Skip to main content

Regulation of intestinal epithelial homeostasis by mesenchymal cells

Abstract

The gastrointestinal tract harbors diverse microorganisms in the lumen. Epithelial cells segregate the luminal microorganisms from immune cells in the lamina propria by constructing chemical and physical barriers through the production of various factors to prevent excessive immune responses against microbes. Therefore, perturbations of epithelial integrity are linked to the development of gastrointestinal disorders. Several mesenchymal stromal cell populations, including fibroblasts, myofibroblasts, pericytes, and myocytes, contribute to the establishment and maintenance of epithelial homeostasis in the gut through regulation of the self-renewal, proliferation, and differentiation of intestinal stem cells. Recent studies have revealed alterations in the composition of intestinal mesenchymal stromal cells in patients with inflammatory bowel disease and colorectal cancer. A better understanding of the interplay between mesenchymal stromal cells and epithelial cells associated with intestinal health and diseases will facilitate identification of novel biomarkers and therapeutic targets for gastrointestinal disorders. This review summarizes the key findings obtained to date on the mechanisms by which functionally distinct mesenchymal stromal cells regulate epithelial integrity in intestinal health and diseases at different developmental stages.

Background

The gastrointestinal tract harbors a large number of commensal bacteria, referred to as microbiota, in the lumen, while various types of immune cells are present in the lamina propria. Intestinal epithelial cells form a physical barrier and a chemical barrier, which prevents direct interaction between the microbiota and the immune cells. As the basis for this barrier, a monolayer of intestinal epithelium composed of a variety of differentiated cell types is constantly regenerated from intestinal stem cells located at the crypt bases [1, 2]. The self-renewal, differentiation, and proliferation of intestinal stem cells are precisely manipulated by various secretory factors, such as WNTs, bone morphogenetic proteins (BMPs), and R-spondins. This manipulation is necessary to prevent the excessive growth of intestinal stem cells and the disruption of epithelial tissue regeneration [2].

Intestinal epithelial cells are located above the basement membrane, which is an extracellular matrix compartment [3, 4]. Fibroblasts and myofibroblasts, which are mesenchymal stromal cell populations, play roles in constructing this compartment by secreting various proteins, including collagens, glycoproteins, proteoglycans, and extracellular matrix remodeling-related proteolytic enzymes, such as matrix metalloproteinases (MMPs) [5]. Among these proteins, collagens are involved in the regulation of epithelial elasticity and tensile strength. Meanwhile, glycoproteins and proteoglycans bind to signaling molecules, including amphiregulin, WNTs, BMPs, EGF, and FGF, thereby contributing to epithelial cell survival, proliferation, differentiation, migration, and apoptosis.

Recent advances in single-cell RNA-sequencing (scRNA-seq) and multi-omics technologies have revealed the diversity of mesenchymal stromal cells, such as fibroblast, pericytes, myofibroblasts, and myocyte, within individual tissues, along with the specific functions of these subsets in health and diseases [6,7,8,9,10,11]. Molecular markers for these cells include PDGFRα, PDGFRβ, vimentin, CD90, podoplanin, collagens, and αSMA, while they lack the expression of markers for hematopoietic cells (CD45), endothelial cells (CD31), and epithelial cells (EpCAM). Vimentin+ PDGFRα+ fibroblasts, a major cell type among cells producing components of the extracellular matrix, are involved in tissue/organ development and wound healing [12]. Studies have shown that tissue-resident fibroblasts contribute to the maintenance of hematopoietic stem cells [13, 14] and non-hematopoietic stem cells [15,16,17,18]. Therefore, perturbations of fibroblast physiology lead to the onset and/or progression of tissue fibrosis, cancer, and chronic inflammatory disorders.

Ulcerative colitis and Crohn’s disease, the main clinical forms of inflammatory bowel disease (IBD), are relapsing and chronic inflammatory disorders of the gastrointestinal tracts. Their incidence and prevalence have been increasing globally [19]. Although the etiology of IBD is largely unknown, accumulating evidence has demonstrated that its pathogenesis is associated with the dysfunction of epithelial barrier integrity, dysregulation of innate/adaptive immune responses, microbiota, environmental factors, dietary habits, and genetics [20, 21]. In the intestine of patients with IBD [22,23,24,25,26], as well as those with colorectal cancer [27,28,29,30], there is a change in the composition of mesenchymal stromal cell populations, while immune cell and epithelial cell populations are also altered.

In this review, we provide an overview of the roles of intestinal mesenchymal stromal cells in the establishment and maintenance of epithelial homeostasis in intestinal health and diseases at different developmental stages.

Functionally distinct cell types in adult intestinal epithelium

The epithelial monolayer in the intestine provides the first line of host defense against pathogens. It is composed of diverse cell types [2, 31]: fast-cycling (active) Lgr5+ stem cells and slow-cycling (quiescent) + 4 long-term label-retaining cells (LRCs), which have multipotency and self-renewal ability, proliferating transit-amplifying (TA) cells, secretory lineage cells (i.e., Paneth cells whose distribution appears to be limited to the small intestine, goblet cells, tuft cells, and enteroendocrine cells as well as microfold cells present in Peyer’s patches), and absorptive enterocytes, which constitute more than 80% of all differentiated intestinal epithelial cells (Fig. 1A and B). The small intestine is made up of a huge number of crypt-villus units [32], while the luminal surface of colonic crypts is flat and lacks villi [2, 32]. Lgr5+ stem cells, + 4 LRCs, and Paneth cells are present at the crypt bases of the intestine, while proliferating TA cells are arranged near these bases. The cells generated from stem cells or their progeny near or at the crypt base migrate toward the villus in the small intestine and the top of the crypt in the colon. In this context, they differentiate and mature into multiple cell types, such as absorptive enterocytes, tuft cells, goblet cells, and enteroendocrine cells.

Fig. 1
figure 1

Intestinal epithelial cells and their regulatory signals. A In the small intestine, Lgr5+ stem cells are intercalated between Paneth cells at the crypt base. These cells continuously give rise to proliferating transit-amplifying (TA) cells, which are present at the remainder of the crypts. The + 4 long-term label-retaining cells (LRCs), which reside at the fourth position from the base of crypts, are quiescent under homeostatic conditions, while they are activated and reconstruct the Lgr5+ stem cell compartment during epithelial regeneration after injury. TA cells differentiate into functionally distinct cell types, such as tuft cells, goblet cells, enteroendocrine cells, and absorptive enterocytes. B In the colon, Lgr5+ stem cells localized at the crypt base generate TA cells, which reside in the lower half part of the crypt and differentiate into multiple cell type, including goblet cells, tuft cells, enteroendocrine cells, and absorptive enterocytes. Homeostatic turnover of epithelial cells occurs every 3–5 days in the small intestine and 5–7 days in the colon. C Lgr5+ stem cells constantly replenish functionally mature cells under homeostatic conditions. Fate determination and differentiation of Lgr5+ stem cell progeny excluded from the stem cell niche are precisely manipulated by various secretory molecules. Dll1+ TA (secretory progenitor) cells mature into goblet cells, tuft cells, enteroendocrine cells, and Paneth cells, while Dll1− TA cells supply absorptive enterocytes. Notch signaling promotes enterocyte differentiation and conversely inhibit differentiation of secretory lineage cells by inducing the expression of Hes1. Secretory progenitor cells receiving WNT signal through Eph receptors return to the crypt base and mature into Paneth cells. Meanwhile, secretory progenitor cells differentiate into goblet cells or enteroendocrine lineage cells when they are not supplied with sufficient WNT proteins. Th2 cytokines IL-4 and IL-13 drive tuft cell differentiation. In the epithelium of Peyer’s patches, RANKL derived from mesenchymal cells elicits microfold cell (M cell) differentiation through induction of transcription factor SpiB expression

The intestinal tract harbors multiple microorganisms, including bacteria, viruses, and fungi, as commensals in the lumen [33,34,35,36,37], while adaptive immune cells, IFN-γ-producing T-helper 1 (Th1) cells, IL-17-producing Th17 cells, Foxp3+ regulatory T cells, IgA-producing plasma cells, and B cells and innate immune cells, macrophages, eosinophils, neutrophils, dendritic cells, and innate lymphoid cells (ILCs) reside in the lamina propria [38,39,40]. To segregate the microorganisms from the immune cells, a mucus layer organized by MUC2 secreted by goblet cells [41] serves as a physical barrier in the colon, whereas antimicrobial peptides (e.g., defensins and cathelicidins) and chemical barrier molecules (e.g., Reg3 family proteins) are secreted by enterocytes [42] and Paneth cells [43] in the small intestine. Patients with IBD show exacerbated immune responses in the intestinal mucosa accompanied by perturbations of the epithelial barrier [44,45,46]. These findings indicate the importance of the precise regulation of epithelial barrier integrity in the intestine to avoid the excessive activation of mucosal immune cells linked to the development of IBD.

The signaling pathways responsible for the establishment and maintenance of epithelial homeostasis

Approximately, 1011 epithelial cells are lost in the human intestine each day [47]. Thus, to maintain homeostasis, Lgr5+ intestinal stem cells need to constantly replenish functionally mature cells. A large body of evidence suggests that the balance among quiescence, proliferation, cell fate decision, and differentiation of intestinal stem cells and their progeny is precisely manipulated by various secretory molecules, such as WNTs, the WNT antagonists DKKs, R-spondins (Rspo1-4), Notch ligands, EGFs, BMPs, and the BMP inhibitors Noggin and Gremlins, in order to maintain homeostasis of intestinal epithelium [2, 31, 32, 48] (Fig. 1C).

WNT signaling

WNT canonical signaling initiates gene transcription by inducing the formation of a complex of β-catenin with TCF/LEF family protein [49, 50] through a frizzled receptor (frizzled1–10) and its co-receptor LRP (LRP5 or LRP6) [51]. This is essential for crypt maintenance by promoting stem cell proliferation as well as maintaining the undifferentiated status of these cells. The WNT signaling pathway induces the expression of Eph receptors in intestinal epithelium [52], including EphB2 and EphB3. Genetic ablation of EphB2 or EphB3 has been shown to lead to the mislocalization of Paneth cells and proliferating cells [52, 53], indicating that WNT signaling ensures the proper localization of stem cells, TA cells, and Paneth cells near or at the crypt base by inducing the expression of EphB receptors. R-spondins produced from endothelial cells and mesenchymal stromal cells [15, 54,55,56,57] are recruited by the receptor LGR5 and enhance WNT signaling, which ensures intestinal stem cell homeostasis and facilitates epithelial regeneration [56,57,58,59,60,61,62]. Lymphatic endothelial cells close to the crypt bottom produce Rspo3, Wnt2, and the extracellular matrix protein reelin [56, 57]. Notably, receptors for reelin, such as VLDLR [57, 63, 64], ApoER2 [57, 64, 65], α3β1-integrin receptor [63, 64, 66], and EphB2 receptor [66, 67], are expressed in myofibroblasts, crypt cells, and enterocytes in mice. It has been shown that the reelin promotes phosphorylation of Disabled-1 in murine intestinal epithelial cells via its receptor [57, 64]. In addition, lack of reelin signaling leads to an increase in the number of proliferating cells within crypt niches [57] and reduction of matured goblet cells [65] in the murine intestine, indicating that reelin plays essential roles in the self-renewal and differentiation of intestinal stem cells.

EGF signaling

EGF produced by Paneth cells and mesenchymal cells stimulates stem cell proliferation through its receptors [68]. Among the four EGF receptors (ErbB1–4), the expression of ErbB1 is restricted to stem cells and TA cells within the crypt of the murine small intestine [69]. To successfully culture intestinal organoids, EGF receptor ligand supplementation is required [70], suggesting that non-epithelial cells elicit sufficient activation of the EGF signaling pathway in intestinal stem cells. In addition to ErbB family members, intestinal stem cells express the negative regulator of ErbB signaling Lrig1 [71]. It has been reported that Lrig1 deficiency in mice facilitates crypt proliferation and increases the numbers of intestinal stem cells and Paneth cells in the small intestine, which are restored to near-normal levels by treatment with ErbB inhibitor [71]. These findings suggest that the EGF-ErbB signaling pathway is tightly regulated to maintain stem cell homeostasis in the intestine.

Notch signaling

The Notch ligands Dll1 and Dll4 are secreted from Paneth cells [31]. Notch1 and Notch2 signaling pathways contribute to the first fate decision of intestinal stem cells. These pathways inhibit differentiation into secretory cells by inducing the expression of Hes1, which suppresses transcription of a master regulator gene of secretory lineage cells, Atoh1 (also known as Math1) [72,73,74,75], and directs the differentiation of TA cells into absorptive progenitors and ultimately the development of enterocytes.

BMP signaling

BMPs are mainly produced by PDGFRαhigh cells [15, 76, 77] or Col6a1+ CD201+ CD34− cells [78] in the crypt top of the colon and the villus of the small intestine. These proteins restrict intestinal stem cell proliferation as well as self-renewal. It has been reported that a complex of Smad1 with Smad4, the formation of which is induced by BMP signaling, constrains the stemness of intestinal stem cells by suppressing the expression of stem cell signature genes through recruiting the histone deacetylase HDAC1 [79]. This in turn leads to the formation of villus tip and crypt top in the small and large intestines in mice, respectively. A BMP gradient along the crypt-villus axis restrains cell proliferation and permits broad zonation of goblet cells, enteroendocrine cells, and enterocytes by switching their gene expression patterns [78, 80,81,82,83,84]. Meanwhile, to maintain stem cell properties, the BMP antagonists Gremlins, Noggin, and Chordins are produced by PDGFRαlow cells surrounding the crypt bottom [15, 17, 62, 76, 78, 85, 86].

Intestinal stem cells and their progeny pick up signals related to regulation of the WNT and BMP pathways provided by secretory cells in the epithelium [2, 31], endothelial cells [62], glial cells [87], and PDGFRα+ fibroblast subsets, including PDGFRαhigh Foxl1+ telocytes [16, 88], PDGFRαlow GREM1+ trophocytes [15], CD81− PDGFRαlow peri-cryptal cells [16], LTβR+ PDGFRαhigh cells [89], GLI1+ cells [17], and pericyte-like cells [90]. Among secretory cells, Paneth cells produce Wnt3, Dll1, Dll4, EGF, and Noggin to maintain epithelial stemness in the small intestine, while enteroendocrine cells secrete Wnt3, Dll1, and Dll4 [31]. However, previous studies showed that a lack of secretory lineage cells caused by Atoh1 deficiency [91, 92] as well as epithelial cell-specific deletion of Wnt3a [93] does not affect intestinal stem cell physiology in vivo but does in vitro. Additionally, studies have shown that the main source of secretory factors related to WNT and BMP signaling is non-epithelial cells in the colon [31, 94]. These findings demonstrate that mesenchymal stromal cells, such as fibroblasts, are essential for the maintenance of epithelial stem cells and their differentiation into several functionally distinct cell types in the intestine.

The interaction between mesenchymal stromal cells and epithelial cells during development of the intestine

Epithelial cells communicate bidirectionally with mesenchymal stromal cells in the developing intestine (Fig. 2). Endodermal pseudostratified epithelium and splanchnic mesoderm-derived mesenchymal cells beneath the epithelium proliferate vigorously from embryonic day 9.5 (E9.5) to E14.5 in mice (corresponding to weeks 4 to 9 of gestation in humans), which enables elongation of the gut tube at this stage [95]. The entire pseudostratified epithelium in the murine fetal intestine proliferates and expresses stem cell markers, such as Lgr5, CD44, Sox9, and Axin2 [96, 97], although proliferative cells are present with an apparently random distribution within the pseudostratified epithelium in the human fetal intestine [98]. A murine study showed the expansion of fibroblast/myofibroblast progenitors with the expression of Mki67, Dek, Cdk1, Aurkb, and Hmmr in the intestine at E14.5 [99], which might be identical to proliferative mesenchymal cells residing in the human fetal intestine [100]. During the same period, the pseudostratified epithelium in the small intestine begins to form villi and construct columnar epithelium [94, 101, 102]. Villus morphogenesis is strongly dependent on the interplay between PDGFRα+ cells and the epithelium [94, 95, 101]. The epithelium produces the PDGF ligand PDGF-A and hedgehog (Hh) proteins, such as sonic hedgehog (Shh) and Indian hedgehog, and elicits the migration and clustering of mesenchymal cells that highly express Pdgfra, the Shh receptor Ptch1 (Ptc1), and the transcription factor Gli1, underneath the sites where villi subsequently emerge [97, 103,104,105,106]. It has been reported that a lack of Hh signaling [106] or PDGF/PDGFR signaling [103] perturbs villus formation and also impairs the proliferation of mesenchymal progenitors at the embryonic stage in mice. The activation of Hh signaling in PDGFRα+ cells directs the Gli2-dependent transcription of genes related to the planar cell polarity (PCP) pathway [107], such as cadherins Fat4 and Dchs1, Wnt5a, and PCP signaling core factors Vangl1 and Vangl2 [108]. Both Fat4-Dchs1 signaling and Wnt5a-Vangl2 signaling are essential for the cluster formation of mesenchymal stromal cells and proper vilification in the murine intestine. Therefore, perturbations of these signaling pathways cause abnormal morphogenesis of villi in mice at E15.5, such as shortened and fused villi. The transcription factor Foxl1, which is specifically expressed in intestinal PDGFRα+ cells, is encoded by one of the Hh-Gli2 axis-dependent genes [109]. Defect of the Foxl1 gene leads to delayed villus formation and a decreased number of villi in mice at E14.5 to E16.5 [110], suggesting that the activation of Foxl1-dependent transcription in PDGFRα+ cells via the Hh signaling pathway is crucial for epithelial remodeling to form villi in the small intestine.

Fig. 2
figure 2

The interplay between epithelial cells and mesenchymal cells in the developing intestine. A In the murine fetal intestine, pseudostratified epithelium that expresses stem cell markers (e.g., Lgr5, CD44, Sox9, and Axin2) and mesenchymal cells beneath the epithelium proliferate vigorously from E9.5 to E14.5. The epithelium produces PDGF-A, WNT proteins, and hedgehog (Hh) proteins and elicits the migration and clustering of mesenchymal stromal cells expressing Pdgfra, the Hh receptor Ptch1, and the transcription factors Gli1 and Gli2. The clustering of mesenchymal cells results in inhibition of cell proliferation in the above epithelial cells. Hh signaling from the epithelium contributes to the continued clustering of mesenchymal cells through induction of the Gli2-dependent transcription of genes related to PCP pathway, which supports vilification. Around E16.5, mesenchymal stromal cells surrounding the villus tip where the shh proteins secreted from epithelial cells are concentrated drive epithelial cell differentiation by producing BMPs, which confines intestinal stem cells to the base of forming villi. Crypt formation in the murine intestine occurs after birth. B Maturation of LTβR+ PDGFRαhigh mesenchymal stromal cells that produce BMPs, retinoic acids (RA), collagens, and laminins is induced by epithelial cell-derived PDGF-A and PDGF-C signals at the early stage after birth in the murine intestine, which is essential for reinforcement of the epithelial barrier and regulation of intestinal immune responses in a steady state as well as during tissue regeneration after injury. LTβR+ PDGFRαhigh mesenchymal stromal cell-mediated accumulation of CD103+ dendritic cells (DCs) is required for induction of Foxp3+ regulatory T-cell development in the lamina propria. In the murine postnatal intestine, CD206+ macrophages, which are differentiated from monocytes in a microbiota-dependent fashion, facilitate the proliferation of pericyte marker Ng2+ mesenchymal cells through Wnt6 production. Ng2+ mesenchymal cells secrete BMP inhibitor Gremlin 1 and induce functionally mature Paneth cells, which is required for prevention of neonatal necrotizing enterocolitis-like pathology in the murine intestine

In the late embryonic developmental stage (around E16.5) in mice, the clustering Ptc1+ PDGFRα+ cells surrounding the villus tip where epithelium-derived Shh proteins are concentrated drive epithelial cell differentiation by producing BMPs through activation of the Hh signaling pathway [97, 105], which confines intestinal stem cells to the base of forming villi. In human fetal intestine, proliferating cells are restricted to the crypt base and intervillus domain by gestational week 12 [98]. Morphogenesis and maturation of human intestinal epithelium are completed by birth, whereas these processes continue into the neonatal stage in mice, and the mature intestinal epithelium is established by postnatal day 28 [111]. In human fetal small intestine from weeks 7 to 21 of gestation, mesenchymal cells residing in the muscularis mucosa, which is close to the base of crypts, produce WNTs and R-spondins to maintain intestinal stem cells, while PDGFRαhigh F3high DLL1high subepithelial mesenchymal cells along the crypt-villus axis promote secretory lineage cell differentiation by producing the EGF family member Neuregulin1 [112]. Similarly, Neuregulin1 is secreted from PDGFRαhigh cells in the villus and crypt top of the murine small intestine [113] and large intestine [76], which is elevated during epithelial regeneration following irradiation-mediated injury [113]. Furthermore, perivascular PDGFRαhigh cells generated from postnatal LTβR+ cells after birth in mice [89] induce maturation of the epithelial barrier and immune system by producing BMPs, retinoic acids, collagens, and laminins in response to epithelial PDGF-A and PDGF-C [89, 114].

Neonatal necrotizing enterocolitis (NEC) is a serious gastrointestinal disorder with a high mortality rate in premature infants [115, 116]. Several lines of evidence have shown that antibiotic use and consequent dysbiosis, such as expansion of Enterobacteriaceae, are associated with the incidence and pathogenesis of NEC [117,118,119,120,121]. A study demonstrated that antibiotic treatment-induced dysbiosis in murine neonates exacerbates NEC-like pathogenesis accompanied by a reduction in Paneth cells [90]. This study also showed that the colonization of Lactobacillus rhamnosus induces Paneth cell differentiation and partially improves NEC-like intestinal pathology. Heat-killed Lactobacillus promoted the differentiation of monocytes into CD206+ macrophages, and Wnt6 derived from CD206+ macrophages promoted the proliferation of mesenchymal stromal cells that express mRNA of pericyte marker Ng2 with Pdgfra, Foxl1, Grem1, and Cd81 in vitro. In intestinal epithelial organoid culture, CD206+ macrophages and Ng2+ mesenchymal stromal cells induced the maturation of Paneth cells. This suggests that interaction among microbiota, immune cells, mesenchymal stromal cells, and epithelial cells during early postnatal development is important for ensuring epithelial barrier integrity, leading to the prevention of intestinal diseases. Taken together, these findings indicate that the bidirectional association between epithelial cells and mesenchymal stromal cells is needed for proper intestinal development and the consequent maintenance of gut homeostasis.

Heterogeneity of PDGFRα high fibroblasts and their functions in the maintenance and differentiation of intestinal stem cells

A recent study [18] showed that large and small intestinal PDGFRα+ fibroblasts are divided into four subsets, including PDGFRαhigh Foxl1+ cells, PDGFRαlow CD81+ CD34high ACKR4+ trophocytes, PDGFRαlow Fgfr2+ CD34− cells, and CD81− CD34high Igfbp5+ (small intestine)/CD90+ (large intestine), in adult mice (Fig. 3). It also predicted that PDGFRαhigh Foxl1+ cells originate from PDGFRαlow Fgfr2+ CD34− cells. Additionally, this study identified three PDGFRαhigh Foxl1+ subsets: Neuregulin1+ CD9high CD141− cells, CXCL12+ αSMA+ CD9low CD141+ cells, and Adamdec1+ Wnt4a+ αSMA+ CD141intermediate cells. PDGFRαhigh Foxl1+ cells present in the developed intestine, which are referred to as both telocytes and αSMA-expressing myofibroblasts, are one of the most investigated cell types of the intestinal mesenchymal niche [16, 77, 88, 114, 122,123,124,125,126,127,128] (Fig. 4). Telocytes are characterized by thin cytoplasmic extensions and are present beneath epithelial cells along the crypt-villus axis in the small intestine, as well as the colonic crypt axis, in a steady state [77, 88, 124]. A study showed that telocytes residing in the crypt base of the murine intestine express the canonical WNT ligand Wnt2b, WNT signaling agonist R-spondin 3, and WNT inhibitor Sfrp1, whereas the cells that express the noncanonical WNT ligands Wnt5a and Bmp5 and the canonical WNT signaling antagonists Dkk2 and Dkk3 are enriched toward the villus tip in the small intestine [88]. Another study identified that Lgr5+ telocytes in the villus tip secrete Wnt5a, Rspo3, BMP2, and BMP4 [77]. Ablation of Lgr5+ telocytes in mice suppresses the expression of enterocyte genes in the region toward the villus tip, including Ada, Nt5e, Slc28a2, Klf4, Cdh1, and Egfr [77]. This suggests that villus tip telocytes are essential for the terminal differentiation of enterocytes.

Fig. 3
figure 3

The subsets of PDGFRα+ fibroblast in the intestine. PDGFRα+ fibroblasts comprise four subsets, including PDGFRαhigh Foxl1+ telocytes, PDGFRαlow CD34high CD81+ACKR4+ trophocytes, PDGFRαlow CD34high CD81− Igfbp5+ (small intestine) or CD90+ (large intestine), and PDGFRαlow CD34− Fgfr2+ cells, in the intestine of adult mice. Trophocytes lying beneath the muscularis mucosae activate WNT signaling in intestinal stem cells through producing WNT signaling agonist R-spondin 1–3, the canonical WNT ligand Wnt2b, and the BMP inhibitor Gremlin 1, thereby maintaining stemness of intestinal stem cells. CD81− CD55low cells secrete noncanonical WNT ligand Wnt4 in the lamina propria near or at the crypt top. Meanwhile, peri-cryptal CD81low CD55high cells produce intestinal stem cell niche factors, such as R-spondins, WNTs, Gremlins, and Noggin. Telocytes, which are predicted to originate from PDGFRαlow CD34− Fgfr2+ cells, contain three subsets: CD9high CD141− cells with expression of Neuregulin1, αSMA+ CD9low CD141+ cells with high expression of CXCL12, and αSMA+ CD141intermediate cells expressing Wnt4. These cells are present beneath epithelial cells and contribute to the self-renewal and proliferation of intestinal stem cells as well as differentiation of epithelial cells through modulation of the WNT and BMP signaling pathways by producing various molecules, including WNTs, Rspo3, the canonical WNT signaling antagonists Dkks, WNT inhibitor Sfrp1, BMPs, and MMPs

Fig. 4
figure 4

Heterogeneity of PDGFRαhigh Foxl1+ telocytes in the intestine. PDGFRαhigh Foxl1+ telocytes are located beneath epithelial cells along the crypt-villus axis in the small intestine. These cells compartmentalize production of signaling molecules. At the crypt bases, telocytes function as one of key components of intestinal stem cell niche through secretion of WNT signaling amplifier Rspo3 and canonical WNT ligand Wnt2b, and they also produce WNT antagonist Sfrp1. Meanwhile, telocytes distributed in the upper portion of the crypt as well as villus produce the noncanonical WNT ligand Wnt5a, BMP5, and the canonical WNT signaling antagonists Dkk2 and Dkk3. In the villus tip, Lgr5+ telocytes supply Wnt5a, Rspo3, BMP2, and BMP4 and contribute to terminal differentiation of enterocytes that express Ada, Nt5e, Slc28a2, Klf4, Cdh1, and Egfr

Foxl1-cre; Wnt5aflox/flox mice exhibit an increased number of apoptotic epithelial cells in the fetal intestine and a short small intestine at birth [123], which might provide some insight into the mechanism underlying the development of short bowel syndrome [129]. Additionally, Foxl1-creERT2; Porcnflox/Y mice, in which Wnt secretion is ablated in Foxl1-expressing cells specifically following tamoxifen treatment, show abnormal epithelial architecture in the small and large intestines accompanied by declines in epithelial stem and TA cell proliferation [88]. Similarly, in Foxl1-hDTR mice, depletion of Foxl1-expressing cells by diphtheria toxin results in the disruption of epithelial tissue, such as shorter villi, reduced numbers of Ki67+ proliferating cells and Olfm4+ stem cells, and crypt collapse in the small intestine [124]. These findings indicate that telocytes function as the cellular source of Wnt4, Wnt5a, and Wnt2b surrounding the stem cell compartment, which is linked to the maintenance of epithelial homeostasis. In the small intestinal crypts of Foxl1−/− mice, the number of Ki67+ proliferating cells is increased [110], while the expression of BMP2 and BMP4 is decreased in mesenchymal stromal cells. Additionally, lack of the Foxl1 gene in APCMin/+ mice was reported to lead to intestinal tumor progression [126]. A murine study also showed that Foxl1 transcriptionally upregulates the expression of BMP4 in telocytes [130]. These findings suggest that Foxl1-dependent transcription of the genes encoding BMPs in telocytes is essential for the regulation of intestinal epithelial cell proliferation, which might be associated with prevention of the intestinal tumorigenesis.

In the human [22] and murine [18] intestines, mesenchymal cells express the chemokine CXCL12. A recent study showed that the transcription factors FOXC1 and FOXC2 elicit CXCL12 expression in blood vascular endothelial cells adjacent to small intestinal crypts after ischemia–reperfusion injury in mice, which induces Rspo3 expression via CXCR4 in lymphatic endothelial cells and thereby promotes tissue regeneration through activation of the WNT signaling pathway in intestinal stem cells [131]. It has been reported that the CXCL12-CXCR4 pathway is involved in immune cell differentiation [13, 14, 132,133,134,135,136] and migration [137,138,139,140,141]. In addition to immune cells, CXCR4 is expressed in intestinal epithelial cells [142]. We found that telocytes without the expression of FOXC1 and FOXC2 highly express CXCL12 in the large and small intestines, and that Cxcl12 deficiency in Foxl1-expressing cells results in the progression of intestinal tumorigenesis in APCMin/+ mice (unpublished). Taking these findings together, the induction of functional telocytes is essential for the prevention of intestinal neoplasia as well as the maintenance of epithelial homeostasis.

The function of PDGFRα low populations as a key component of the intestinal stem cell niche

PDFGRαlow Gli1+ CD90+ fibroblasts include CD34high CD81+ ACKR4+ trophocytes and CD34high CD81− cells [15,16,17,18] (Fig. 5). In mice, trophocytes lying beneath the muscularis mucosae secrete Rspo1, Rspo2, Rspo3, Wnt2b, and the BMP inhibitor Gremlin 1, and thus, they are alone able to maintain the growth of intestinal crypt organoids [15, 16, 114]. In addition to trophocytes, muscularis mucosa cells supply Rspo3, Gremlin 1, and Gremlin 2 in human and murine intestine at the postnatal stage [114]. Diphtheria toxin-induced ablation of smooth muscle in the murine intestine at postnatal day 14, which leads to the loss of muscularis mucosae and muscularis propria containing Noggin-producing cells residing underneath trophocytes, is associated with impaired crypt fission and a reduced number of intestinal stem cells through the facilitation of BMP signaling [114]. In this period, trophocytes do not support the growth of small intestinal crypt organoids without growth factors, unlike adult cells. These findings suggest that the layered structure composed of muscularis mucosa cells, trophocytes, and muscularis lamina propria cells becomes established as a complex stem cell niche in the developing intestine.

Fig. 5
figure 5

The diversity of PDGFRαlow fibroblasts in the intestine. PDFGRαlow Gli1+ CD90+ fibroblasts in the intestine can be divided into CD34high CD81high ACKR4high trophocytes and CD34high CD81low subset. The latter comprises CD55low cells and CD55high cells. PDFGRαlow CD55low cells localize in the lamina propria near or at the crypt top and produce noncanonical WNT ligand Wnt4. BMP signaling from telocytes facilitates expression of Wnt4 in PDGFRαlow CD81− cells via the BMP receptors Bmpr1 and Bmpr2, while it inhibits the expression of intestinal stem cell niche factors Rspo3 and Grem1. Trophocytes, CD81low CD55high cells, muscularis mucosae cells, and superficial muscle cells in the muscularis propria cells secrete intestinal stem cell niche factors, including R-spondins, WNTs, Gremlins, and Noggin. In addition, Gli1+ CD90+ peri-cryptal mesenchymal cells produce Sema3 as well as niche factors and accelerate intestinal stem cell proliferation via the receptor Nrp2. Lymphatic endothelial cells reside close to Gremlin 1+ Rspo3+ mesenchymal cells at the crypt base in the intestines, and they produce Rspo3. Gremlin 1+ Rspo3+ mesenchymal cells maintain lymphatic endothelial cells by secreting vascular endothelial cell growth factor D (VEGF-D) and IGF-1 under homeostatic conditions, and they induce expansion of lymphatic endothelial cells by producing VEGF-D, IGF-1, FGF2, IL-6, ANGPTL4, and epiregulin after injury, which support the maintenance of epithelial homeostasis

The PDFGRαlow CD34high CD81− population comprises CD55low cells and CD55high cells in the colon [16]. PDFGRαlow CD55low cells localize in the lamina propria near or at the crypt top and highly express Wnt4 and the BMP-inducible gene Id1. Peri-cryptal PDFGRαlow CD55high cells exhibit the upregulated mRNA expression of Grem1, Rspo3, and Wnt2b. BMP signaling derived from telocytes distributed in the upper portion of the colonic crypt suppresses the expression of intestinal stem cell niche factors Rspo3 and Gremlin 1 and facilitates the expression of Wnt4 in PDGFRαlow CD81− cells with expression of the BMP receptors Bmpr1 and Bmpr2 [16], suggesting that the heterogeneity of PDGFRαlow CD81− cells is determined by their distance from telocytes, but not factors intrinsic to these cells. Diphtheria toxin-mediated elimination of Grem1+ cells results in the development of edema-like dilation in the small intestine accompanied by reduced numbers of intestinal stem cells and Paneth cells, increased numbers of TA cells and goblet cells, and inflammatory cell infiltration due to the loss of a low-BMP environment at the crypt base [15]. In addition to signaling molecules related to the WNT and BMP pathways, Sema3 secreted from PDFGRαlow Gli1+ CD90+ cells surrounding the crypt base promotes intestinal stem cell proliferation via the receptor Nrp2 [143]. Prostaglandin E2 (PGE2) secreted from peri-cryptal PDFGRαlow Fgfr2+ Rspo1+ cells induces Sca-1+ epithelial stem cells with a tumorigenic program through YAP/TAZ activation via the receptor Ptger4, which is linked to early tumor progression [144].

PDFGRαlow cells modulate immune cell physiology. Intestinal myofibroblasts, which are characterized as PDFGRαlow αSMAhigh cells [22], suppress CD4+ T-cell proliferation through the TLR4-dependent induction of PD-L1 expression [145, 146]. PDFGRαlow CD34+ cells transiently increase expression of the adhesion molecule VCAM-1 as well as the inflammatory mediators CCL2 and Csf1 at the acute phase of dextran sodium sulfate-induced colitis in mice, which attracts immune cells to the mucosa [86]. During colitis, the expression of lymphocyte survival/proliferation factor IL-7 and chemokine CCL19 persists in PDFGRαlow CD34+ cells at the chronic phase, which contributes to induction of the prolonged survival of T cells [86]. Although the impacts of PDGFRαlow cell reprogramming on gut homeostasis and disease pathophysiology require further evaluation, the coordination of PDGFRαlow CD34+ cell populations with distinct functions and distributions as well as PDGFRαhigh fibroblasts might be required for epithelial homeostasis and regeneration through maintenance of the intestinal stem cell niche.

The roles of fibroblasts in IBD pathogenesis

Damage to the epithelium drives a complex process of degradation and biosynthesis of the extracellular matrix to regenerate the epithelial layer under physiological conditions [147,148,149]. In the gut of patients with IBD, epithelial regenerative ability as well as the epithelial barrier is impaired [150,151,152], which accelerates inflammatory responses through activating the uptake of luminal antigens by dendritic cells and macrophages in the mucosa. The migratory potential of fibroblasts in the colonic mucosa is known to be reduced in patients with Crohn’s disease and ulcerative colitis [153, 154]. However, proliferation and collagen secretion in intestinal fibroblasts are facilitated in IBD patients [155]. A study showed that Grem1+ Grem2+ myofibroblasts highly express COL18A1 and COL23A1 mRNA in the ileum of patients with Crohn’s disease [156]. Myofibroblast reprogramming into Grem1+ Grem2+ cells is induced by CHMP1A, TBX3, and RNF168 [156], which might contribute not only to wound healing but also fibrosis in Crohn’s disease [157]. Imbalanced MMP proteolytic activity in the extracellular matrix is associated with failure of wound healing through an excess of extracellular matrix deposition. MMP-2 and MMP-9 promote fragmentation of the extracellular matrix and thereby impair reepithelialization [158]. An increased concentration of MMP-9 is observed in the plasma from patients with active ulcerative colitis [159] and active Crohn’s disease [160]. Additionally, elevated concentrations of MMP-2 and MMP-9 in the plasma positively correlate with high intestinal permeability in ulcerative colitis patients [161]. The expression of MMP-7, which increases epithelial barrier permeability through posttranscriptional downregulation of Claudin 7 in vitro, is upregulated in the colon of ulcerative colitis patients [162]. MMP-13, which promotes goblet cell loss, destabilizes tight junctions, and increases intestinal permeability in mice [163], is elevated at the protein level in the intestinal mucosa of patients with ulcerative colitis and Crohn’s disease [164]. Meanwhile, TNF-α stimulation drives the expression of MMP-1, MMP-3, and TIMP1 and consequently initiates the production of type I and IV collagens in myofibroblasts from the murine [165, 166] and human intestines [167]. TNF∆ARE/+ mice, in which TNF-α is constitutively overproduced, exhibit the spontaneous development of Crohn’s disease-like ileitis [168]. Additionally, ColVI-cre; TnfRIfloxneo/floxneoTNF∆ARE/+ mice, in which TNF receptor 1 is specifically expressed in intestinal mesenchymal stromal cells, spontaneously develop ileal inflammation [169] while TNF∆ARE/+ mice with constitutive knockout of TnfRI do not. Intriguingly, treatment with an acellular extracellular matrix derived from the porcine small intestine was shown to effectively close anorectal fistulas in patients with Crohn’s disease [170, 171] and promote cecal wound healing in rats [172]. Furthermore, ECM hydrogel therapy was found to ameliorate dextran sodium sulfate-induced colitis in rats through the inhibition of TNF-α production by macrophages [173]. These findings indicate that intestinal mesenchymal stromal cells are one of the target cell types for TNF-α in the onset and/or progression of IBD.

Toll-like receptors (TLRs) recognize pathogen-associated molecular patterns, which are conserved structures on the surface of microbes, and damage-associated molecular patterns [174]. In addition to epithelial cells and immune cells, human CD90+ fibroblasts [175] and murine fibroblasts [176] in the intestine also express TLRs. Activation of the TLR-adaptor molecule MyD88/TRIF signaling pathway initiates the expression of pro-inflammatory mediators. MAP3K8, known as Tpl2, is activated through the TLR, IL-1R, and TNFR signaling pathways. It has been shown that Tpl2 kinase-mediated activation of the Cox-2/PGE2 pathway in intestinal myofibroblasts, which is initiated by macrophage-derived IL-1β as well as TLR4 ligand lipopolysaccharide, is essential for homeostatic responses in the epithelium through promoting epithelial cell proliferation in mice during dextran sodium sulfate-induced colitis [177]. In ileal myofibroblasts from Crohn’s disease patients, the expression level of MAP3K8 mRNA was shown to be reduced [177]. Additionally, COX2-expressing CD45− CD44+ mesenchymal stromal cells were found to migrate to wound sites of the epithelium via a MyD88-dependent mechanism in mice following dextran sodium sulfate administration and accelerate epithelial tissue regeneration [178]. Furthermore, tissue-resident Vimentin+ mesenchymal stromal cells secrete PGE2 via the TLR4-Tpl2-COX2 pathway in the propagation phase of dextran sodium sulfate-induced colitis in mice and alleviate pathology of the large intestine [179]. These findings demonstrate that intestinal mesenchymal stromal cells are involved in the maintenance of gut homeostasis and tissue regeneration as well as the pathogenesis of IBD.

Excessive accumulation of immune cells is known to occur in the mucosa of patients with IBD [23, 24, 156, 180, 181]. These infiltrating immune cells produce various pro-inflammatory cytokines, including TNF-α, IL-6, IL-1, IL-23p19, IL-12p40, IFN-γ, and IL-17 [23], which initiates tissue damage and chronic inflammation. Therefore, drugs targeting these cytokines and their downstream signaling molecules (e.g., the JAK family members) are clinically approved or being evaluated for IBD therapy [182, 183]. In patients with IBD, the concentration of TNF-α is elevated in the plasma and the intestinal mucosa. TNF-α produced by various types of cells, including immune cells, adipocytes, fibroblasts, and endothelial cells, promote inflammation through chemotaxis [184, 185], proliferation and activation of immune cells [186,187,188,189], degradation of the extracellular matrix [166], and angiogenesis [190] in the intestine. The fact that therapies with anti-TNF antibodies (infliximab and adalimumab, approved for ulcerative colitis and Crohn’s disease; certolizumab pegol, approved for Crohn’s disease; and golimumab, approved for ulcerative colitis) are effective for IBD underscores that TNF plays a crucial role in the progression of this disease [191]. A previous study demonstrated a positive correlation between the numbers of apoptotic mucosal cells and the therapeutic efficacy of infliximab in patients with active Crohn’s disease [192]. Accordingly, the anti-TNF agents infliximab, adalimumab, and certolizumab pegol induce the apoptosis of TNFR2-expressing T cells isolated from IBD patients by binding to transmembrane TNF-α (mTNF-α)-expressing CD14+ macrophages [193]. These findings suggest that the induction of T-cell apoptosis through inhibition of the TNFR2-mTNF-α pathway is one of the mechanisms by which anti-TNF agents control intestinal pathology in patients with IBD. Although the use of anti-TNF agents improves the quality of life of IBD patients by promoting mucosal healing [194], 10–40% of IBD patients do not respond to primary anti-TNF therapy [195,196,197,198]; moreover, a subset of responders lose their responsiveness to anti-TNF treatment [199,200,201]. One study showed that expression of the cytokine oncostatin M is increased in inflamed sites of the mucosa of patients with Crohn’s disease and ulcerative colitis; it also revealed that a high level of oncostatin M expression prior to treatment is associated with the failure of anti-TNF therapy [202]. CD45− CD31− EpCAM− mesenchymal stromal cells in the mucosa of patients with IBD express the receptor for oncostatin M (OSMR) and produce pro-inflammatory cytokines, such as IL-1α, IL-1β, IL-6, and IL-17A, and chemokines in response to oncostatin M [202]. In patients with ulcerative colitis, anti-TNF resistance signature is enriched in IL-24+ IL-13A2+ IL-11+ inflammation-associated fibroblasts, monocytes, and type2 conventional dendritic cells [23]. Intriguingly, inflammation-associated fibroblasts highly express OSMR, while inflammatory monocytes express oncostatin M. Similarly, activated mononuclear phagocytes that produce oncostatin M, TNF-α, IL-1α, and IL-1β in inflamed sites of the ileal mucosa of Crohn’s disease patients [24] are associated with resistance to anti-TNF therapy. These findings suggest that cross talk between monocytes and fibroblasts via the oncostatin M-OSMR axis is involved in resistance to anti-TNF therapy in IBD patients.

The roles of fibroblasts in the intestinal tumor environment

In human lung cancer, group 3 ILCs (ILC3s) were found to contribute to antitumor immunity by supporting the formation of the tertiary lymphoid structure (TLS) composed of a network of CD21+ follicular dendritic cells surrounded by T cells and CD21+ B cells through the expression of TNF-α, LTα, and LTβ [203]. We found that NKp44+ ILC3s highly express LTA, LTB, and TNF in the tumors of patients with colorectal cancers at the T1 or T2 stage, as well as normal colon, and that these cells induce the expression of lymphoid structure formation-related chemokines (e.g., CXCL13, CCL9, and CCL21) and adhesion molecules (e.g., VCAM-1 and ICAM-1) in CD90+ mesenchymal stromal cells [204]. In T3/T4 colorectal cancer, the number of NKp44+ ILC3s and the expression of LTA, LTB, and TNF, but not ILC3-related molecules (e.g., RORC, IL22, and IL17A), in these cells were decreased. Accordingly, the expression of CXCL13, CCL21, ICAM-1, and VCAM-1 was drastically reduced in CD90+ cells within T3/T4 tumor tissues. We found that TLSs are more abundant in the early stages of colorectal cancer, and their number gradually declines as the tumor progresses, indicating that the decreasing number of NKp44+ ILC3s expressing LTA, LTB, and TNF mRNAs correlates with the density of TLSs during colorectal cancer progression. These findings suggest that NKp44+ ILC3-mediated CD90+ mesenchymal stromal cell reprogramming contributes to prevent tumor progression through the induction of TLS formation.

Mesenchymal stomal cells producing the WNT signaling amplifier Rspo3 are increased during intestinal epithelial regeneration to maintain the stem cell compartment and restore epithelial tissue homeostasis [17, 62, 86, 128, 205, 206]. However, the inadequate amplification of WNT signaling is associated with the incidence and/or progression of colorectal cancer. A study showed that colorectal cancer is divided into WNT ligand-independent and ligand-dependent subgroups [207]. The former has mutations in the APC gene and CTNNB1 gene encoding β-catenin in epithelial cells, which causes constitutive activation of the canonical WNT signaling pathway [208]. Meanwhile, the latter shows the upregulation of Rspo3 expression in fibroblast but lacks epithelial mutations in the APC and CTNNB1 genes.

Regarding findings on mesenchymal stromal cells in the context of cancer, it has been revealed that the interactions of several types of cancer-associated fibroblasts (CAFs) with other cell types are associated with metastasis and cancer progression by facilitating angiogenesis and promoting the accumulation of regulatory T cells and anti-inflammatory tumor-associated macrophages [209]. In patients with colorectal cancer, the augmentation of CAFs secreting the BMP signaling antagonist Gremlin 1 is associated with poor survival, whereas CAF secreting the BMP signaling activator ISLR inhibit colorectal cancer progression [210]. Gremlin 1 expression is upregulated in intestinal epithelial cells of patients with hereditary mixed polyposis syndrome characterized by the development of multiple types of colorectal tumors, which provides Lgr5− cells excluded from the stem cell niche with stem-like properties by disrupting the homeostatic Gremlin 1 gradient at the crypt base-villus axis [211]. Elsewhere, it has been reported that CAFs within APC-mutant colorectal cancer secrete a large amount of WNT2 [29, 212, 213], which provokes the expression of extracellular matrix molecules and proangiogenic factors through autocrine WNT signaling via the receptor FZD8 in CAFs and fibroblasts. This in turn leads to the progression of tumor growth and metastasis [213, 214]. Low WNT signaling activity is associated with the induction of CXCL12-producing inflammatory CAFs (iCAFs) that induce expression of the genes involved in epithelial-mesenchymal transition in tumor cells, leading to cancer dissemination [209, 215]. Meanwhile, it has been shown that CXCL12+ iCAFs are increased during colorectal cancer progression, and the pathway related to drug metabolism is activated in these cells [216], suggesting that CXCL12+ iCAFs are associated with chemotherapy resistance in patients with colorectal cancer. TGF-β-induced suppression of PKCζ activity induces CAFs expressing WNT antagonists SFRP1 and SFRP2 in a SOX2-dependent fashion, which is associated with poor prognosis in colorectal cancer by establishing an immunosuppressive tumor environment [217]. These findings indicate that the WNT signaling-induced generation and expansion of several types of CAFs as well as CAF-mediated modulation of WNT signaling exacerbate the progression of intestinal cancer.

Conclusion

Mesenchymal stromal cells have been primarily conceptualized as structural cells that maintain the architecture of organs and tissues and the positioning of the cells within them [12, 157, 209, 218, 219]. However, accumulating evidence indicates that these cells also function as key regulators of immune responses and epithelial integrity in the intestine [55, 94, 101, 102, 220, 221]. As summarized in this review, intestinal mesenchymal stromal cells are a heterogeneous population with distinct spatial localizations and functions in different developmental stages, homeostatic conditions, and intestinal diseases. In patients with intestinal diseases, besides alterations in the character of tissue-resident mesenchymal stromal cells, the microbiota community and metabolomic profiles in biological samples are also changed [222, 223]. Microbiota-derived bioactive metabolites exert beneficial and deleterious effects on the host immune system, epithelial barrier, and microbial ecosystems [224,225,226,227,228]. However, the impacts of microbiota-supplied metabolites on mesenchymal stromal cell physiology remain poorly understood. Moreover, the cellular origin of tissue-resident and disease-associated mesenchymal stromal cells and the mechanism regulating their differentiation, localization, and activation remain incompletely defined. There is thus a need for a more detailed and complete fate map of intestinal mesenchymal stromal cell populations and the local signaling molecules to manipulate their differentiation and function both in a steady state and under inflammatory conditions. This will aid the identification of novel targets for the development of medications for IBD and colorectal cancer, as well as other intractable diseases. There is also a need to define how interactions between factors associated with the intestinal environment and IBD susceptibility genes in mesenchymal cells are involved in the maintenance of intestinal homeostasis and/or the pathogenesis of IBD. This might help to unravel some of the complex etiology of IBD.

Availability of data materials

Not applicable.

Abbreviations

APC:

APC regulator of WNT signaling pathway

ANGPTL4:

Angiopoietin-like 4

BMP:

Bone morphogenetic proteins

CAF:

Cancer-associated fibroblast

CCL:

C–C motif chemokine ligand

Col:

Collagen

CXCL:

C-X-C motif chemokine ligand

CXCR:

C-X-C motif chemokine receptor

DKK:

Dickkopf

Dll1:

Delta-like 1

EGF:

Epidermal growth factor

Fgfr2:

Fibroblast growth factor receptor 2

Ereg:

Epiregulin

Grem:

Gremlin

hDTR:

Human diphtheria toxin receptor

Hh:

Hedgehog

HHIP:

Hedgehog-interacting protein

IBD:

Inflammatory bowel disease

IGF:

Insulin-like growth factor

IFN-γ:

Interferon gamma

IL:

Interleukin

Lgr5:

Leucine-rich repeat-containing G-protein coupled receptor 5

LRC:

Long-term label-retaining cell

Lrp5:

LDL receptor-related protein 5

LTβR:

Lymphotoxin B receptor

M cell:

Microfold cell

MMP:

Matrix metalloproteinase

MAPK:

Mitogen-activated protein kinase

MAP3K8:

Mitogen-activated protein kinase kinase kinase 8

NEC:

Neonatal necrotizing enterocolitis

Ng2:

Neuron-glial antigen 2

NPNT:

Nephronectin

PCP:

Planar cell polarity

PGE2 :

Prostaglandin E2

Porcn:

Porcupine O-acyltransferase

RANKL:

Receptor activator of nuclear factor-kappa B ligand

Rspo:

R-spondin

scRNA-seq:

Single-cell RNA sequencing

Sema3:

Semaphorin 3

SGLT1:

Sodium-dependent glucose transporter 1

Shh:

Sonic hedgehog

TA cell:

Transit-amplifying cell

TCF/LEF:

DNA-binding T-cell factor/lymphoid enhancer factor

TLR4:

Toll-like receptor 4

References

  1. Santos AJM, Lo YH, Mah AT, Kuo CJ. The intestinal stem cell niche: homeostasis and adaptations. Trends Cell Biol. 2018;28(12):1062–78.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Barker N. Adult intestinal stem cells: critical drivers of epithelial homeostasis and regeneration. Nat Rev Mol Cell Biol. 2014;15(1):19–33.

    Article  CAS  PubMed  Google Scholar 

  3. Miner JH, Li C, Mudd JL, Go G, Sutherland AE. Compositional and structural requirements for laminin and basement membranes during mouse embryo implantation and gastrulation. Development. 2004;131(10):2247–56.

    Article  CAS  PubMed  Google Scholar 

  4. Poschl E, Schlotzer-Schrehardt U, Brachvogel B, Saito K, Ninomiya Y, Mayer U. Collagen IV is essential for basement membrane stability but dispensable for initiation of its assembly during early development. Development. 2004;131(7):1619–28.

    Article  PubMed  Google Scholar 

  5. Hussey GS, Keane TJ, Badylak SF. The extracellular matrix of the gastrointestinal tract: a regenerative medicine platform. Nat Rev Gastroenterol Hepatol. 2017;14(9):540–52.

    Article  PubMed  Google Scholar 

  6. Loe AKH, Rao-Bhatia A, Kim JE, Kim TH. Mesenchymal niches for digestive organ development, homeostasis, and disease. Trends Cell Biol. 2021;31(3):152–65.

    Article  CAS  PubMed  Google Scholar 

  7. Plikus MV, Wang X, Sinha S, Forte E, Thompson SM, Herzog EL, Driskell RR, Rosenthal N, Biernaskie J, Horsley V. Fibroblasts: origins, definitions, and functions in health and disease. Cell. 2021;184(15):3852–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Buechler MB, Pradhan RN, Krishnamurty AT, Cox C, Calviello AK, Wang AW, Yang YA, Tam L, Caothien R, Roose-Girma M, et al. Cross-tissue organization of the fibroblast lineage. Nature. 2021;593(7860):575–9.

    Article  CAS  PubMed  Google Scholar 

  9. Krausgruber T, Fortelny N, Fife-Gernedl V, Senekowitsch M, Schuster LC, Lercher A, Nemc A, Schmidl C, Rendeiro AF, Bergthaler A, et al. Structural cells are key regulators of organ-specific immune responses. Nature. 2020;583(7815):296–302.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Zhu S, Chen M, Ying Y, Wu Q, Huang Z, Ni W, Wang X, Xu H, Bennett S, Xiao J, et al. Versatile subtypes of pericytes and their roles in spinal cord injury repair, bone development and repair. Bone Res. 2022;10(1):30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Avolio E, Campagnolo P, Katare R, Madeddu P. The role of cardiac pericytes in health and disease: therapeutic targets for myocardial infarction. Nat Rev Cardiol. 2024;21(2):106–18.

    Article  PubMed  Google Scholar 

  12. Koliaraki V, Prados A, Armaka M, Kollias G. The mesenchymal context in inflammation, immunity and cancer. Nat Immunol. 2020;21(9):974–82.

    Article  CAS  PubMed  Google Scholar 

  13. Sugiyama T, Kohara H, Noda M, Nagasawa T. Maintenance of the hematopoietic stem cell pool by CXCL12-CXCR4 chemokine signaling in bone marrow stromal cell niches. Immunity. 2006;25(6):977–88.

    Article  CAS  PubMed  Google Scholar 

  14. Nakatani T, Sugiyama T, Omatsu Y, Watanabe H, Kondoh G, Nagasawa T. Ebf3(+) niche-derived CXCL12 is required for the localization and maintenance of hematopoietic stem cells. Nat Commun. 2023;14(1):6402.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. McCarthy N, Manieri E, Storm EE, Saadatpour A, Luoma AM, Kapoor VN, Madha S, Gaynor LT, Cox C, Keerthivasan S, et al. Distinct mesenchymal cell populations generate the essential intestinal BMP signaling gradient. Cell Stem Cell. 2020;26(3):391-402 e395.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Kraiczy J, McCarthy N, Malagola E, Tie G, Madha S, Boffelli D, Wagner DE, Wang TC, Shivdasani RA. Graded BMP signaling within intestinal crypt architecture directs self-organization of the Wnt-secreting stem cell niche. Cell Stem Cell. 2023;30(4):433-449 e438.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Degirmenci B, Valenta T, Dimitrieva S, Hausmann G, Basler K. GLI1-expressing mesenchymal cells form the essential Wnt-secreting niche for colon stem cells. Nature. 2018;558(7710):449–53.

    Article  CAS  PubMed  Google Scholar 

  18. Paerregaard SI, Wulff L, Schussek S, Niss K, Morbe U, Jendholm J, Wendland K, Andrusaite AT, Brulois KF, Nibbs RJB, et al. The small and large intestine contain related mesenchymal subsets that derive from embryonic Gli1(+) precursors. Nat Commun. 2023;14(1):2307.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Collaborators GBDIBD. The global, regional, and national burden of inflammatory bowel disease in 195 countries and territories, 1990–2017: a systematic analysis for the global burden of disease study 2017. Lancet Gastroenterol Hepatol. 2020;5(1):17–30.

    Article  Google Scholar 

  20. Ni J, Wu GD, Albenberg L, Tomov VT. Gut microbiota and IBD: causation or correlation? Nat Rev Gastroenterol Hepatol. 2017;14(10):573–84.

    Article  PubMed  PubMed Central  Google Scholar 

  21. Ananthakrishnan AN. Epidemiology and risk factors for IBD. Nat Rev Gastroenterol Hepatol. 2015;12(4):205–17.

    Article  PubMed  Google Scholar 

  22. Kinchen J, Chen HH, Parikh K, Antanaviciute A, Jagielowicz M, Fawkner-Corbett D, Ashley N, Cubitt L, Mellado-Gomez E, Attar M, et al. Structural remodeling of the human colonic mesenchyme in inflammatory bowel disease. Cell. 2018;175(2):372-386 e317.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Smillie CS, Biton M, Ordovas-Montanes J, Sullivan KM, Burgin G, Graham DB, Herbst RH, Rogel N, Slyper M, Waldman J, et al. Intra- and inter-cellular rewiring of the human colon during ulcerative colitis. Cell. 2019;178(3):714-730 e722.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Martin JC, Chang C, Boschetti G, Ungaro R, Giri M, Grout JA, Gettler K, Chuang LS, Nayar S, Greenstein AJ, et al. Single-cell analysis of Crohn’s disease lesions identifies a pathogenic cellular module associated with resistance to anti-TNF therapy. Cell. 2019;178(6):1493-1508 e1420.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Nayar S, Morrison JK, Giri M, Gettler K, Chuang LS, Walker LA, Ko HM, Kenigsberg E, Kugathasan S, Merad M, et al. A myeloid-stromal niche and gp130 rescue in NOD2-driven Crohn’s disease. Nature. 2021;593(7858):275–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Friedrich M, Pohin M, Jackson MA, Korsunsky I, Bullers SJ, Rue-Albrecht K, Christoforidou Z, Sathananthan D, Thomas T, Ravindran R, et al. IL-1-driven stromal-neutrophil interactions define a subset of patients with inflammatory bowel disease that does not respond to therapies. Nat Med. 2021;27(11):1970–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Wang F, Long J, Li L, Wu ZX, Da TT, Wang XQ, Huang C, Jiang YH, Yao XQ, Ma HQ, et al. Single-cell and spatial transcriptome analysis reveals the cellular heterogeneity of liver metastatic colorectal cancer. Sci Adv. 2023;9(24):eadf5464.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Zhou Y, Bian S, Zhou X, Cui Y, Wang W, Wen L, Guo L, Fu W, Tang F. Single-cell multiomics sequencing reveals prevalent genomic alterations in tumor stromal cells of human colorectal cancer. Cancer Cell. 2020;38(6):818-828 e815.

    Article  CAS  PubMed  Google Scholar 

  29. Becker WR, Nevins SA, Chen DC, Chiu R, Horning AM, Guha TK, Laquindanum R, Mills M, Chaib H, Ladabaum U, et al. Single-cell analyses define a continuum of cell state and composition changes in the malignant transformation of polyps to colorectal cancer. Nat Genet. 2022;54(7):985–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Qi J, Sun H, Zhang Y, Wang Z, Xun Z, Li Z, Ding X, Bao R, Hong L, Jia W, et al. Single-cell and spatial analysis reveal interaction of FAP(+) fibroblasts and SPP1(+) macrophages in colorectal cancer. Nat Commun. 2022;13(1):1742.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Beumer J, Clevers H. Cell fate specification and differentiation in the adult mammalian intestine. Nat Rev Mol Cell Biol. 2021;22(1):39–53.

    Article  CAS  PubMed  Google Scholar 

  32. Clevers H. The intestinal crypt, a prototype stem cell compartment. Cell. 2013;154(2):274–84.

    Article  CAS  PubMed  Google Scholar 

  33. Underhill DM, Iliev ID. The mycobiota: interactions between commensal fungi and the host immune system. Nat Rev Immunol. 2014;14(6):405–16.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Berg RD. The indigenous gastrointestinal microflora. Trends Microbiol. 1996;4(11):430–5.

    Article  CAS  PubMed  Google Scholar 

  35. Iliev ID, Leonardi I. Fungal dysbiosis: immunity and interactions at mucosal barriers. Nat Rev Immunol. 2017;17(10):635–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Norman JM, Handley SA, Baldridge MT, Droit L, Liu CY, Keller BC, Kambal A, Monaco CL, Zhao G, Fleshner P, et al. Disease-specific alterations in the enteric virome in inflammatory bowel disease. Cell. 2015;160(3):447–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Cadwell K. The virome in host health and disease. Immunity. 2015;42(5):805–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Powell N, Walker MM, Talley NJ. The mucosal immune system: master regulator of bidirectional gut-brain communications. Nat Rev Gastroenterol Hepatol. 2017;14(3):143–59.

    Article  CAS  PubMed  Google Scholar 

  39. Mowat AM, Agace WW. Regional specialization within the intestinal immune system. Nat Rev Immunol. 2014;14(10):667–85.

    Article  CAS  PubMed  Google Scholar 

  40. Huus KE, Petersen C, Finlay BB. Diversity and dynamism of IgA-microbiota interactions. Nat Rev Immunol. 2021;21(8):514–25.

    Article  CAS  PubMed  Google Scholar 

  41. Gustafsson JK, Johansson MEV. The role of goblet cells and mucus in intestinal homeostasis. Nat Rev Gastroenterol Hepatol. 2022;19(12):785–803.

    Article  PubMed  Google Scholar 

  42. Gribble FM, Reimann F. Function and mechanisms of enteroendocrine cells and gut hormones in metabolism. Nat Rev Endocrinol. 2019;15(4):226–37.

    Article  CAS  PubMed  Google Scholar 

  43. Wallaeys C, Garcia-Gonzalez N, Libert C. Paneth cells as the cornerstones of intestinal and organismal health: a primer. EMBO Mol Med. 2023;15(2): e16427.

    Article  CAS  PubMed  Google Scholar 

  44. Neurath MF. Targeting immune cell circuits and trafficking in inflammatory bowel disease. Nat Immunol. 2019;20(8):970–9.

    Article  CAS  PubMed  Google Scholar 

  45. Patankar JV, Becker C. Cell death in the gut epithelium and implications for chronic inflammation. Nat Rev Gastroenterol Hepatol. 2020;17(9):543–56.

    Article  PubMed  Google Scholar 

  46. de Souza HS, Fiocchi C. Immunopathogenesis of IBD: current state of the art. Nat Rev Gastroenterol Hepatol. 2016;13(1):13–27.

    Article  PubMed  Google Scholar 

  47. Leblond CP, Walker BE. Renewal of cell populations. Physiol Rev. 1956;36(2):255–76.

    Article  CAS  PubMed  Google Scholar 

  48. Merenda A, Fenderico N, Maurice MM. Wnt signaling in 3D: recent advances in the applications of intestinal organoids. Trends Cell Biol. 2020;30(1):60–73.

    Article  CAS  PubMed  Google Scholar 

  49. van Es JH, Haegebarth A, Kujala P, Itzkovitz S, Koo BK, Boj SF, Korving J, van den Born M, van Oudenaarden A, Robine S, et al. A critical role for the Wnt effector Tcf4 in adult intestinal homeostatic self-renewal. Mol Cell Biol. 2012;32(10):1918–27.

    Article  PubMed  PubMed Central  Google Scholar 

  50. Korinek V, Barker N, Willert K, Molenaar M, Roose J, Wagenaar G, Markman M, Lamers W, Destree O, Clevers H. Two members of the Tcf family implicated in Wnt/beta-catenin signaling during embryogenesis in the mouse. Mol Cell Biol. 1998;18(3):1248–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Niehrs C. The complex world of WNT receptor signalling. Nat Rev Mol Cell Biol. 2012;13(12):767–79.

    Article  CAS  PubMed  Google Scholar 

  52. Batlle E, Henderson JT, Beghtel H, van den Born MM, Sancho E, Huls G, Meeldijk J, Robertson J, van de Wetering M, Pawson T, et al. Beta-catenin and TCF mediate cell positioning in the intestinal epithelium by controlling the expression of EphB/ephrinB. Cell. 2002;111(2):251–63.

    Article  CAS  PubMed  Google Scholar 

  53. Holmberg J, Genander M, Halford MM, Anneren C, Sondell M, Chumley MJ, Silvany RE, Henkemeyer M, Frisen J. EphB receptors coordinate migration and proliferation in the intestinal stem cell niche. Cell. 2006;125(6):1151–63.

    Article  CAS  PubMed  Google Scholar 

  54. Ogasawara R, Hashimoto D, Kimura S, Hayase E, Ara T, Takahashi S, Ohigashi H, Yoshioka K, Tateno T, Yokoyama E, et al. Intestinal lymphatic endothelial cells produce R-Spondin3. Sci Rep. 2018;8(1):10719.

    Article  PubMed  PubMed Central  Google Scholar 

  55. McCarthy N, Kraiczy J, Shivdasani RA. Cellular and molecular architecture of the intestinal stem cell niche. Nat Cell Biol. 2020;22(9):1033–41.

    Article  CAS  PubMed  Google Scholar 

  56. Palikuqi B, Rispal J, Reyes EA, Vaka D, Boffelli D, Klein O. Lymphangiocrine signals are required for proper intestinal repair after cytotoxic injury. Cell Stem Cell. 2022;29(8):1262-1272 e1265.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Niec RE, Chu T, Schernthanner M, Gur-Cohen S, Hidalgo L, Pasolli HA, Luckett KA, Wang Z, Bhalla SR, Cambuli F, et al. Lymphatics act as a signaling hub to regulate intestinal stem cell activity. Cell Stem Cell. 2022;29(7):1067-1082 e1018.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. de Lau W, Barker N, Low TY, Koo BK, Li VS, Teunissen H, Kujala P, Haegebarth A, Peters PJ, van de Wetering M, et al. Lgr5 homologues associate with Wnt receptors and mediate R-spondin signalling. Nature. 2011;476(7360):293–7.

    Article  PubMed  Google Scholar 

  59. Carmon KS, Gong X, Lin Q, Thomas A, Liu Q. R-spondins function as ligands of the orphan receptors LGR4 and LGR5 to regulate Wnt/beta-catenin signaling. Proc Natl Acad Sci U S A. 2011;108(28):11452–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Glinka A, Dolde C, Kirsch N, Huang YL, Kazanskaya O, Ingelfinger D, Boutros M, Cruciat CM, Niehrs C. LGR4 and LGR5 are R-spondin receptors mediating Wnt/beta-catenin and Wnt/PCP signalling. EMBO Rep. 2011;12(10):1055–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Carmon KS, Lin Q, Gong X, Thomas A, Liu Q. LGR5 interacts and cointernalizes with Wnt receptors to modulate Wnt/beta-catenin signaling. Mol Cell Biol. 2012;32(11):2054–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Goto N, Goto S, Imada S, Hosseini S, Deshpande V, Yilmaz OH. Lymphatics and fibroblasts support intestinal stem cells in homeostasis and injury. Cell Stem Cell. 2022;29(8):1246-1261 e1246.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Bottner M, Ghorbani P, Harde J, Barrenschee M, Hellwig I, Vogel I, Ebsen M, Forster E, Wedel T. Expression and regulation of reelin and its receptors in the enteric nervous system. Mol Cell Neurosci. 2014;61:23–33.

    Article  PubMed  Google Scholar 

  64. Garcia-Miranda P, Peral MJ, Ilundain AA. Rat small intestine expresses the reelin-disabled-1 signalling pathway. Exp Physiol. 2010;95(4):498–507.

    Article  CAS  PubMed  Google Scholar 

  65. Carvajal AE, Serrano-Morales JM, Vazquez-Carretero MD, Garcia-Miranda P, Calonge ML, Peral MJ, Ilundain AA. Reelin protects from colon pathology by maintaining the intestinal barrier integrity and repressing tumorigenic genes. Biochim Biophys Acta Mol Basis Dis. 2017;1863(9):2126–34.

    Article  CAS  PubMed  Google Scholar 

  66. Qureshi FG, Leaphart C, Cetin S, Li J, Grishin A, Watkins S, Ford HR, Hackam DJ. Increased expression and function of integrins in enterocytes by endotoxin impairs epithelial restitution. Gastroenterology. 2005;128(4):1012–22.

    Article  CAS  PubMed  Google Scholar 

  67. Islam S, Loizides AM, Fialkovich JJ, Grand RJ, Montgomery RK. Developmental expression of Eph and ephrin family genes in mammalian small intestine. Dig Dis Sci. 2010;55(9):2478–88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Sato T, van Es JH, Snippert HJ, Stange DE, Vries RG, van den Born M, Barker N, Shroyer NF, van de Wetering M, Clevers H. Paneth cells constitute the niche for Lgr5 stem cells in intestinal crypts. Nature. 2011;469(7330):415–8.

    Article  CAS  PubMed  Google Scholar 

  69. Suzuki A, Sekiya S, Gunshima E, Fujii S, Taniguchi H. EGF signaling activates proliferation and blocks apoptosis of mouse and human intestinal stem/progenitor cells in long-term monolayer cell culture. Lab Invest. 2010;90(10):1425–36.

    Article  CAS  PubMed  Google Scholar 

  70. 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(2):177–190 e174.

    Article  CAS  PubMed  Google Scholar 

  71. Wong VW, Stange DE, Page ME, Buczacki S, Wabik A, Itami S, van de Wetering M, Poulsom R, Wright NA, Trotter MW, et al. Lrig1 controls intestinal stem-cell homeostasis by negative regulation of ErbB signalling. Nat Cell Biol. 2012;14(4):401–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. VanDussen KL, Carulli AJ, Keeley TM, Patel SR, Puthoff BJ, Magness ST, Tran IT, Maillard I, Siebel C, Kolterud A, et al. Notch signaling modulates proliferation and differentiation of intestinal crypt base columnar stem cells. Development. 2012;139(3):488–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Pellegrinet L, Rodilla V, Liu Z, Chen S, Koch U, Espinosa L, Kaestner KH, Kopan R, Lewis J, Radtke F. Dll1- and dll4-mediated notch signaling are required for homeostasis of intestinal stem cells. Gastroenterology. 2011;140(4):1230–40 e1231–1237.

    Article  CAS  PubMed  Google Scholar 

  74. Milano J, McKay J, Dagenais C, Foster-Brown L, Pognan F, Gadient R, Jacobs RT, Zacco A, Greenberg B, Ciaccio PJ. Modulation of notch processing by gamma-secretase inhibitors causes intestinal goblet cell metaplasia and induction of genes known to specify gut secretory lineage differentiation. Toxicol Sci. 2004;82(1):341–58.

    Article  CAS  PubMed  Google Scholar 

  75. van Es JH, van Gijn ME, Riccio O, van den Born M, Vooijs M, Begthel H, Cozijnsen M, Robine S, Winton DJ, Radtke F, et al. Notch/gamma-secretase inhibition turns proliferative cells in intestinal crypts and adenomas into goblet cells. Nature. 2005;435(7044):959–63.

    Article  PubMed  Google Scholar 

  76. Brugger MD, Valenta T, Fazilaty H, Hausmann G, Basler K. Distinct populations of crypt-associated fibroblasts act as signaling hubs to control colon homeostasis. PLoS Biol. 2020;18(12): e3001032.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Bahar Halpern K, Massalha H, Zwick RK, Moor AE, Castillo-Azofeifa D, Rozenberg M, Farack L, Egozi A, Miller DR, Averbukh I, et al. Lgr5+ telocytes are a signaling source at the intestinal villus tip. Nat Commun. 2020;11(1):1936.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Melissari MT, Henriques A, Tzaferis C, Prados A, Sarris ME, Chalkidi N, Mavroeidi D, Chouvardas P, Grammenoudi S, Kollias G, et al. Col6a1(+)/CD201(+) mesenchymal cells regulate intestinal morphogenesis and homeostasis. Cell Mol Life Sci. 2021;79(1):1.

    Article  PubMed  PubMed Central  Google Scholar 

  79. Qi Z, Li Y, Zhao B, Xu C, Liu Y, Li H, Zhang B, Wang X, Yang X, Xie W, et al. BMP restricts stemness of intestinal Lgr5(+) stem cells by directly suppressing their signature genes. Nat Commun. 2017;8:13824.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Beumer J, Puschhof J, Yengej FY, Zhao L, Martinez-Silgado A, Blotenburg M, Begthel H, Boot C, van Oudenaarden A, Chen YG, et al. BMP gradient along the intestinal villus axis controls zonated enterocyte and goblet cell states. Cell Rep. 2022;38(9):110438.

    Article  CAS  PubMed  Google Scholar 

  81. Moor AE, Harnik Y, Ben-Moshe S, Massasa EE, Rozenberg M, Eilam R, Bahar Halpern K, Itzkovitz S. Spatial reconstruction of single enterocytes uncovers broad zonation along the intestinal villus axis. Cell. 2018;175(4):1156-1167 e1115.

    Article  CAS  PubMed  Google Scholar 

  82. Beumer J, Artegiani B, Post Y, Reimann F, Gribble F, Nguyen TN, Zeng H, Van den Born M, Van Es JH, Clevers H. Enteroendocrine cells switch hormone expression along the crypt-to-villus BMP signalling gradient. Nat Cell Biol. 2018;20(8):909–16.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Beumer J, Puschhof J, Bauza-Martinez J, Martinez-Silgado A, Elmentaite R, James KR, Ross A, Hendriks D, Artegiani B, Busslinger GA, et al. High-resolution mRNA and secretome atlas of human enteroendocrine cells. Cell. 2020;181(6):1291-1306 e1219.

    Article  CAS  PubMed  Google Scholar 

  84. Berkova L, Fazilaty H, Yang Q, Kubovciak J, Stastna M, Hrckulak D, Vojtechova M, Dalessi T, Brugger MD, Hausmann G, et al. Terminal differentiation of villus tip enterocytes is governed by distinct Tgfbeta superfamily members. EMBO Rep. 2023;24(9): e56454.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Kosinski C, Li VS, Chan AS, Zhang J, Ho C, Tsui WY, Chan TL, Mifflin RC, Powell DW, Yuen ST, et al. Gene expression patterns of human colon tops and basal crypts and BMP antagonists as intestinal stem cell niche factors. Proc Natl Acad Sci U S A. 2007;104(39):15418–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Stzepourginski I, Nigro G, Jacob JM, Dulauroy S, Sansonetti PJ, Eberl G, Peduto L. CD34+ mesenchymal cells are a major component of the intestinal stem cells niche at homeostasis and after injury. Proc Natl Acad Sci U S A. 2017;114(4):E506–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Baghdadi MB, Ayyaz A, Coquenlorge S, Chu B, Kumar S, Streutker C, Wrana JL, Kim TH. Enteric glial cell heterogeneity regulates intestinal stem cell niches. Cell Stem Cell. 2022;29(1):86-100 e106.

    Article  CAS  PubMed  Google Scholar 

  88. Shoshkes-Carmel M, Wang YJ, Wangensteen KJ, Toth B, Kondo A, Massasa EE, Itzkovitz S, Kaestner KH. Subepithelial telocytes are an important source of Wnts that supports intestinal crypts. Nature. 2018;557(7704):242–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Jacob JM, Di Carlo SE, Stzepourginski I, Lepelletier A, Ndiaye PD, Varet H, Legendre R, Kornobis E, Benabid A, Nigro G, et al. PDGFRalpha-induced stromal maturation is required to restrain postnatal intestinal epithelial stemness and promote defense mechanisms. Cell Stem Cell. 2022;29(5):856-868 e855.

    Article  CAS  PubMed  Google Scholar 

  90. Kim JE, Li B, Fei L, Horne R, Lee D, Loe AK, Miyake H, Ayar E, Kim DK, Surette MG, et al. Gut microbiota promotes stem cell differentiation through macrophage and mesenchymal niches in early postnatal development. Immunity. 2022;55(12):2300-2317 e2306.

    Article  CAS  PubMed  Google Scholar 

  91. Durand A, Donahue B, Peignon G, Letourneur F, Cagnard N, Slomianny C, Perret C, Shroyer NF, Romagnolo B. Functional intestinal stem cells after Paneth cell ablation induced by the loss of transcription factor Math1 (Atoh1). Proc Natl Acad Sci U S A. 2012;109(23):8965–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Kim TH, Escudero S, Shivdasani RA. Intact function of Lgr5 receptor-expressing intestinal stem cells in the absence of Paneth cells. Proc Natl Acad Sci U S A. 2012;109(10):3932–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Farin HF, Van Es JH. Clevers H: redundant sources of Wnt regulate intestinal stem cells and promote formation of paneth cells. Gastroenterology. 2012;143(6):1518-1529 e1517.

    Article  CAS  PubMed  Google Scholar 

  94. Kolev HM, Kaestner KH. Mammalian intestinal development and differentiation-the state of the art. Cell Mol Gastroenterol Hepatol. 2023;16(5):809–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Shyer AE, Tallinen T, Nerurkar NL, Wei Z, Gil ES, Kaplan DL, Tabin CJ, Mahadevan L. Villification: how the gut gets its villi. Science. 2013;342(6155):212–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Chin AM, Tsai YH, Finkbeiner SR, Nagy MS, Walker EM, Ethen NJ, Williams BO, Battle MA, Spence JR. A dynamic WNT/beta-catenin signaling environment leads to WNT-independent and WNT-dependent proliferation of embryonic intestinal progenitor cells. Stem Cell Reports. 2016;7(5):826–39.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Shyer AE, Huycke TR, Lee C, Mahadevan L, Tabin CJ. Bending gradients: how the intestinal stem cell gets its home. Cell. 2015;161(3):569–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Arsenault P, Menard D. Cell proliferation in developing human jejunum. Biol Neonate. 1987;51(6):297–304.

    Article  CAS  PubMed  Google Scholar 

  99. Fazilaty H, Brugger MD, Valenta T, Szczerba BM, Berkova L, Doumpas N, Hausmann G, Scharl M, Basler K. Tracing colonic embryonic transcriptional profiles and their reactivation upon intestinal damage. Cell Rep. 2021;36(5): 109484.

    Article  CAS  PubMed  Google Scholar 

  100. Fawkner-Corbett D, Antanaviciute A, Parikh K, Jagielowicz M, Geros AS, Gupta T, Ashley N, Khamis D, Fowler D, Morrissey E, et al. Spatiotemporal analysis of human intestinal development at single-cell resolution. Cell. 2021;184(3):810-826 e823.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Brugger MD, Basler K. The diverse nature of intestinal fibroblasts in development, homeostasis, and disease. Trends Cell Biol. 2023;33(10):834–49.

    Article  PubMed  Google Scholar 

  102. Felsenthal N, Vignjevic DM. Stand by me: fibroblasts regulation of the intestinal epithelium during development and homeostasis. Curr Opin Cell Biol. 2022;78: 102116.

    Article  CAS  PubMed  Google Scholar 

  103. Karlsson L, Lindahl P, Heath JK, Betsholtz C. Abnormal gastrointestinal development in PDGF-A and PDGFR-(alpha) deficient mice implicates a novel mesenchymal structure with putative instructive properties in villus morphogenesis. Development. 2000;127(16):3457–66.

    Article  CAS  PubMed  Google Scholar 

  104. Walton KD, Kolterud A, Czerwinski MJ, Bell MJ, Prakash A, Kushwaha J, Grosse AS, Schnell S, Gumucio DL. Hedgehog-responsive mesenchymal clusters direct patterning and emergence of intestinal villi. Proc Natl Acad Sci U S A. 2012;109(39):15817–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Walton KD, Whidden M, Kolterud A, Shoffner SK, Czerwinski MJ, Kushwaha J, Parmar N, Chandhrasekhar D, Freddo AM, Schnell S, et al. Villification in the mouse: BMP signals control intestinal villus patterning. Development. 2016;143(3):427–36.

    CAS  PubMed  PubMed Central  Google Scholar 

  106. Mao J, Kim BM, Rajurkar M, Shivdasani RA, McMahon AP. Hedgehog signaling controls mesenchymal growth in the developing mammalian digestive tract. Development. 2010;137(10):1721–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Rao-Bhatia A, Zhu M, Yin WC, Coquenlorge S, Zhang X, Woo J, Sun Y, Dean CH, Liu A, Hui CC, et al. Hedgehog-activated Fat4 and PCP pathways mediate mesenchymal cell clustering and villus formation in gut development. Dev Cell. 2020;52(5):647-658 e646.

    Article  CAS  PubMed  Google Scholar 

  108. Daulat AM, Borg JP. Wnt/planar cell polarity signaling: new opportunities for cancer treatment. Trends Cancer. 2017;3(2):113–25.

    Article  CAS  PubMed  Google Scholar 

  109. Madison BB, McKenna LB, Dolson D, Epstein DJ, Kaestner KH. FoxF1 and FoxL1 link hedgehog signaling and the control of epithelial proliferation in the developing stomach and intestine. J Biol Chem. 2009;284(9):5936–44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Kaestner KH, Silberg DG, Traber PG, Schutz G. The mesenchymal winged helix transcription factor Fkh6 is required for the control of gastrointestinal proliferation and differentiation. Genes Dev. 1997;11(12):1583–95.

    Article  CAS  PubMed  Google Scholar 

  111. Calvert R, Pothier P. Migration of fetal intestinal intervillous cells in neonatal mice. Anat Rec. 1990;227(2):199–206.

    Article  CAS  PubMed  Google Scholar 

  112. Holloway EM, Czerwinski M, Tsai YH, Wu JH, Wu A, Childs CJ, Walton KD, Sweet CW, Yu Q, Glass I, et al. Mapping development of the human intestinal niche at single-cell resolution. Cell Stem Cell. 2021;28(3):568-580 e564.

    Article  CAS  PubMed  Google Scholar 

  113. Jarde T, Chan WH, Rossello FJ, Kaur Kahlon T, Theocharous M, Kurian Arackal T, Flores T, Giraud M, Richards E, Chan E, et al. Mesenchymal niche-derived Neuregulin-1 drives intestinal stem cell proliferation and regeneration of damaged epithelium. Cell Stem Cell. 2020;27(4):646-662 e647.

    Article  CAS  PubMed  Google Scholar 

  114. McCarthy N, Tie G, Madha S, He R, Kraiczy J, Maglieri A, Shivdasani RA. Smooth muscle contributes to the development and function of a layered intestinal stem cell niche. Dev Cell. 2023;58(7):550-564 e556.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Neu J, Walker WA. Necrotizing enterocolitis. N Engl J Med. 2011;364(3):255–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Bowker RM, Yan X, De Plaen IG. Intestinal microcirculation and necrotizing enterocolitis: the vascular endothelial growth factor system. Semin Fetal Neonatal Med. 2018;23(6):411–5.

    Article  PubMed  Google Scholar 

  117. Wang Y, Hoenig JD, Malin KJ, Qamar S, Petrof EO, Sun J, Antonopoulos DA, Chang EB, Claud EC. 16S rRNA gene-based analysis of fecal microbiota from preterm infants with and without necrotizing enterocolitis. ISME J. 2009;3(8):944–54.

    Article  CAS  PubMed  Google Scholar 

  118. Weintraub AS, Ferrara L, Deluca L, Moshier E, Green RS, Oakman E, Lee MJ, Rand L. Antenatal antibiotic exposure in preterm infants with necrotizing enterocolitis. J Perinatol. 2012;32(9):705–9.

    Article  CAS  PubMed  Google Scholar 

  119. Brower-Sinning R, Zhong D, Good M, Firek B, Baker R, Sodhi CP, Hackam DJ, Morowitz MJ. Mucosa-associated bacterial diversity in necrotizing enterocolitis. PLoS ONE. 2014;9(9): e105046.

    Article  PubMed  PubMed Central  Google Scholar 

  120. Cotten CM, Taylor S, Stoll B, Goldberg RN, Hansen NI, Sanchez PJ, Ambalavanan N, Benjamin DK Jr, Network NNR. Prolonged duration of initial empirical antibiotic treatment is associated with increased rates of necrotizing enterocolitis and death for extremely low birth weight infants. Pediatrics. 2009;123(1):58–66.

    Article  PubMed  Google Scholar 

  121. Mai V, Young CM, Ukhanova M, Wang X, Sun Y, Casella G, Theriaque D, Li N, Sharma R, Hudak M, et al. Fecal microbiota in premature infants prior to necrotizing enterocolitis. PLoS ONE. 2011;6(6): e20647.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Powell DW, Mifflin RC, Valentich JD, Crowe SE, Saada JI, West AB. Myofibroblasts. I. Paracrine cells important in health and disease. Am J Physiol. 1999;277(1):C1-9.

    Article  CAS  PubMed  Google Scholar 

  123. Kondo A, Kaestner KH. FoxL1(+) mesenchymal cells are a critical source of Wnt5a for midgut elongation during mouse embryonic intestinal development. Cells Dev. 2021;165: 203662.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Aoki R, Shoshkes-Carmel M, Gao N, Shin S, May CL, Golson ML, Zahm AM, Ray M, Wiser CL, Wright CV, et al. Foxl1-expressing mesenchymal cells constitute the intestinal stem cell niche. Cell Mol Gastroenterol Hepatol. 2016;2(2):175–88.

    Article  PubMed  Google Scholar 

  125. Katz JP, Perreault N, Goldstein BG, Chao HH, Ferraris RP, Kaestner KH. Foxl1 null mice have abnormal intestinal epithelia, postnatal growth retardation, and defective intestinal glucose uptake. Am J Physiol Gastrointest Liver Physiol. 2004;287(4):G856-864.

    Article  CAS  PubMed  Google Scholar 

  126. Perreault N, Sackett SD, Katz JP, Furth EE, Kaestner KH. Foxl1 is a mesenchymal Modifier of Min in carcinogenesis of stomach and colon. Genes Dev. 2005;19(3):311–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Takano-Maruyama M, Hase K, Fukamachi H, Kato Y, Koseki H, Ohno H. Foxl1-deficient mice exhibit aberrant epithelial cell positioning resulting from dysregulated EphB/EphrinB expression in the small intestine. Am J Physiol Gastrointest Liver Physiol. 2006;291(1):G163-170.

    Article  CAS  PubMed  Google Scholar 

  128. Greicius G, Kabiri Z, Sigmundsson K, Liang C, Bunte R, Singh MK, Virshup DM. PDGFRalpha(+) pericryptal stromal cells are the critical source of Wnts and RSPO3 for murine intestinal stem cells in vivo. Proc Natl Acad Sci U S A. 2018;115(14):E3173–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Uko V, Radhakrishnan K, Alkhouri N. Short bowel syndrome in children: current and potential therapies. Paediatr Drugs. 2012;14(3):179–88.

    Article  PubMed  Google Scholar 

  130. Zhu G, Lahori D, Schug J, Kaestner KH. Villification of the intestinal epithelium is driven by Foxl1. bioRxiv. 2024. https://doi.org/10.1101/2024.02.27.582300.

  131. Tan C, Norden PR, Yu W, Liu T, Ujiie N, Lee SK, Yan X, Dyakiv Y, Aoto K, Ortega S, et al. Endothelial FOXC1 and FOXC2 promote intestinal regeneration after ischemia-reperfusion injury. EMBO Rep. 2023;24(7): e56030.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Noda M, Omatsu Y, Sugiyama T, Oishi S, Fujii N, Nagasawa T. CXCL12-CXCR4 chemokine signaling is essential for NK-cell development in adult mice. Blood. 2011;117(2):451–8.

    Article  CAS  PubMed  Google Scholar 

  133. Cordeiro Gomes A, Hara T, Lim VY, Herndler-Brandstetter D, Nevius E, Sugiyama T, Tani-Ichi S, Schlenner S, Richie E, Rodewald HR, et al. Hematopoietic stem cell niches produce lineage-instructive signals to control multipotent progenitor differentiation. Immunity. 2016;45(6):1219–31.

    Article  CAS  PubMed  Google Scholar 

  134. Tokoyoda K, Egawa T, Sugiyama T, Choi BI, Nagasawa T. Cellular niches controlling B lymphocyte behavior within bone marrow during development. Immunity. 2004;20(6):707–18.

    Article  CAS  PubMed  Google Scholar 

  135. Kohara H, Omatsu Y, Sugiyama T, Noda M, Fujii N, Nagasawa T. Development of plasmacytoid dendritic cells in bone marrow stromal cell niches requires CXCL12-CXCR4 chemokine signaling. Blood. 2007;110(13):4153–60.

    Article  CAS  PubMed  Google Scholar 

  136. Mandal M, Okoreeh MK, Kennedy DE, Maienschein-Cline M, Ai J, McLean KC, Kaverina N, Veselits M, Aifantis I, Gounari F, et al. CXCR4 signaling directs Igk recombination and the molecular mechanisms of late B lymphopoiesis. Nat Immunol. 2019;20(10):1393–403.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Cali B, Deygas M, Munari F, Marcuzzi E, Cassara A, Toffali L, Vetralla M, Bernard M, Piel M, Gagliano O, et al. Atypical CXCL12 signaling enhances neutrophil migration by modulating nuclear deformability. Sci Signal. 2022;15(761):eabk2552.

    Article  CAS  PubMed  Google Scholar 

  138. Hampton HR, Bailey J, Tomura M, Brink R, Chtanova T. Microbe-dependent lymphatic migration of neutrophils modulates lymphocyte proliferation in lymph nodes. Nat Commun. 2015;6:7139.

    Article  CAS  PubMed  Google Scholar 

  139. Suratt BT, Petty JM, Young SK, Malcolm KC, Lieber JG, Nick JA, Gonzalo JA, Henson PM, Worthen GS. Role of the CXCR4/SDF-1 chemokine axis in circulating neutrophil homeostasis. Blood. 2004;104(2):565–71.

    Article  CAS  PubMed  Google Scholar 

  140. Contento RL, Molon B, Boularan C, Pozzan T, Manes S, Marullo S, Viola A. CXCR4-CCR5: a couple modulating T cell functions. Proc Natl Acad Sci U S A. 2008;105(29):10101–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Kabashima K, Shiraishi N, Sugita K, Mori T, Onoue A, Kobayashi M, Sakabe J, Yoshiki R, Tamamura H, Fujii N, et al. CXCL12-CXCR4 engagement is required for migration of cutaneous dendritic cells. Am J Pathol. 2007;171(4):1249–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Zimmerman NP, Vongsa RA, Faherty SL, Salzman NH, Dwinell MB. Targeted intestinal epithelial deletion of the chemokine receptor CXCR4 reveals important roles for extracellular-regulated kinase-1/2 in restitution. Lab Invest. 2011;91(7):1040–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Karpus ON, Westendorp BF, Vermeulen JLM, Meisner S, Koster J, Muncan V, Wildenberg ME, van den Brink GR. Colonic CD90+Crypt fibroblasts secrete semaphorins to support epithelial growth. Cell Rep. 2019;26(13):3698-+.

    Article  CAS  PubMed  Google Scholar 

  144. Roulis M, Kaklamanos A, Schernthanner M, Bielecki P, Zhao J, Kaffe E, Frommelt LS, Qu R, Knapp MS, Henriques A, et al. Paracrine orchestration of intestinal tumorigenesis by a mesenchymal niche. Nature. 2020;580(7804):524–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Pinchuk IV, Saada JI, Beswick EJ, Boya G, Qiu SM, Mifflin RC, Raju GS, Reyes VE, Powell DW. PD-1 ligand expression by human colonic myofibroblasts/fibroblasts regulates CD4+ T-cell activity. Gastroenterology. 2008;135(4):1228–37.

    Article  CAS  PubMed  Google Scholar 

  146. Beswick EJ, Johnson JR, Saada JI, Humen M, House J, Dann S, Qiu S, Brasier AR, Powell DW, Reyes VE, et al. TLR4 activation enhances the PD-L1-mediated tolerogenic capacity of colonic CD90+ stromal cells. J Immunol. 2014;193(5):2218–29.

    Article  CAS  PubMed  Google Scholar 

  147. Goke M, Zuk A, Podolsky DK. Regulation and function of extracellular matrix intestinal epithelial restitution in vitro. Am J Physiol. 1996;271(5 Pt 1):G729-740.

    CAS  PubMed  Google Scholar 

  148. Voytik-Harbin SL, Brightman AO, Kraine MR, Waisner B, Badylak SF. Identification of extractable growth factors from small intestinal submucosa. J Cell Biochem. 1997;67(4):478–91.

    Article  CAS  PubMed  Google Scholar 

  149. Okamoto R, Watanabe M. Molecular and clinical basis for the regeneration of human gastrointestinal epithelia. J Gastroenterol. 2004;39(1):1–6.

    Article  PubMed  Google Scholar 

  150. Gitter AH, Wullstein F, Fromm M, Schulzke JD. Epithelial barrier defects in ulcerative colitis: characterization and quantification by electrophysiological imaging. Gastroenterology. 2001;121(6):1320–8.

    Article  CAS  PubMed  Google Scholar 

  151. Schmitz H, Barmeyer C, Gitter AH, Wullstein F, Bentzel CJ, Fromm M, Riecken EO, Schulzke JD. Epithelial barrier and transport function of the colon in ulcerative colitis. Ann N Y Acad Sci. 2000;915:312–26.

    Article  CAS  PubMed  Google Scholar 

  152. Lee C, Hong SN, Kim ER, Chang DK, Kim YH. Epithelial regeneration ability of Crohn’s disease assessed using patient-derived intestinal organoids. Int J Mol Sci. 2021;22(11):6013.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Leeb SN, Vogl D, Gunckel M, Kiessling S, Falk W, Goke M, Scholmerich J, Gelbmann CM, Rogler G. Reduced migration of fibroblasts in inflammatory bowel disease: role of inflammatory mediators and focal adhesion kinase. Gastroenterology. 2003;125(5):1341–54.

    Article  CAS  PubMed  Google Scholar 

  154. Meier JK, Scharl M, Miller SN, Brenmoehl J, Hausmann M, Kellermeier S, Scholmerich J, Rogler G. Specific differences in migratory function of myofibroblasts isolated from Crohn’s disease fistulae and strictures. Inflamm Bowel Dis. 2011;17(1):202–12.

    Article  PubMed  Google Scholar 

  155. Lawrance IC, Maxwell L, Doe W. Altered response of intestinal mucosal fibroblasts to profibrogenic cytokines in inflammatory bowel disease. Inflamm Bowel Dis. 2001;7(3):226–36.

    Article  CAS  PubMed  Google Scholar 

  156. Kong L, Pokatayev V, Lefkovith A, Carter GT, Creasey EA, Krishna C, Subramanian S, Kochar B, Ashenberg O, Lau H, et al. The landscape of immune dysregulation in Crohn’s disease revealed through single-cell transcriptomic profiling in the ileum and colon. Immunity. 2023;56(2):444-458 e445.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. D’Alessio S, Ungaro F, Noviello D, Lovisa S, Peyrin-Biroulet L, Danese S. Revisiting fibrosis in inflammatory bowel disease: the gut thickens. Nat Rev Gastroenterol Hepatol. 2022;19(3):169–84.

    Article  CAS  PubMed  Google Scholar 

  158. Lindholm M, Manon-Jensen T, Madsen GI, Krag A, Karsdal MA, Kjeldsen J, Mortensen JH. Extracellular matrix fragments of the basement membrane and the interstitial matrix are serological markers of intestinal tissue remodeling and disease activity in dextran sulfate sodium colitis. Dig Dis Sci. 2019;64(11):3134–42.

    Article  CAS  PubMed  Google Scholar 

  159. de Bruyn M, Arijs I, Wollants WJ, Machiels K, Van Steen K, Van Assche G, Ferrante M, Rutgeerts P, Vermeire S, Opdenakker G. Neutrophil gelatinase B-associated lipocalin and matrix metalloproteinase-9 complex as a surrogate serum marker of mucosal healing in ulcerative colitis. Inflamm Bowel Dis. 2014;20(7):1198–207.

    Article  PubMed  Google Scholar 

  160. de Bruyn M, Arijs I, De Hertogh G, Ferrante M, Van Assche G, Rutgeerts P, Vermeire S, Opdenakker G. Serum neutrophil gelatinase B-associated lipocalin and matrix metalloproteinase-9 complex as a surrogate marker for mucosal healing in patients with Crohn’s disease. J Crohns Colitis. 2015;9(12):1079–87.

    Article  PubMed  Google Scholar 

  161. Bai X, Bai G, Tang L, Liu L, Li Y, Jiang W. Changes in MMP-2, MMP-9, inflammation, blood coagulation and intestinal mucosal permeability in patients with active ulcerative colitis. Exp Ther Med. 2020;20(1):269–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  162. Xiao Y, Lian H, Zhong XS, Krishnachaitanya SS, Cong Y, Dashwood RH, Savidge TC, Powell DW, Liu X, Li Q. Matrix metalloproteinase 7 contributes to intestinal barrier dysfunction by degrading tight junction protein Claudin-7. Front Immunol. 2022;13: 1020902.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Vandenbroucke RE, Dejonckheere E, Van Hauwermeiren F, Lodens S, De Rycke R, Van Wonterghem E, Staes A, Gevaert K, Lopez-Otin C, Libert C. Matrix metalloproteinase 13 modulates intestinal epithelial barrier integrity in inflammatory diseases by activating TNF. EMBO Mol Med. 2013;5(7):1000–16.

    Article  PubMed  Google Scholar 

  164. Vizoso FJ, Gonzalez LO, Corte MD, Corte MG, Bongera M, Martinez A, Martin A, Andicoechea A, Gava RR. Collagenase-3 (MMP-13) expression by inflamed mucosa in inflammatory bowel disease. Scand J Gastroenterol. 2006;41(9):1050–5.

    Article  CAS  PubMed  Google Scholar 

  165. Theiss AL, Simmons JG, Jobin C, Lund PK. Tumor necrosis factor (TNF) alpha increases collagen accumulation and proliferation in intestinal myofibroblasts via TNF receptor 2. J Biol Chem. 2005;280(43):36099–109.

    Article  CAS  PubMed  Google Scholar 

  166. Okuno T, Andoh A, Bamba S, Araki Y, Fujiyama Y, Fujiyama M, Bamba T. Interleukin-1beta and tumor necrosis factor-alpha induce chemokine and matrix metalloproteinase gene expression in human colonic subepithelial myofibroblasts. Scand J Gastroenterol. 2002;37(3):317–24.

    Article  CAS  PubMed  Google Scholar 

  167. Bamba S, Andoh A, Yasui H, Araki Y, Bamba T, Fujiyama Y. Matrix metalloproteinase-3 secretion from human colonic subepithelial myofibroblasts: role of interleukin-17. J Gastroenterol. 2003;38(6):548–54.

    Article  CAS  PubMed  Google Scholar 

  168. Kontoyiannis D, Pasparakis M, Pizarro TT, Cominelli F, Kollias G. Impaired on/off regulation of TNF biosynthesis in mice lacking TNF AU-rich elements: implications for joint and gut-associated immunopathologies. Immunity. 1999;10(3):387–98.

    Article  CAS  PubMed  Google Scholar 

  169. Armaka M, Apostolaki M, Jacques P, Kontoyiannis DL, Elewaut D, Kollias G. Mesenchymal cell targeting by TNF as a common pathogenic principle in chronic inflammatory joint and intestinal diseases. J Exp Med. 2008;205(2):331–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Cintron JR, Abcarian H, Chaudhry V, Singer M, Hunt S, Birnbaum E, Mutch MG, Fleshman J. Treatment of fistula-in-ano using a porcine small intestinal submucosa anal fistula plug. Tech Coloproctol. 2013;17(2):187–91.

    Article  CAS  PubMed  Google Scholar 

  171. O’Connor L, Champagne BJ, Ferguson MA, Orangio GR, Schertzer ME, Armstrong DN. Efficacy of anal fistula plug in closure of Crohn’s anorectal fistulas. Dis Colon Rectum. 2006;49(10):1569–73.

    Article  PubMed  Google Scholar 

  172. Ueno T, Oga A, Takahashi T, Pappas TN. Small intestinal submucosa (SIS) in the repair of a cecal wound in unprepared bowel in rats. J Gastrointest Surg. 2007;11(7):918–22.

    Article  PubMed  Google Scholar 

  173. Keane TJ, Dziki J, Sobieski E, Smoulder A, Castleton A, Turner N, White LJ, Badylak SF. Restoring mucosal barrier function and modifying macrophage phenotype with an extracellular matrix hydrogel: potential therapy for ulcerative colitis. J Crohns Colitis. 2017;11(3):360–8.

    PubMed  Google Scholar 

  174. Takeuchi O, Akira S. Pattern recognition receptors and inflammation. Cell. 2010;140(6):805–20.

    Article  CAS  PubMed  Google Scholar 

  175. Otte JM, Rosenberg IM, Podolsky DK. Intestinal myofibroblasts in innate immune responses of the intestine. Gastroenterology. 2003;124(7):1866–78.

    Article  CAS  PubMed  Google Scholar 

  176. Walton KL, Holt L, Sartor RB. Lipopolysaccharide activates innate immune responses in murine intestinal myofibroblasts through multiple signaling pathways. Am J Physiol Gastrointest Liver Physiol. 2009;296(3):G601-611.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. Roulis M, Nikolaou C, Kotsaki E, Kaffe E, Karagianni N, Koliaraki V, Salpea K, Ragoussis J, Aidinis V, Martini E, et al. Intestinal myofibroblast-specific Tpl2-Cox-2-PGE2 pathway links innate sensing to epithelial homeostasis. Proc Natl Acad Sci U S A. 2014;111(43):E4658-4667.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Brown SL, Riehl TE, Walker MR, Geske MJ, Doherty JM, Stenson WF, Stappenbeck TS. Myd88-dependent positioning of Ptgs2-expressing stromal cells maintains colonic epithelial proliferation during injury. J Clin Invest. 2007;117(1):258–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Gao L, Yu Q, Zhang H, Wang Z, Zhang T, Xiang J, Yu S, Zhang S, Wu H, Xu Y, et al. A resident stromal cell population actively restrains innate immune response in the propagation phase of colitis pathogenesis in mice. Sci Transl Med. 2021;13(603):eabb5071.

    Article  CAS  PubMed  Google Scholar 

  180. Gurtner A, Borrelli C, Gonzalez-Perez I, Bach K, Acar IE, Nunez NG, Crepaz D, Handler K, Vu VP, Lafzi A, et al. Active eosinophils regulate host defence and immune responses in colitis. Nature. 2023;615(7950):151–7.

    Article  CAS  PubMed  Google Scholar 

  181. Garrido-Trigo A, Corraliza AM, Veny M, Dotti I, Melon-Ardanaz E, Rill A, Crowell HL, Corbi A, Gudino V, Esteller M, et al. Macrophage and neutrophil heterogeneity at single-cell spatial resolution in human inflammatory bowel disease. Nat Commun. 2023;14(1):4506.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Neurath MF. Strategies for targeting cytokines in inflammatory bowel disease. Nat Rev Immunol. 2024;24:559–76.

    Article  CAS  PubMed  Google Scholar 

  183. Digby-Bell JL, Atreya R, Monteleone G, Powell N. Interrogating host immunity to predict treatment response in inflammatory bowel disease. Nat Rev Gastroenterol Hepatol. 2020;17(1):9–20.

    Article  PubMed  Google Scholar 

  184. Ming WJ, Bersani L, Mantovani A. Tumor necrosis factor is chemotactic for monocytes and polymorphonuclear leukocytes. J Immunol. 1987;138(5):1469–74.

    Article  CAS  PubMed  Google Scholar 

  185. Wedemeyer J, Lorentz A, Goke M, Meier PN, Flemming P, Dahinden CA, Manns MP, Bischoff SC. Enhanced production of monocyte chemotactic protein 3 in inflammatory bowel disease mucosa. Gut. 1999;44(5):629–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  186. Scheurich P, Thoma B, Ucer U, Pfizenmaier K. Immunoregulatory activity of recombinant human tumor necrosis factor (TNF)-alpha: induction of TNF receptors on human T cells and TNF-alpha-mediated enhancement of T cell responses. J Immunol. 1987;138(6):1786–90.

    Article  CAS  PubMed  Google Scholar 

  187. Ranges GE, Zlotnik A, Espevik T, Dinarello CA, Cerami A, Palladino MA. Jr.: umor necrosis factor alpha/cachectin is a growth factor for thymocytes. Synergistic interactions with other cytokines. J Exp Med. 1988;167(4):1472–8.

    Article  CAS  PubMed  Google Scholar 

  188. Hurme M. Both interleukin 1 and tumor necrosis factor enhance thymocyte proliferation. Eur J Immunol. 1988;18(8):1303–6.

    Article  CAS  PubMed  Google Scholar 

  189. Yokota S, Geppert TD, Lipsky PE. Enhancement of antigen- and mitogen-induced human T lymphocyte proliferation by tumor necrosis factor-alpha. J Immunol. 1988;140(2):531–6.

    Article  CAS  PubMed  Google Scholar 

  190. Danese S, Sans M, de la Motte C, Graziani C, West G, Phillips MH, Pola R, Rutella S, Willis J, Gasbarrini A, et al. Angiogenesis as a novel component of inflammatory bowel disease pathogenesis. Gastroenterology. 2006;130(7):2060–73.

    Article  CAS  PubMed  Google Scholar 

  191. Danese S, Vuitton L, Peyrin-Biroulet L. Biologic agents for IBD: practical insights. Nat Rev Gastroenterol Hepatol. 2015;12(9):537–45.

    Article  CAS  PubMed  Google Scholar 

  192. Van den Brande JM, Koehler TC, Zelinkova Z, Bennink RJ, te Velde AA, ten Cate FJ, van Deventer SJ, Peppelenbosch MP, Hommes DW. Prediction of antitumour necrosis factor clinical efficacy by real-time visualisation of apoptosis in patients with Crohn’s disease. Gut. 2007;56(4):509–17.

    Article  PubMed  Google Scholar 

  193. Atreya R, Zimmer M, Bartsch B, Waldner MJ, Atreya I, Neumann H, Hildner K, Hoffman A, Kiesslich R, Rink AD, et al. Antibodies against tumor necrosis factor (TNF) induce T-cell apoptosis in patients with inflammatory bowel diseases via TNF receptor 2 and intestinal CD14(+) macrophages. Gastroenterology. 2011;141(6):2026–38.

    Article  CAS  PubMed  Google Scholar 

  194. Gomollon F, Dignass A, Annese V, Tilg H, Van Assche G, Lindsay JO, Peyrin-Biroulet L, Cullen GJ, Daperno M, Kucharzik T, et al. 3rd European evidence-based consensus on the diagnosis and management of Crohn’s disease 2016: part 1: diagnosis and medical management. J Crohns Colitis. 2017;11(1):3–25.

    Article  PubMed  Google Scholar 

  195. Atreya R, Neurath MF, Siegmund B. Personalizing treatment in IBD: hype or reality in 2020? Can we predict response to anti-TNF? Front Med (Lausanne). 2020;7:517.

    Article  PubMed  Google Scholar 

  196. Eder P, Michalak M, Katulska K, Lykowska-Szuber L, Krela-Kazmierczak I, Stawczyk-Eder K, Klimczak K, Szymczak A, Linke K. Magnetic resonance enterographic predictors of one-year outcome in ileal and ileocolonic Crohn’s disease treated with anti-tumor necrosis factor antibodies. Sci Rep. 2015;5: 10223.

    Article  PubMed  PubMed Central  Google Scholar 

  197. Sochal M, Krzywdzinska M, Gabryelska A, Talar-Wojnarowska R, Bialasiewicz P, Malecka-Panas E. A simple index to predict the efficiency of adalimumab treatment in Crohn disease with a limited duration of therapy. Pol Arch Intern Med. 2020;130(10):910–2.

    PubMed  Google Scholar 

  198. Sparrow MP, Papamichael K, Ward MG, Riviere P, Laharie D, Paul S, Roblin X. Therapeutic drug monitoring of biologics during induction to prevent primary non-response. J Crohns Colitis. 2020;14(4):542–56.

    Article  PubMed  Google Scholar 

  199. Gisbert JP, Panes J. Loss of response and requirement of infliximab dose intensification in Crohn’s disease: a review. Am J Gastroenterol. 2009;104(3):760–7.

    CAS  PubMed  Google Scholar 

  200. Fine S, Papamichael K, Cheifetz AS. Etiology and management of lack or loss of response to anti-tumor necrosis factor therapy in patients with inflammatory bowel disease. Gastroenterol Hepatol (N Y). 2019;15(12):656–65.

    PubMed  Google Scholar 

  201. Ben-Horin S, Chowers Y. Review article: loss of response to anti-TNF treatments in Crohn’s disease. Aliment Pharmacol Ther. 2011;33(9):987–95.

    Article  CAS  PubMed  Google Scholar 

  202. West NR, Hegazy AN, Owens BMJ, Bullers SJ, Linggi B, Buonocore S, Coccia M, Gortz D, This S, Stockenhuber K, et al. Oncostatin M drives intestinal inflammation and predicts response to tumor necrosis factor-neutralizing therapy in patients with inflammatory bowel disease. Nat Med. 2017;23(5):579–89.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  203. Carrega P, Loiacono F, Di Carlo E, Scaramuccia A, Mora M, Conte R, Benelli R, Spaggiari GM, Cantoni C, Campana S, et al. NCR(+)ILC3 concentrate in human lung cancer and associate with intratumoral lymphoid structures. Nat Commun. 2015;6:8280.

    Article  CAS  PubMed  Google Scholar 

  204. Ikeda A, Ogino T, Kayama H, Okuzaki D, Nishimura J, Fujino S, Miyoshi N, Takahashi H, Uemura M, Matsuda C, et al. Human NKp44(+) Group 3 innate lymphoid cells associate with tumor-associated tertiary lymphoid structures in colorectal cancer. Cancer Immunol Res. 2020;8(6):724–31.

    Article  CAS  PubMed  Google Scholar 

  205. Harnack C, Berger H, Antanaviciute A, Vidal R, Sauer S, Simmons A, Meyer TF, Sigal M. R-spondin 3 promotes stem cell recovery and epithelial regeneration in the colon. Nat Commun. 2019;10(1):4368.

    Article  PubMed  PubMed Central  Google Scholar 

  206. Cox CB, Storm EE, Kapoor VN, Chavarria-Smith J, Lin DL, Wang L, Li Y, Kljavin N, Ota N, Bainbridge TW, et al. IL-1R1-dependent signaling coordinates epithelial regeneration in response to intestinal damage. Sci Immunol. 2021;6(59):eabe8856.

    Article  CAS  PubMed  Google Scholar 

  207. Kleeman SO, Koelzer VH, Jones HJ, Vazquez EG, Davis H, East JE, Arnold R, Koppens MA, Blake A, Domingo E, et al. Exploiting differential Wnt target gene expression to generate a molecular biomarker for colorectal cancer stratification. Gut. 2020;69(6):1092–103.

    Article  CAS  PubMed  Google Scholar 

  208. Groden J, Thliveris A, Samowitz W, Carlson M, Gelbert L, Albertsen H, Joslyn G, Stevens J, Spirio L, Robertson M, et al. Identification and characterization of the familial adenomatous polyposis coli gene. Cell. 1991;66(3):589–600.

    Article  CAS  PubMed  Google Scholar 

  209. Chen Y, McAndrews KM, Kalluri R. Clinical and therapeutic relevance of cancer-associated fibroblasts. Nat Rev Clin Oncol. 2021;18(12):792–804.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  210. Kobayashi H, Gieniec KA, Wright JA, Wang T, Asai N, Mizutani Y, Lida T, Ando R, Suzuki N, Lannagan TRM, et al. The balance of stromal BMP signaling mediated by GREM1 and ISLR drives colorectal carcinogenesis. Gastroenterology. 2021;160(4):1224-1239 e1230.

    Article  CAS  PubMed  Google Scholar 

  211. Davis H, Irshad S, Bansal M, Rafferty H, Boitsova T, Bardella C, Jaeger E, Lewis A, Freeman-Mills L, Giner FC, et al. Aberrant epithelial GREM1 expression initiates colonic tumorigenesis from cells outside the stem cell niche. Nat Med. 2015;21(1):62–70.

    Article  CAS  PubMed  Google Scholar 

  212. Aizawa T, Karasawa H, Funayama R, Shirota M, Suzuki T, Maeda S, Suzuki H, Yamamura A, Naitoh T, Nakayama K, et al. Cancer-associated fibroblasts secrete Wnt2 to promote cancer progression in colorectal cancer. Cancer Med. 2019;8(14):6370–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  213. Kramer N, Schmollerl J, Unger C, Nivarthi H, Rudisch A, Unterleuthner D, Scherzer M, Riedl A, Artaker M, Crncec I, et al. Autocrine WNT2 signaling in fibroblasts promotes colorectal cancer progression. Oncogene. 2017;36(39):5460–72.

    Article  CAS  PubMed  Google Scholar 

  214. Unterleuthner D, Neuhold P, Schwarz K, Janker L, Neuditschko B, Nivarthi H, Crncec I, Kramer N, Unger C, Hengstschlager M, et al. Cancer-associated fibroblast-derived WNT2 increases tumor angiogenesis in colon cancer. Angiogenesis. 2020;23(2):159–77.

    Article  CAS  PubMed  Google Scholar 

  215. Mosa MH, Michels BE, Menche C, Nicolas AM, Darvishi T, Greten FR, Farin HF. A Wnt-induced phenotypic switch in cancer-associated fibroblasts inhibits EMT in colorectal cancer. Cancer Res. 2020;80(24):5569–82.

    Article  CAS  PubMed  Google Scholar 

  216. Ke H, Li Z, Li P, Ye S, Huang J, Hu T, Zhang C, Yuan M, Chen Y, Wu X, et al. Dynamic heterogeneity of colorectal cancer during progression revealed clinical risk-associated cell types and regulations in single-cell resolution and spatial context. Gastroenterol Rep (Oxf). 2023;11:goad034.

    Article  PubMed  Google Scholar 

  217. Kasashima H, Duran A, Martinez-Ordonez A, Nakanishi Y, Kinoshita H, Linares JF, Reina-Campos M, Kudo Y, L'Hermitte A, Yashiro M et al.: Stromal SOX2 upregulation promotes tumorigenesis through the generation of a SFRP1/2-expressing cancer-associated fibroblast population. Dev Cell 2021, 56(1):95–110 e110.

    Article  CAS  PubMed  Google Scholar 

  218. Roulis M, Flavell RA. Fibroblasts and myofibroblasts of the intestinal lamina propria in physiology and disease. Differentiation. 2016;92(3):116–31.

    Article  CAS  PubMed  Google Scholar 

  219. De Jaeghere EA, Denys HG, De Wever O. Fibroblasts fuel immune escape in the tumor microenvironment. Trends Cancer. 2019;5(11):704–23.

    Article  PubMed  Google Scholar 

  220. Davidson S, Coles M, Thomas T, Kollias G, Ludewig B, Turley S, Brenner M, Buckley CD. Fibroblasts as immune regulators in infection, inflammation and cancer. Nat Rev Immunol. 2021;21(11):704–17.

    Article  CAS  PubMed  Google Scholar 

  221. Powell DW, Pinchuk IV, Saada JI, Chen X, Mifflin RC. Mesenchymal cells of the intestinal lamina propria. Annu Rev Physiol. 2011;73:213–37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  222. Hou K, Wu ZX, Chen XY, Wang JQ, Zhang D, Xiao C, Zhu D, Koya JB, Wei L, Li J, et al. Microbiota in health and diseases. Signal Transduct Target Ther. 2022;7(1):135.

    Article  PubMed  PubMed Central  Google Scholar 

  223. Fan Y, Pedersen O. Gut microbiota in human metabolic health and disease. Nat Rev Microbiol. 2021;19(1):55–71.

    Article  CAS  PubMed  Google Scholar 

  224. Kayama H, Okumura R, Takeda K. Interaction between the microbiota, epithelia, and immune cells in the intestine. Annu Rev Immunol. 2020;38:23–48.

    Article  CAS  PubMed  Google Scholar 

  225. Kayama H, Takeda K. Manipulation of epithelial integrity and mucosal immunity by host and microbiota-derived metabolites. Eur J Immunol. 2020;50(7):921–31.

    Article  CAS  PubMed  Google Scholar 

  226. Lavelle A, Sokol H. Gut microbiota-derived metabolites as key actors in inflammatory bowel disease. Nat Rev Gastroenterol Hepatol. 2020;17(4):223–37.

    Article  PubMed  Google Scholar 

  227. Rooks MG, Garrett WS. Gut microbiota, metabolites and host immunity. Nat Rev Immunol. 2016;16(6):341–52.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  228. Krautkramer KA, Fan J, Backhed F. Gut microbial metabolites as multi-kingdom intermediates. Nat Rev Microbiol. 2021;19(2):77–94.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We would like to thank Edanz Group (https://jp.edanz.com/ac) for editing a draft of this manuscript and C. Hidaka for secretarial assistance.

Funding

This work was funded by CREST, the Japan Agency for Medical Research and Development (21gm1010004), a Grant-in-Aid for Scientific Research (S) (JP21H050430), a Grant-in-Aid for Scientific Research Fund for the Promotion of Joint International Research (International Leading Research) (22K21354), and a Japan Society for the Promotion of Science Core-to-Core Program (JPJSCCA20210008) (to K. Takeda) and PRIME, the Japan Agency for Medical Research and Development (22gm6210016), JST FOREST Program (JPMJFR215W), and a Grant-in-Aid for Scientific Research (B) (21H02747) (to H. Kayama).

Author information

Authors and Affiliations

Authors

Contributions

HK conceptualized this review, wrote the manuscript, and designed the figures, and KT reviewed the manuscript and helped to supervise the study.

Corresponding author

Correspondence to Hisako Kayama.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kayama, H., Takeda, K. Regulation of intestinal epithelial homeostasis by mesenchymal cells. Inflamm Regener 44, 42 (2024). https://doi.org/10.1186/s41232-024-00355-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s41232-024-00355-0

Keywords