Functional characteristics of basolateral peptide transporter in the human intestinal cell line Caco-2. Am J Physiol. 1999;276:G1435-G1441.  [PubMed]  [DOI]  [Cited in This Article: ]
10.  Irie M, Terada T, Okuda M, Inui K. Efflux properties of basolateral peptide transporter in human intestinal cell line Caco-2. Pflugers Arch. 2004;449:186-194.  [PubMed]  [DOI]  [Cited in This Article: ]
11.  Daniel H. Molecular and integrative physiology of intestinal peptide transport. Annu Rev Physiol. 2004;66:361-384.  [PubMed]  [DOI]  [Cited in This Article: ]
12.  Daniel H, Kottra G. The proton oligopeptide cotransporter family SLC15 in physiology and pharmacology. Pflugers Arch. 2004;447:610-618.  [PubMed]  [DOI]  [Cited in This Article: ]
13.  Shimakura J, Terada T, Shimada Y, Katsura T, Inui K. The transcription factor Cdx2 regulates the intestine-specific expression of human peptide transporter 1 through functional interaction with Sp1. Biochem Pharmacol. 2006;71:1581-1588.  [PubMed]  [DOI]  [Cited in This Article: ]
14.  Liang R, Fei YJ, Prasad PD, Ramamoorthy S, Han H, Yang-Feng TL, Hediger MA, Ganapathy V, Leibach FH. Human intestinal H+/peptide cotransporter. Cloning, functional expression, and chromosomal localization. J Biol Chem. 1995;270:6456-6463.  [PubMed]  [DOI]  [Cited in This Article: ]
15.  Brodin B, Nielsen CU, Steffansen B, Frøkjaer S. Transport of peptidomimetic drugs by the intestinal Di/tri-peptide transporter, PepT1. Pharmacol Toxicol. 2002;90:285-296.  [PubMed]  [DOI]  [Cited in This Article: ]
16.  Steffansen B, Nielsen CU, Brodin B, Eriksson AH, Andersen R, Frokjaer S. Intestinal solute carriers: an overview of trends and strategies for improving oral drug absorption. Eur J Pharm Sci. 2004;21:3-16.  [PubMed]  [DOI]  [Cited in This Article: ]
17.  Brandsch M. Drug transport via the intestinal peptide transporter PepT1. Curr Opin Pharmacol. 2013;13:881-887.  [PubMed]  [DOI]  [Cited in This Article: ]
18.  Smith DE, Clémençon B, Hediger MA. Proton-coupled oligopeptide transporter family SLC15: physiological, pharmacological and pathological implications. Mol Aspects Med. 2013;34:323-336.  [PubMed]  [DOI]  [Cited in This Article: ]
19.  Newstead S. Molecular insights into proton coupled peptide transport in the PTR family of oligopeptide transporters. Biochim Biophys Acta. 2015;1850:488-499.  [PubMed]  [DOI]  [Cited in This Article: ]
20.  Ferraris RP, Diamond JM. Specific regulation of intestinal nutrient transporters by their dietary substrates. Annu Rev Physiol. 1989;51:125-141.  [PubMed]  [DOI]  [Cited in This Article: ]
21.  Ferraris RP, Carey HV. Intestinal transport during fasting and malnutrition. Annu Rev Nutr. 2000;20:195-219.  [PubMed]  [DOI]  [Cited in This Article: ]
22.  Pan X, Terada T, Irie M, Saito H, Inui K. Diurnal rhythm of H+-peptide cotransporter in rat small intestine. Am J Physiol Gastrointest Liver Physiol. 2002;283:G57-G64.  [PubMed]  [DOI]  [Cited in This Article: ]
23.  Pan X, Terada T, Okuda M, Inui K. Altered diurnal rhythm of intestinal peptide transporter by fasting and its effects on the pharmacokinetics of ceftibuten. J Pharmacol Exp Ther. 2003;307:626-632.  [PubMed]  [DOI]  [Cited in This Article: ]
24.  Pan X, Terada T, Okuda M, Inui K. The diurnal rhythm of the intestinal transporters SGLT1 and PEPT1 is regulated by the feeding conditions in rats. J Nutr. 2004;134:2211-2215.  [PubMed]  [DOI]  [Cited in This Article: ]
25.  Pácha J. Development of intestinal transport function in mammals. Physiol Rev. 2000;80:1633-1667.  [PubMed]  [DOI]  [Cited in This Article: ]
26.  Ford D, Howard A, Hirst BH. Expression of the peptide transporter hPepT1 in human colon: a potential route for colonic protein nitrogen and drug absorption. Histochem Cell Biol. 2003;119:37-43.  [PubMed]  [DOI]  [Cited in This Article: ]
27.  Ziegler TR, Fernández-Estívariz C, Gu LH, Bazargan N, Umeakunne K, Wallace TM, Diaz EE, Rosado KE, Pascal RR, Galloway JR. Distribution of the H+/peptide transporter PepT1 in human intestine: up-regulated expression in the colonic mucosa of patients with short-bowel syndrome. Am J Clin Nutr. 2002;75:922-930.  [PubMed]  [DOI]  [Cited in This Article: ]
28.  Merlin D, Si-Tahar M, Sitaraman SV, Eastburn K, Williams I, Liu X, Hediger MA, Madara JL. Colonic epithelial hPepT1 expression occurs in inflammatory bowel disease: transport of bacterial peptides influences expression of MHC class 1 molecules. Gastroenterology. 2001;120:1666-1679.  [PubMed]  [DOI]  [Cited in This Article: ]
29.  Dalmasso G, Nguyen HT, Charrier-Hisamuddin L, Yan Y, Laroui H, Demoulin B, Sitaraman SV, Merlin D. PepT1 mediates transport of the proinflammatory bacterial tripeptide L-Ala-{gamma}-D-Glu-meso-DAP in intestinal epithelial cells. Am J Physiol Gastrointest Liver Physiol. 2010;299:G687-G696.  [PubMed]  [DOI]  [Cited in This Article: ]
30.  Ingersoll SA, Ayyadurai S, Charania MA, Laroui H, Yan Y, Merlin D. The role and pathophysiological relevance of membrane transporter PepT1 in intestinal inflammation and inflammatory bowel disease. Am J Physiol Gastrointest Liver Physiol. 2012;302:G484-G492.  [PubMed]  [DOI]  [Cited in This Article: ]
31.  Wuensch T, Ullrich S, Schulz S, Chamaillard M, Schaltenberg N, Rath E, Goebel U, Sartor RB, Prager M, Büning C. Colonic expression of the peptide transporter PEPT1 is downregulated during intestinal inflammation and is not required for NOD2-dependent immune activation. Inflamm Bowel Dis. 2014;20:671-684.  [PubMed]  [DOI]  [Cited in This Article: ]
32.  Kanneganti TD, Lamkanfi M, Núñez G. Intracellular NOD-like receptors in host defense and disease. Immunity. 2007;27:549-559.  [PubMed]  [DOI]  [Cited in This Article: ]
33.  Cuthbert AP, Fisher SA, Mirza MM, King K, Hampe J, Croucher PJ, Mascheretti S, Sanderson J, Forbes A, Mansfield J. The contribution of NOD2 gene mutations to the risk and site of disease in inflammatory bowel disease. Gastroenterology. 2002;122:867-874.  [PubMed]  [DOI]  [Cited in This Article: ]
34.  Hugot JP, Chamaillard M, Zouali H, Lesage S, Cézard JP, Belaiche J, Almer S, Tysk C, O’Morain CA, Gassull M. Association of NOD2 leucine-rich repeat variants with susceptibility to Crohn’s disease. Nature. 2001;411:599-603.  [PubMed]  [DOI]  [Cited in This Article: ]
35.  Ogura Y, Bonen DK, Inohara N, Nicolae DL, Chen FF, Ramos R, Britton H, Moran T, Karaliuskas R, Duerr RH. A frameshift mutation in NOD2 associated with susceptibility to Crohn’s disease. Nature. 2001;411:603-606.  [PubMed]  [DOI]  [Cited in This Article: ]
36.  Strober W, Asano N, Fuss I, Kitani A, Watanabe T. Cellular and molecular mechanisms underlying NOD2 risk-associated polymorphisms in Crohn’s disease. Immunol Rev. 2014;260:249-260.  [PubMed]  [DOI]  [Cited in This Article: ]
37.  Zucchelli M, Torkvist L, Bresso F, Halfvarson J, Hellquist A, Anedda F, Assadi G, Lindgren GB, Svanfeldt M, Janson M. PepT1 oligopeptide transporter (SLC15A1) gene polymorphism in inflammatory bowel disease. Inflamm Bowel Dis. 2009;15:1562-1569.  [PubMed]  [DOI]  [Cited in This Article: ]
38.  Dalmasso G, Charrier-Hisamuddin L, Nguyen HT, Yan Y, Sitaraman S, Merlin D. PepT1-mediated tripeptide KPV uptake reduces intestinal inflammation. Gastroenterology. 2008;134:166-178.  [PubMed]  [DOI]  [Cited in This Article: ]
39.  Laroui H, Dalmasso G, Nguyen HT, Yan Y, Sitaraman SV, Merlin D. Drug-loaded nanoparticles targeted to the colon with polysaccharide hydrogel reduce colitis in a mouse model. Gastroenterology. 2010;138:843-853.e1-2.  [PubMed]  [DOI]  [Cited in This Article: ]
40.  Kovacs-Nolan J, Zhang H, Ibuki M, Nakamori T, Yoshiura K, Turner PV, Matsui T, Mine Y. The PepT1-transportable soy tripeptide VPY reduces intestinal inflammation. Biochim Biophys Acta. 2012;1820:1753-1763.  [PubMed]  [DOI]  [Cited in This Article: ]
41.  Proal AD, Albert PJ, Marshall TG. Inflammatory disease and the human microbiome. Discov Med. 2014;17:257-265.  [PubMed]  [DOI]  [Cited in This Article: ]