Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 3170 2022-04-22 04:28:19 |
2 format correct + 6 word(s) 3176 2022-04-22 04:34:42 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Corr, S.; Stiegeler, S.; Mercurio, K.; Iancu, M. The Impact of MicroRNAs during Inflammatory Bowel Disease. Encyclopedia. Available online: https://encyclopedia.pub/entry/22149 (accessed on 21 May 2024).
Corr S, Stiegeler S, Mercurio K, Iancu M. The Impact of MicroRNAs during Inflammatory Bowel Disease. Encyclopedia. Available at: https://encyclopedia.pub/entry/22149. Accessed May 21, 2024.
Corr, Sinéad, Sarah Stiegeler, Kevin Mercurio, Miruna Iancu. "The Impact of MicroRNAs during Inflammatory Bowel Disease" Encyclopedia, https://encyclopedia.pub/entry/22149 (accessed May 21, 2024).
Corr, S., Stiegeler, S., Mercurio, K., & Iancu, M. (2022, April 22). The Impact of MicroRNAs during Inflammatory Bowel Disease. In Encyclopedia. https://encyclopedia.pub/entry/22149
Corr, Sinéad, et al. "The Impact of MicroRNAs during Inflammatory Bowel Disease." Encyclopedia. Web. 22 April, 2022.
The Impact of MicroRNAs during Inflammatory Bowel Disease
Edit

Inflammatory bowel disease (IBD) is a debilitating autoimmune disease characterised by chronic inflammation along the GI tract. Patients diagnosed with IBD are symptomatic for recurrent intestinal inflammation, diarrhoea, abdominal pain, rectal bleeding, weight loss and anaemia. Due to its complexity, a number of factors are attributed to IBD aetiology, including patients’ genetics and makeup of microbiota, food and pharmaceutical consumption, and even limiting antigen exposure due to excessive sanitation. All these aspects further contribute to changes in miRNA expression.

microRNAs inflammatory bowel disease

1. Inflammatory Bowel Disease

Inflammatory bowel disease (IBD) is a debilitating autoimmune disease characterised by chronic inflammation along the GI tract. Patients diagnosed with IBD are symptomatic for recurrent intestinal inflammation, diarrhoea, abdominal pain, rectal bleeding, weight loss and anaemia. Due to its complexity, a number of factors are attributed to IBD aetiology, including patients’ genetics and makeup of microbiota, food and pharmaceutical consumption, and even limiting antigen exposure due to excessive sanitation [1][2]. All these aspects further contribute to changes in miRNA expression.
IBD is caused by the overactivation of the mucosal immune system driven mainly by increased exposure to the gut microbiota as a result of compromised gut permeability. Importantly, host genetics are a major factor attributed to the manifestation of disease and linked to multiple regions of the genome [2]. For example, it is more likely for IBD patients to possess variants of the NOD2/CARD15 gene on chromosome 16 than healthy individuals, with this gene encoding a specific pattern-recognition receptor (PRR) for bacterial lipopolysaccharide that regulates macrophage activation of nuclear factor-κB (NF-κB) [3]. Additionally, researchers consistently observe that the IBD3 gene on chromosome 6, which encodes the major histocompatibility complex (MHC), has a genetic linkage to IBD [3]. Other environmental factors that increase IBD incidence include the use of non-steroidal anti-inflammatory drugs, antibiotics and smoking, which are all known to alter the gut microbiota [4].
Clinically, IBD is segregated into two main types known as Crohn’s disease (CD) and ulcerative colitis (UC). CD can impact any part of the GI tract with patchy regions of inflammation, whereas UC has inflammation typically localised to the colon or rectum [5]. Incidence of CD or UC can also lead to an increased risk of other diseases such as colorectal cancer [6]. The activation of central immune cell populations leads to the recruitment of non-specific mediators of the inflammatory response, such as the formation of metabolites such as prostaglandins and leukotrienes, along with damaging compounds such as reactive oxygen species (ROS) [2]. All these factors can compromise the gut epithelial barrier where the majority of host and gut microbiota interactions occur. In such a disease state, luminal antigens gain access to the lamina propria, triggering a response by innate and adaptive immune cells via various PRRs, causing professional (i.e., dendritic cells) and non-professional (i.e., intestinal epithelial cells, IECs) antigen presenting cells (APCs) to further activate central effector immune cells as well as other pro-inflammatory mediators. This cascade thereby perpetuates a positive feedback loop of leukocyte recruitment and increasing tissue damage in both types of IBD [1].
Research on IBD has highlighted changes in gut immunity and the overall functioning of cells that participate in the characteristic overactive immune response. Immune cells can be a part of innate immunity, which serves as the initial rapid defence upon the recognition of foreign pathogens, or adaptive immunity, which leads to slow but long-lasting defensive measures [7]. An effective inflammatory response to any pathogenic invasion is conducted mainly by innate immune cells, which include neutrophils, dendritic cells, monocytes, macrophages and natural killer cells. For an immunological response to specific pathogens, adaptive immune cells are more critical and include effector T-cells, regulatory T-cells and B-cells [7]. The function within, and communication between, the two arms of immunity contribute to chronic inflammation associated with IBD, leading to tissue damage through ROS production, fibrosis and continuous feedback loops of pro-inflammatory cytokine signalling [1].
There are notable immunological differences between CD and UC. Previously accepted notions were that CD typically has a major CD4+ lymphocyte population with a type-1 helper T-cell (Th1) phenotype, while UC has a type-2 helper T-cell (Th2) phenotype [8]. Due to these differences in lymphocyte populations, CD is driven by interferon-γ (IFN-γ) and interleukin-12 (IL-12) expression, while UC is driven by transforming growth factor (TGF)-β, IL-4, IL-5 and IL-13 expression. More recently, this paradigm has been expanded to incorporate the IL-23/Th17 axis to further distinguish the two types [8]. The interactions between immune cells and cells that constitute the gut epithelial barrier are paramount in understanding permeability within CD and UC disease states.

2. Biomarkers and Treatments

The archetypical method of diagnosis for IBD is endoscopy in conjunction with biopsies [9]. Despite the established belief that CD is mainly attributed to the overactivation of the inflammatory response throughout the GI tract while UC is mainly confined to the rectum and colon, there remain issues in the proper diagnosis of the two conditions. Serological biomarkers include the presence of perinuclear staining anti-neutrophil cytoplasmic antibodies in 70% of UC patients, or the presence of anti-Saccharomyces cerevisiae antibodies in 50% of CD patients [10]. Other biomarkers include C-reactive protein in serum and granulocyte proteins lactoferrin and calprotectin in faeces [11]. Recently, a panel of 51 protein biomarkers in serum was determined to be predictive of CD within up to five years before diagnosis that included associated changes to complement cascade, lysosomes, innate immune response and glycosaminoglycan metabolism [12]. Further development on IBD biomarkers is still required for differentiation between CD and UC, along with indication of their severity and active state.
Due to their high stability in bodily fluids, miRNAs have been studied as potential biomarkers of disease [13]. Research has expanded novel diagnostic tools to determine differences in the presence of miRNAs within various samples, with miRNAs being observed in serum, urine, tears, saliva and breast milk [14]. Total serum RNA from active and inactive CD and UC patients demonstrated that miR-595 and miR-1246 were significantly upregulated in IBD and could serve as non-specific biomarkers [15]. In plasma samples, only significant downregulation of miR-16 was validated for diagnosing CD [16]. Whole blood samples obtained from IBD patients showed that CD4+ T-cell expression of miR-1307-3p, miR3615 and miR-4792 predicted disease progression of IBD [17]. In testing miRNA profiles of saliva samples taken from IBD patients, there was a significantly altered expression of miR-101 in CD patients and miR-21, miR-31, miR-142-3p and miR-142-5p in UC [18]. Interestingly, freshly frozen colonic mucosa tissues from IBD patients displayed high levels of miR-31, miR-146a, miR-206 and miR-424, with miR-31 also highly expressed in formalin-fixed, paraffin-embedded tissues [19]. A recent study investigated the mucosal and serum expression of miRNAs in the colon of a canine IBD model. In canine IBD, miR-16, miR-21, miR-122 and miR-147 were elevated in serum and the colonic mucosa, while miR-146a, miR-192 and miR-223 were upregulated in the serum only compared to their controls [20]. In human blood or biopsy samples of IBD patients, among others, an elevation of miR-16, miR-21, miR-106a, miR-122, miR-151-5p, miR-155, miR-199a-5p, miR-320 and miR-362-3p was observed [21][22][23][24]. The variety of samples in which miRNAs exist provide several options for studying biomarkers in IBD.
There remains an emphasis on biomarker studies that show distinguishable aspects between CD and UC for novel diagnostic tools. Some research has looked at developing miRNA panels for CD and UC diagnosis with great accuracy. One study identified an 11-miRNA panel for CD using serum samples [25], while another used platelet-derived miRNAs to determine a 31-miRNA panel for UC [26]. A further study used a six-miRNA panel to distinguish between CD and UC from colon biopsies [18]. Using peripheral blood, an eight-miRNA panel was found to distinguish between active UC and CD [27]. In addition, differentiation between types of IBD and intestinal colitis is also crucial in furthering diagnostic methods. Differential expression of miR-24 allowed researchers to distinguish between UC and L2 CD within rectal biopsies [28]. These advances are necessary for identifying specific treatments tailored uniquely to the patient.
Determining IBD activity is also crucial in prescribing treatments and predicting patient health impacts. One study found that miR-150, miR-196b, miR-199a-3p, miR-199b-5p, miR-223 and miR-320a displayed significant differential expression in non-inflamed UC compared to non-inflamed CD colonic tissues [29]. Another study demonstrated that miR-20b, miR-26b, miR-98, miR-99a and miR-203 were significantly upregulated in colonic mucosal pinch biopsies obtained from patients with active UC compared to quiescent UC [30]. Significant downregulation of miR-192, miR-375 and miR-422b and a significant upregulation in miR-16, miR-21, miR-23a, miR-24, miR-29a, miR-126, miR-195 and let-7f have been observed in sigmoid colon pinch biopsies [31]. Levels of miR-192 were significantly upregulated and miR-16 significantly downregulated in active UC [31][32]. A downregulation in miR-4284 in colonic tissue samples from active UC patients was also observed in a separate study [33]. Moreover, levels of miR-142-5p, miR-595 and miR-1246 in serum samples could differentiate active and non-active CD with high accuracy [15]. Interestingly, miR-31-5p and miR-203 were identified as inflammation-independent diagnostic markers for CD in colonic tissue samples, while miR-215 predicted a specific penetrating/fistulising CD phenotype in the ileum [34]. An overview of altered expression patterns of miRNAs in the context of IBD can be found in Table 1.
Table 1. Summary of altered microRNA expression patterns in IBD.
MicroRNA Expression Level Sample Biomarker References
let-7f upregulated colonic tissue diagnosed UC patients [31]
miR-16 downregulated colonic tissue; plasma active UC; diagnosis of CD [31][32]
upregulated serum and colonic mucosa; blood; biopsy; colonic tissue canine IBD model; diagnosed IBD patients; diagnosed UC patients [21][20][23][31]
miR-20b differential pattern colonic mucosa active vs. quiescence UC [30]
miR-21 upregulated colonic tissue; blood; serum; saliva diagnosed UC patients; diagnosed IBD patients; canine IBD model [21][18][20][23][31]
miR-23a upregulated colonic tissue diagnosed UC patients [31]
miR-24 upregulated colonic tissue diagnosed UC patients [31]
miR-26b differential pattern colonic mucosa active vs. quiescence UC [30]
miR-29a upregulated colonic tissue diagnosed UC patients [31]
miR-31 upregulated colonic mucosa, saliva diagnosed IBD and UC patients [18][19]
miR-31-5p differential pattern colonic tissue diagnostic marker for CD [34]
miR-98 differential pattern colonic mucosa active vs. quiescence UC [30]
miR-99a differential pattern colonic mucosa active vs. quiescence UC [30]
miR-101 upregulated saliva CD [18]
miR-106a upregulated blood/biopsy diagnosed IBD patients [21][23][24]
miR-122 upregulated blood/biopsy; serum and colonic mucosa diagnosed IBD patients; canine IBD model [21][20][23]
miR-126 upregulated colonic tissue diagnosed UC patients [31]
miR-142-3p upregulated saliva UC [18]
miR-142-5p differential pattern serum active vs. quiescence CD [15]
downregulated saliva UC [18]
miR-146a upregulated colonic mucosa; serum diagnosed IBD patients; canine IBD model [19][20]
miR-147 upregulated serum and colonic mucosa canine IBD model [20]
miR-150 differential pattern colonic tissue non-inflamed UC vs. non-inflamed CD [29]
miR-151-5p upregulated blood/biopsy diagnosed IBD patients [21][22][23]
miR-155 upregulated blood/biopsy diagnosed IBD patients [21][22][23]
miR-192 upregulated serum; colonic tissue canine IBD model; active UC [20][31][32]
miR-192 downregulated colonic tissue diagnosed UC patients [31]
miR-195 upregulated colonic tissue diagnosed UC patients [31]
miR-196b differential pattern colonic tissue non-inflamed UC vs. non-inflamed CD [29]
miR-199a-3p differential pattern colonic tissue non-inflamed UC vs. non-inflamed CD [29]
miR-199a-5p upregulated blood/biopsy diagnosed IBD patients [21][23]
miR-199b-5p differential pattern colonic tissue non-inflamed UC vs. non-inflamed CD [29]
miR-203 differential pattern colonic tissue active vs. quiescence UC; diagnostic marker for CD [30][34]
miR-206 upregulated colonic mucosa diagnosed IBD patients [19]
miR-223 upregulated serum canine IBD model [20]
differential pattern colonic tissue non-inflamed UC vs. non-inflamed CD [29]
miR-320 upregulated blood/biopsy diagnosed IBD patients [21][22][23]
miR-320a differential pattern colonic tissue non-inflamed UC vs. non-inflamed CD [29]
miR-362-3p upregulated blood/biopsy diagnosed IBD patients [21][23][24]
miR-375 downregulated colonic tissue diagnosed UC patients [31]
miR-422b downregulated colonic tissue diagnosed UC patients [31]
miR-424 upregulated colonic mucosa diagnosed IBD patients [19]
miR-595 differential pattern serum active vs. quiescence CD [15]
upregulated serum non-specific biomarker for IBD [15]
miR-1246 upregulated serum non-specific biomarker for IBD [15]
differential pattern serum active vs. quiescence CD and UC [15]
miR-1307-3p upregulated blood (CD4+ T-cells) disease progression in IBD [17]
miR-3615 upregulated blood (CD4+ T-cells) disease progression in IBD [17]
miR-4284 downregulated colonic tissue active UC [33]
miR-4792 expression blood (CD4+ T-cells) disease progression in IBD [17]
Treatments for IBD include pharmaceuticals, antibody therapies and full surgical procedures. One frequently prescribed treatment is 5-aminosalicylate, which was suggested to block the production of prostaglandins and leukotrienes, inhibit bacterial peptide-induced neutrophil chemotaxis, scavenge circulating ROS and further inhibit the activation of NF-κB [2]. Others include the use of corticosteroids, though it is important to note that continuous monitoring of prevalent side effects such as corticosteroid-induced osteoporosis, hypertension or diabetes must be undertaken [2]. Immunosuppressants such as azathioprine are also prescribed; however, this often puts patients at risk of opportunistic infections and toxic side effects such as neutropenia, pancreatitis and drug-induced hepatotoxicity [2]. Distinguishing between types of IBD is crucial as immunoregulatory treatments could sometimes be beneficial for one and not the other. This is indeed the case for cyclosporin in treating UC, or anti-tumour necrosis factor (TNF) therapy via infliximab combined with thiopurines for CD [2][9]. Even appendectomies have been associated with an increased risk of developing strictures in CD patients while having improved effects in UC patients [1].
MiRNAs have been investigated as new therapeutic targets and indicators of drug treatment suitability. For example, utilising miR-200c-3p mimics may reduce levels of inflammation caused by IL-8 or NF-κB response to TLR4 activation in IBD [35]. Overexpression of miR-122 was suggested to downregulate NOD2 in IECs, inhibiting apoptosis and destruction of the intestinal barrier [36]. Inhibiting miR-155 expression may be a therapeutic avenue due to its targeting of SOCS1 in UC or the inhibition of the hypoxia-inducible factor (HIF)-1α/trefoil factor (TFF)-3 axis observed in animal models [37][38]. Moreover, it was suggested that using miR-195 as a biomarker may help track therapeutic steroid resistance in IBD patients [39]. Other miRNAs such as let-7d, let-7e, miR-28-5p, miR-221 and miR-224 were found to be significantly increased in patients after six weeks of infliximab treatment, with let-7d and let-7e also showing upregulation in patients undergoing remission [40]. Further details on other miRNAs studied for their regulation in response to specific therapies and as treatment biomarkers are described elsewhere [41].

3. Gut Immunity

Research on IBD and gut immunity uses multiple in vivo models. For example, the dextran sulfate sodium (DSS) mouse model pertains more closely to UC with colitis manifesting due to destruction of the intestinal barrier, while the T-cell adoptive transfer mouse model more closely resembles CD with a focus on early immunological factors [13]. Interestingly, drastic impacts occur upon the deletion of Dicer1 within the intestinal epithelium of mice, one of the major enzymes in miRNA biogenesis, including the spontaneous development of colitis and a dramatic increase in epithelial cell apoptosis [42], highlighting the relevance of miRNA-mediated regulation of the GI barrier. Other chemically induced or knockout (KO) mouse models, along with alternative model organisms used in the study of IBD pathogenicity, are well-described elsewhere [43][44].
A difficulty in studying miRNA relevance in IBD and innate immunity is the species- and cell-specific features. For instance, Toll-like receptors (TLRs) are key for IECs to identify pathogen-associated molecular pathways (PAMPs). TLR4-activated NF-κB induction of microRNA-9 (miR-9) occurs primarily in human monocytes and neutrophils, while miR-210 plays a regulatory role in the NF-κB feedback pathway typically in murine macrophages [45][46]. Additionally, nucleotide-binding and oligomerization domain (NOD)-like receptors (NLRs) are important in the surveillance of the intracellular environment for signs of possible infection. Modulations by miR-29 in general dendritic cells, miR-122 in HT-29 cells, miR-146a in muramyl dipeptide (MDP)-activated macrophages and miR-192/miR-495/miR-512/miR-671 in HCT116 cells were shown to play critical roles in the regulation of NLRs [36][47][48][49]. It is essential that researchers understand the limitations to their study models and explore all alternatives to study miRNA relevance in IBD.
One of the most widely studied miRNAs regarding health and disease is miR-21. Among the most abundantly expressed miRNAs in various mammalian cell types, miR-21 is considered an oncomiR within the intronic region of the protein coding gene TMEM49 [50]. The regulation of miR-21 is still not fully understood, as there are multiple layers towards maturity that can be regulated, including several transcription factors that bind to its promoter region or bind to the pri-miR-21 form [51]. Elevated levels of miR-21 are suggested to be pathological in IBD [52][53]. Epigenome-wide whole blood DNA methylation profiles of paediatric CD treatment-naïve patients showed that hypomethylation of the miR-21 locus correlated with increased expression in leukocytes and inflamed intestinal tissue [54]. Importantly, several studies showed that the ablation of miR-21 in mice led to protection against DSS-induced colitis [55][56]. For UC patients in remission, miR-21 was found to be downregulated while known target programmed cell death protein (PDCD)-4 was upregulated in CD3+ T-cells compared to active disease and healthy controls [57]. Further work is required to understand the full scope of miR-21 influence within the inflamed gut.
Other miRNA KO models demonstrated amelioration during DSS-induced colitis. Like miR-21, the deletion of miR-155 in mice protected against DSS-induced colitis [58]. Additionally, the deletion of miR-301a also protects mice against DSS-induced colitis by rescuing BTG anti-proliferation factor 1 (BTG1) expression and is associated with lowering levels of pro-inflammatory markers such as IL-1β, IL-6, IL-8 and tumour necrosis factor (TNF) [59]. Genetic studies on the consequences of combined KO models could elaborate more on the negative roles these miRNAs have in IBD.
Another hallmark factor in those afflicted by IBD is the presence of oxygen, either as increased levels of ROS through continuous activation of macrophages or the sensing of oxygen in the gut environment. Several miRNAs were shown to be involved in regulating nitric oxide synthase-2 (NOS2) in IBD tissues. Induction of the nitric oxide pathway by miR-21, miR-126, miR-146a, miR-221 and miR-223 led to senescence among adjacent epithelial cells via the upregulation of HP1γ [60]. Regarding the sensing of environmental oxygen in the gut, HIF was demonstrated to be a key regulator of barrier integrity and induced expression of miR-320a to improve barrier function in T84 cells [61]. Developing methods for measuring levels of oxygen as damaging free radicals and in its gaseous state within the inflamed gut may help researchers track the progression of the disease.
Numerous miRNAs demonstrated relevance to IBD and adaptive immunity. T-cells have significant roles in the genesis and development of IBD. The deletion of miR-21 exacerbates CD4+ T-cell-mediated models of colitis, while loss of miR-155 tends to decrease Th1/Th17, showing that these are key regulators in regulatory T-cell (Treg) homeostasis [58][62]. Continued work has demonstrated other miRNAs in Treg regulation such as miR-10a, miR-17-92 cluster, miR-146a and miR-212/132 [63][64][65][66]. MiRNAs implicated in Th1 and Th2 differentiation include miR-17-92 cluster, miR-27b, miR-29, miR-128, miR-146a, miR-155 and miR-340 [65][67][68][69]. For Th17 differentiation, miRNAs that show impact when imbalanced are miR-10a, miR-155 and miR-326, as well as miR-301a as an indirect inducer [70][71][72][73]. Overexpression of miR-210 may negatively impact Th17 differentiation due to targeting hypoxia-induced inhibitor HIF1α [74]. Finally, B-cell maturation was shown to be regulated by miR-10a, miR-17-92 cluster and miR-181a [70][75][76]. The coordinated interplay between immunity regulation and IECs is essential in controlling barrier permeability. Further information on relevant research models used in the study of IBD and their conclusions regarding miRNAs’ impact on pathogenesis has been summarised elsewhere [77].

References

  1. Baumgart, D.C.; Carding, S.R. Inflammatory bowel disease: Cause and immunobiology. Lancet 2007, 369, 1627–1640.
  2. Podolsky, D.K. Inflammatory bowel disease. N. Engl. J. Med. 2002, 347, 417–429.
  3. van Heel, D.A.; Fisher, S.A.; Kirby, A.; Daly, M.J.; Rioux, J.D.; Lewis, C.M.; Genome Scan Meta-Analysis Group of the, I.B.D.I.G.C. Inflammatory bowel disease susceptibility loci defined by genome scan meta-analysis of 1952 affected relative pairs. Hum. Mol. Genet. 2004, 13, 763–770.
  4. Ramos, G.P.; Papadakis, K.A. Mechanisms of disease: Inflammatory bowel diseases. Mayo Clin. Proc. 2019, 94, 155–165.
  5. Abraham, C.; Cho, J.H. Inflammatory bowel disease. N. Engl. J. Med. 2009, 361, 2066–2078.
  6. Van Der Kraak, L.; Gros, P.; Beauchemin, N. Colitis-associated colon cancer: Is it in your genes? World J. Gastroenterol. 2015, 21, 11688–11699.
  7. Choy, M.C.; Visvanathan, K.; De Cruz, P. An overview of the Innate and adaptive immune system in inflammatory bowel disease. Inflamm. Bowel Dis. 2017, 23, 2–13.
  8. Di Sabatino, A.; Biancheri, P.; Rovedatti, L.; MacDonald, T.T.; Corazza, G.R. New pathogenic paradigms in inflammatory bowel disease. Inflamm. Bowel Dis. 2012, 18, 368–371.
  9. Baumgart, D.C.; Sandborn, W.J. Crohn’s disease. Lancet 2012, 380, 1590–1605.
  10. Peeters, M.; Joossens, S.; Vermeire, S.; Vlietinck, R.; Bossuyt, X.; Rutgeerts, P. Diagnostic value of anti-saccharomyces cerevisiae and antineutrophil cytoplasmic autoantibodies in inflammatory bowel disease. Am. J. Gastroenterol. 2001, 96, 730–734.
  11. Lewis, J.D. The utility of biomarkers in the diagnosis and therapy of inflammatory bowel disease. Gastroenterology 2011, 140, 1817–1826.
  12. Torres, J.; Petralia, F.; Sato, T.; Wang, P.; Telesco, S.E.; Choung, R.S.; Strauss, R.; Li, X.J.; Laird, R.M.; Gutierrez, R.L.; et al. Serum biomarkers identify patients who will develop inflammatory bowel diseases up to 5 years before diagnosis. Gastroenterology 2020, 159, 96–104.
  13. Cichon, C.; Sabharwal, H.; Ruter, C.; Schmidt, M.A. MicroRNAs regulate tight junction proteins and modulate epithelial/endothelial barrier functions. Tissue Barriers 2014, 2, e944446.
  14. Weber, J.A.; Baxter, D.H.; Zhang, S.; Huang, D.Y.; Huang, K.H.; Lee, M.J.; Galas, D.J.; Wang, K. The microRNA spectrum in 12 body fluids. Clin. Chem. 2010, 56, 1733–1741.
  15. Krissansen, G.W.; Yang, Y.; McQueen, F.M.; Leung, E.; Peek, D.; Chan, Y.C.; Print, C.; Dalbeth, N.; Williams, M.; Fraser, A.G. Overexpression of miR-595 and miR-1246 in the sera of patients with active forms of inflammatory bowel disease. Inflamm. Bowel Dis. 2015, 21, 520–530.
  16. Jensen, M.D.; Andersen, R.F.; Christensen, H.; Nathan, T.; Kjeldsen, J.; Madsen, J.S. Circulating microRNAs as biomarkers of adult Crohn’s disease. Eur. J. Gastroenterol. Hepatol. 2015, 27, 1038–1044.
  17. Kalla, R.; Adams, A.T.; Ventham, N.T.; Kennedy, N.A.; White, R.; Clarke, C.; Ivens, A.; Bergemalm, D.; Vatn, S.; Lopez-Jimena, B.; et al. Whole blood profiling of T-cell derived miRNA allows the development of prognostic models in inflammatory bowel disease. J. Crohns Colitis 2020, 14, 1724–1733.
  18. Schaefer, J.S.; Attumi, T.; Opekun, A.R.; Abraham, B.; Hou, J.; Shelby, H.; Graham, D.Y.; Streckfus, C.; Klein, J.R. MicroRNA signatures differentiate Crohn’s disease from ulcerative colitis. BMC Immunol. 2015, 16, 5.
  19. Lin, J.; Welker, N.C.; Zhao, Z.; Li, Y.; Zhang, J.; Reuss, S.A.; Zhang, X.; Lee, H.; Liu, Y.; Bronner, M.P. Novel specific microRNA biomarkers in idiopathic inflammatory bowel disease unrelated to disease activity. Mod. Pathol. 2014, 27, 602–608.
  20. Konstantinidis, A.; Pardali, D.; Adamama-Moraitou, K.K.; Gazouli, M.; Dovas, C.I.; Legaki, E.; Brellou, G.D.; Savvas, I.; Jergens, A.E.; Rallis, T.S.; et al. Colonic mucosal and serum expression of microRNAs in canine large intestinal inflammatory bowel disease. BMC Vet. Res. 2020, 16, 69.
  21. Paraskevi, A.; Theodoropoulos, G.; Papaconstantinou, I.; Mantzaris, G.; Nikiteas, N.; Gazouli, M. Circulating microRNA in inflammatory bowel disease. J. Crohns Colitis 2012, 6, 900–904.
  22. Cordes, F.; Demmig, C.; Bokemeyer, A.; Bruckner, M.; Lenze, F.; Lenz, P.; Nowacki, T.; Tepasse, P.; Schmidt, H.H.; Schmidt, M.A.; et al. MicroRNA-320a monitors intestinal disease activity in patients with inflammatory bowel disease. Clin. Transl. Gastroenterol. 2020, 11, e00134.
  23. James, J.P.; Riis, L.B.; Malham, M.; Hogdall, E.; Langholz, E.; Nielsen, B.S. MicroRNA biomarkers in IBD-differential diagnosis and prediction of Colitis-associated cancer. Int. J. Mol. Sci. 2020, 21, 7893.
  24. Omidbakhsh, A.; Saeedi, M.; Khoshnia, M.; Marjani, A.; Hakimi, S. Micro-RNAs -106a and -362-3p in Peripheral blood of inflammatory bowel disease patients. Open. Biochem. J. 2018, 12, 78–86.
  25. Zahm, A.M.; Thayu, M.; Hand, N.J.; Horner, A.; Leonard, M.B.; Friedman, J.R. Circulating microRNA is a biomarker of pediatric Crohn disease. J. Pediatr. Gastroenterol. Nutr. 2011, 53, 26–33.
  26. Duttagupta, R.; DiRienzo, S.; Jiang, R.; Bowers, J.; Gollub, J.; Kao, J.; Kearney, K.; Rudolph, D.; Dawany, N.B.; Showe, M.K.; et al. Genome-wide maps of circulating miRNA biomarkers for ulcerative colitis. PLoS ONE 2012, 7, e31241.
  27. Wu, F.; Guo, N.J.; Tian, H.; Marohn, M.; Gearhart, S.; Bayless, T.M.; Brant, S.R.; Kwon, J.H. Peripheral blood microRNAs distinguish active ulcerative colitis and Crohn’s disease. Inflamm. Bowel Dis. 2011, 17, 241–250.
  28. Zahm, A.M.; Hand, N.J.; Tsoucas, D.M.; Le Guen, C.L.; Baldassano, R.N.; Friedman, J.R. Rectal microRNAs are perturbed in pediatric inflammatory bowel disease of the colon. J. Crohns Colitis 2014, 8, 1108–1117.
  29. Fasseu, M.; Treton, X.; Guichard, C.; Pedruzzi, E.; Cazals-Hatem, D.; Richard, C.; Aparicio, T.; Daniel, F.; Soule, J.C.; Moreau, R.; et al. Identification of restricted subsets of mature microRNA abnormally expressed in inactive colonic mucosa of patients with inflammatory bowel disease. PLoS ONE 2010, 5, e13160.
  30. Coskun, M.; Bjerrum, J.T.; Seidelin, J.B.; Troelsen, J.T.; Olsen, J.; Nielsen, O.H. MiR-20b, miR-98, miR-125b-1*, and let-7e* as new potential diagnostic biomarkers in ulcerative colitis. World J. Gastroenterol. 2013, 19, 4289–4299.
  31. Wu, F.; Zikusoka, M.; Trindade, A.; Dassopoulos, T.; Harris, M.L.; Bayless, T.M.; Brant, S.R.; Chakravarti, S.; Kwon, J.H. MicroRNAs are differentially expressed in ulcerative colitis and alter expression of macrophage inflammatory peptide-2 alpha. Gastroenterology 2008, 135, 1624–1635.
  32. Wu, F.; Zhang, S.; Dassopoulos, T.; Harris, M.L.; Bayless, T.M.; Meltzer, S.J.; Brant, S.R.; Kwon, J.H. Identification of microRNAs associated with ileal and colonic Crohn’s disease. Inflamm. Bowel Dis. 2010, 16, 1729–1738.
  33. Koukos, G.; Polytarchou, C.; Kaplan, J.L.; Oikonomopoulos, A.; Ziring, D.; Hommes, D.W.; Wahed, R.; Kokkotou, E.; Pothoulakis, C.; Winter, H.S.; et al. A microRNA signature in pediatric ulcerative colitis: Deregulation of the miR-4284/CXCL5 pathway in the intestinal epithelium. Inflamm. Bowel Dis. 2015, 21, 996–1005.
  34. Peck, B.C.; Weiser, M.; Lee, S.E.; Gipson, G.R.; Iyer, V.B.; Sartor, R.B.; Herfarth, H.H.; Long, M.D.; Hansen, J.J.; Isaacs, K.L.; et al. MicroRNAs classify different disease behavior phenotypes of Crohn’s Disease and may have prognostic utility. Inflamm. Bowel Dis. 2015, 21, 2178–2187.
  35. Van der Goten, J.; Vanhove, W.; Lemaire, K.; Van Lommel, L.; Machiels, K.; Wollants, W.J.; De Preter, V.; De Hertogh, G.; Ferrante, M.; Van Assche, G.; et al. Integrated miRNA and mRNA expression profiling in inflamed colon of patients with ulcerative colitis. PLoS ONE 2014, 9, e116117.
  36. Chen, Y.; Wang, C.; Liu, Y.; Tang, L.; Zheng, M.; Xu, C.; Song, J.; Meng, X. MiR-122 targets NOD2 to decrease intestinal epithelial cell injury in Crohn’s disease. Biochem. Biophys. Res. Commun. 2013, 438, 133–139.
  37. Pathak, S.; Grillo, A.R.; Scarpa, M.; Brun, P.; D’Inca, R.; Nai, L.; Banerjee, A.; Cavallo, D.; Barzon, L.; Palu, G.; et al. MiR-155 modulates the inflammatory phenotype of intestinal myofibroblasts by targeting SOCS1 in ulcerative colitis. Exp. Mol. Med. 2015, 47, e164.
  38. Liu, Y.; Zhu, F.; Li, H.; Fan, H.; Wu, H.; Dong, Y.; Chu, S.; Tan, C.; Wang, Q.; He, H.; et al. MiR-155 contributes to intestinal barrier dysfunction in DSS-induced mice colitis via targeting HIF-1alpha/TFF-3 axis. Aging 2020, 12, 14966–14977.
  39. Chen, G.; Cao, S.; Liu, F.; Liu, Y. miR-195 plays a role in steroid resistance of ulcerative colitis by targeting Smad7. Biochem. J. 2015, 471, 357–367.
  40. Fujioka, S.; Nakamichi, I.; Esaki, M.; Asano, K.; Matsumoto, T.; Kitazono, T. Serum microRNA levels in patients with Crohn’s disease during induction therapy by infliximab. J. Gastroenterol. Hepatol. 2014, 29, 1207–1214.
  41. Cao, B.; Zhou, X.; Ma, J.; Zhou, W.; Yang, W.; Fan, D.; Hong, L. Role of MiRNAs in inflammatory bowel disease. Dig. Dis. Sci. 2017, 62, 1426–1438.
  42. McKenna, L.B.; Schug, J.; Vourekas, A.; McKenna, J.B.; Bramswig, N.C.; Friedman, J.R.; Kaestner, K.H. MicroRNAs control intestinal epithelial differentiation, architecture, and barrier function. Gastroenterology 2010, 139, 1654–1664.
  43. Mizoguchi, E.; Low, D.; Ezaki, Y.; Okada, T. Recent updates on the basic mechanisms and pathogenesis of inflammatory bowel diseases in experimental animal models. Intest. Res. 2020, 18, 151–167.
  44. Jiminez, J.A.; Uwiera, T.C.; Douglas Inglis, G.; Uwiera, R.R. Animal models to study acute and chronic intestinal inflammation in mammals. Gut. Pathog. 2015, 7, 29.
  45. Bazzoni, F.; Rossato, M.; Fabbri, M.; Gaudiosi, D.; Mirolo, M.; Mori, L.; Tamassia, N.; Mantovani, A.; Cassatella, M.A.; Locati, M. Induction and regulatory function of miR-9 in human monocytes and neutrophils exposed to proinflammatory signals. Proc. Natl. Acad. Sci. USA 2009, 106, 5282–5287.
  46. Qi, J.; Qiao, Y.; Wang, P.; Li, S.; Zhao, W.; Gao, C. MicroRNA-210 negatively regulates LPS-induced production of proinflammatory cytokines by targeting NF-kappaB1 in murine macrophages. FEBS Lett. 2012, 586, 1201–1207.
  47. Brain, O.; Owens, B.M.; Pichulik, T.; Allan, P.; Khatamzas, E.; Leslie, A.; Steevels, T.; Sharma, S.; Mayer, A.; Catuneanu, A.M.; et al. The intracellular sensor NOD2 induces microRNA-29 expression in human dendritic cells to limit IL-23 release. Immunity 2013, 39, 521–536.
  48. Ghorpade, D.S.; Sinha, A.Y.; Holla, S.; Singh, V.; Balaji, K.N. NOD2-nitric oxide-responsive microRNA-146a activates Sonic hedgehog signaling to orchestrate inflammatory responses in murine model of inflammatory bowel disease. J. Biol. Chem. 2013, 288, 33037–33048.
  49. Chuang, A.Y.; Chuang, J.C.; Zhai, Z.; Wu, F.; Kwon, J.H. NOD2 expression is regulated by microRNAs in colonic epithelial HCT116 cells. Inflamm. Bowel Dis. 2014, 20, 126–135.
  50. Fujita, S.; Ito, T.; Mizutani, T.; Minoguchi, S.; Yamamichi, N.; Sakurai, K.; Iba, H. MiR-21 Gene expression triggered by AP-1 is sustained through a double-negative feedback mechanism. J. Mol. Biol. 2008, 378, 492–504.
  51. Sheedy, F.J. Turning 21: Induction of miR-21 as a Key Switch in the inflammatory response. Front. Immunol. 2015, 6, 19.
  52. Barrett, J.C.; Hansoul, S.; Nicolae, D.L.; Cho, J.H.; Duerr, R.H.; Rioux, J.D.; Brant, S.R.; Silverberg, M.S.; Taylor, K.D.; Barmada, M.M.; et al. Genome-wide association defines more than 30 distinct susceptibility loci for Crohn’s disease. Nat. Genet. 2008, 40, 955–962.
  53. Yang, Y.; Ma, Y.; Shi, C.; Chen, H.; Zhang, H.; Chen, N.; Zhang, P.; Wang, F.; Yang, J.; Yang, J.; et al. Overexpression of miR-21 in patients with ulcerative colitis impairs intestinal epithelial barrier function through targeting the Rho GTPase RhoB. Biochem. Biophys. Res. Commun. 2013, 434, 746–752.
  54. Adams, A.T.; Kennedy, N.A.; Hansen, R.; Ventham, N.T.; O’Leary, K.R.; Drummond, H.E.; Noble, C.L.; El-Omar, E.; Russell, R.K.; Wilson, D.C.; et al. Two-stage genome-wide methylation profiling in childhood-onset Crohn’s Disease implicates epigenetic alterations at the VMP1/MIR21 and HLA loci. Inflamm. Bowel Dis. 2014, 20, 1784–1793.
  55. Shi, C.; Liang, Y.; Yang, J.; Xia, Y.; Chen, H.; Han, H.; Yang, Y.; Wu, W.; Gao, R.; Qin, H. MicroRNA-21 knockout improve the survival rate in DSS induced fatal colitis through protecting against inflammation and tissue injury. PLoS ONE 2013, 8, e66814.
  56. Johnston, D.G.W.; Williams, M.A.; Thaiss, C.A.; Cabrera-Rubio, R.; Raverdeau, M.; McEntee, C.; Cotter, P.D.; Elinav, E.; O’Neill, L.A.J.; Corr, S.C. Loss of microRNA-21 influences the gut microbiota, causing reduced susceptibility in a murine model of colitis. J. Crohns Colitis 2018, 12, 835–848.
  57. Ando, Y.; Mazzurana, L.; Forkel, M.; Okazaki, K.; Aoi, M.; Schmidt, P.T.; Mjosberg, J.; Bresso, F. Downregulation of microRNA-21 in Colonic CD3+ T cells in UC remission. Inflamm. Bowel Dis. 2016, 22, 2788–2793.
  58. Singh, U.P.; Murphy, A.E.; Enos, R.T.; Shamran, H.A.; Singh, N.P.; Guan, H.; Hegde, V.L.; Fan, D.; Price, R.L.; Taub, D.D.; et al. MiR-155 deficiency protects mice from experimental colitis by reducing T helper type 1/type 17 responses. Immunology 2014, 143, 478–489.
  59. He, C.; Yu, T.; Shi, Y.; Ma, C.; Yang, W.; Fang, L.; Sun, M.; Wu, W.; Xiao, F.; Guo, F.; et al. MicroRNA 301A promotes intestinal inflammation and colitis-associated cancer development by inhibiting BTG1. Gastroenterology 2017, 152, 1434–1448.
  60. Sohn, J.J.; Schetter, A.J.; Yfantis, H.G.; Ridnour, L.A.; Horikawa, I.; Khan, M.A.; Robles, A.I.; Hussain, S.P.; Goto, A.; Bowman, E.D.; et al. Macrophages, nitric oxide and microRNAs are associated with DNA damage response pathway and senescence in inflammatory bowel disease. PLoS ONE 2012, 7, e44156.
  61. Muenchau, S.; Deutsch, R.; de Castro, I.J.; Hielscher, T.; Heber, N.; Niesler, B.; Lusic, M.; Stanifer, M.L.; Boulant, S. Hypoxic environment promotes barrier formation in human intestinal epithelial cells through regulation of microRNA 320a expression. Mol. Cell Biol. 2019, 39, e00553.
  62. Wu, F.; Dong, F.; Arendovich, N.; Zhang, J.; Huang, Y.; Kwon, J.H. Divergent influence of microRNA-21 deletion on murine colitis phenotypes. Inflamm. Bowel Dis. 2014, 20, 1972–1985.
  63. Jeker, L.T.; Zhou, X.; Gershberg, K.; de Kouchkovsky, D.; Morar, M.M.; Stadthagen, G.; Lund, A.H.; Bluestone, J.A. MicroRNA 10a marks regulatory T cells. PLoS ONE 2012, 7, e36684.
  64. de Kouchkovsky, D.; Esensten, J.H.; Rosenthal, W.L.; Morar, M.M.; Bluestone, J.A.; Jeker, L.T. MicroRNA-17-92 regulates IL-10 production by regulatory T cells and control of experimental autoimmune encephalomyelitis. J. Immunol. 2013, 191, 1594–1605.
  65. Lu, L.F.; Boldin, M.P.; Chaudhry, A.; Lin, L.L.; Taganov, K.D.; Hanada, T.; Yoshimura, A.; Baltimore, D.; Rudensky, A.Y. Function of miR-146a in controlling Treg cell-mediated regulation of Th1 responses. Cell 2010, 142, 914–929.
  66. Chinen, I.; Nakahama, T.; Kimura, A.; Nguyen, N.T.; Takemori, H.; Kumagai, A.; Kayama, H.; Takeda, K.; Lee, S.; Hanieh, H.; et al. The aryl hydrocarbon receptor/microRNA-212/132 axis in T cells regulates IL-10 production to maintain intestinal homeostasis. Int. Immunol. 2015, 27, 405–415.
  67. Steiner, D.F.; Thomas, M.F.; Hu, J.K.; Yang, Z.; Babiarz, J.E.; Allen, C.D.; Matloubian, M.; Blelloch, R.; Ansel, K.M. MicroRNA-29 regulates T-box transcription factors and interferon-gamma production in helper T cells. Immunity 2011, 35, 169–181.
  68. Guerau-de-Arellano, M.; Smith, K.M.; Godlewski, J.; Liu, Y.; Winger, R.; Lawler, S.E.; Whitacre, C.C.; Racke, M.K.; Lovett-Racke, A.E. Micro-RNA dysregulation in multiple sclerosis favours pro-inflammatory T-cell-mediated autoimmunity. Brain 2011, 134, 3578–3589.
  69. Banerjee, A.; Schambach, F.; DeJong, C.S.; Hammond, S.M.; Reiner, S.L. Micro-RNA-155 inhibits IFN-gamma signaling in CD4+ T cells. Eur. J. Immunol. 2010, 40, 225–231.
  70. Takahashi, H.; Kanno, T.; Nakayamada, S.; Hirahara, K.; Sciume, G.; Muljo, S.A.; Kuchen, S.; Casellas, R.; Wei, L.; Kanno, Y.; et al. TGF-beta and retinoic acid induce the microRNA miR-10a, which targets Bcl-6 and constrains the plasticity of helper T cells. Nat. Immunol. 2012, 13, 587–595.
  71. O’Connell, R.M.; Kahn, D.; Gibson, W.S.; Round, J.L.; Scholz, R.L.; Chaudhuri, A.A.; Kahn, M.E.; Rao, D.S.; Baltimore, D. MicroRNA-155 promotes autoimmune inflammation by enhancing inflammatory T cell development. Immunity 2010, 33, 607–619.
  72. Du, C.; Liu, C.; Kang, J.; Zhao, G.; Ye, Z.; Huang, S.; Li, Z.; Wu, Z.; Pei, G. MicroRNA miR-326 regulates TH-17 differentiation and is associated with the pathogenesis of multiple sclerosis. Nat. Immunol. 2009, 10, 1252–1259.
  73. Mycko, M.P.; Cichalewska, M.; Machlanska, A.; Cwiklinska, H.; Mariasiewicz, M.; Selmaj, K.W. MicroRNA-301a regulation of a T-helper 17 immune response controls autoimmune demyelination. Proc. Natl. Acad. Sci. USA 2012, 109, E1248–E1257.
  74. Wang, H.; Flach, H.; Onizawa, M.; Wei, L.; McManus, M.T.; Weiss, A. Negative regulation of Hif1a expression and TH17 differentiation by the hypoxia-regulated microRNA miR-210. Nat. Immunol. 2014, 15, 393–401.
  75. Koralov, S.B.; Muljo, S.A.; Galler, G.R.; Krek, A.; Chakraborty, T.; Kanellopoulou, C.; Jensen, K.; Cobb, B.S.; Merkenschlager, M.; Rajewsky, N.; et al. Dicer ablation affects antibody diversity and cell survival in the B lymphocyte lineage. Cell 2008, 132, 860–874.
  76. Chen, C.Z.; Li, L.; Lodish, H.F.; Bartel, D.P. MicroRNAs modulate hematopoietic lineage differentiation. Science 2004, 303, 83–86.
  77. Suri, K.; Bubier, J.A.; Wiles, M.V.; Shultz, L.D.; Amiji, M.M.; Hosur, V. Role of microRNA in inflammatory bowel disease: Clinical evidence and the development of preclinical animal models. Cells 2021, 10, 2204.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , ,
View Times: 471
Revisions: 2 times (View History)
Update Date: 24 Apr 2022
1000/1000
Video Production Service