Next Article in Journal
Safety Profile of the Trastuzumab-Based ADCs: Analysis of Real-World Data Registered in EudraVigilance
Previous Article in Journal
The Congestion “Pandemic” in Acute Heart Failure Patients
Previous Article in Special Issue
T Cell-Induced Colitis Is Exacerbated by Prolonged Stress: A Comparison in Male and Female Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cytokine Profile in Predicting the Effectiveness of Advanced Therapy for Ulcerative Colitis: A Narrative Review

by
Hiroki Kurumi
1,2,
Yoshihiro Yokoyama
1,
Takehiro Hirano
1,
Kotaro Akita
1,
Yuki Hayashi
1,
Tomoe Kazama
1,
Hajime Isomoto
2 and
Hiroshi Nakase
1,*
1
Department of Gastroenterology and Hepatology, Sapporo Medical University School of Medicine, S-1, W-16, Chuo-ku, Sapporo 060-8543, Hokkaido, Japan
2
Division of Gastroenterology and Nephrology, Department of Multidisciplinary Internal Medicine, Tottori University Faculty of Medicine, 36-1, Nishi-cho, Yonago 683-8504, Tottori, Japan
*
Author to whom correspondence should be addressed.
Biomedicines 2024, 12(5), 952; https://doi.org/10.3390/biomedicines12050952
Submission received: 31 March 2024 / Revised: 19 April 2024 / Accepted: 23 April 2024 / Published: 25 April 2024

Abstract

:
Cytokine-targeted therapies have shown efficacy in treating patients with ulcerative colitis (UC), but responses to these advanced therapies can vary. This variability may be due to differences in cytokine profiles among patients with UC. While the etiology of UC is not fully understood, abnormalities of the cytokine profiles are deeply involved in its pathophysiology. Therefore, an approach focused on the cytokine profile of individual patients with UC is ideal. Recent studies have demonstrated that molecular analysis of cytokine profiles in UC can predict response to each advanced therapy. This narrative review summarizes the molecules involved in the efficacy of various advanced therapies for UC. Understanding these associations may be helpful in selecting optimal therapeutic agents.

1. Introduction

Ulcerative colitis (UC) is a chronic, remitting/relapsing inflammatory disease of the intestinal tract that requires lifelong monitoring and treatment [1]. Although the etiology of UC is not fully understood, abnormalities in the cytokine network contribute to the pathophysiology of UC. Several inflammatory bowel disease (IBD) risk loci are located in regions encoding cytokines and their downstream signaling mediators [2,3,4]. Therefore, advanced therapies such as cytokine-targeted therapies for UC have been developed and clinically used [5,6].
These advanced therapies include (1) anti-tumor necrosis factor (TNF)-α antibodies (infliximab, adalimumab, and golimumab), anti-interleukin (IL)-12/23p40 antibodies (ustekinumab), and anti-IL-23p19 antibodies (mirikizumab and guselkumab) that neutralize inflammatory cytokines; (2) Janus kinase (JAK) inhibitors (tofacitinib, filgotinib, and upadacitinib) that block downstream signaling of cytokine receptors; and (3) anti-integrins (vedolizumab and carotegrast) and sphingosine-1-phosphate (S1P) modulators (ozanimod) that inhibit migration of effector immune cells [7,8,9,10,11]. These therapies can directly or indirectly alter key inflammation pathways, receptors, and some crucial adhesion molecules [12]. They have shown efficacy in treating UC [13], but there are differences in how patients respond to each treatment. For instance, up to 40% of patients with UC do not respond to treatment with anti-TNF-α therapy [14]. This might reflect the fact that individual patients with UC have different cytokine profiles [15]. Several reports have explored the association between differences in cytokine profiles among individual patients with UC and therapeutic efficacy [16,17]. Multiple factors, including genetic background, environmental factors, luminal factors, and intestinal microbiota, are intricately involved in the etiology of UC, which could contribute to individual cytokine profile heterogeneity [18,19]. Therefore, it is reasonable that an approach focused on the cytokine profile of individual patients with UC would enhance the likelihood of success, minimize off-target effects, and produce sustained effects of cytokine-targeted therapy [20]. For instance, recent studies have demonstrated that Oncostatin M (OSM) is not only correlated with the severity of UC but can also predict the response to treatment with anti-TNF-α antibodies [21,22]. However, there is currently no evidence that individual cytokine profiles contribute to optimal biologics selection.
This narrative review provides a summary from the perspective of cytokine profiles related to intestinal homeostasis and the pathophysiology of UC, and previous reports on cytokine profiles that predict therapeutic efficacy.

2. Role of Cytokines in the Pathophysiology of UC

2.1. Intestinal Homeostasis and Cytokinesis

Intestinal epithelial cells (IECs), immune cells, gut microbiota, and mesenchymal cells orchestrate the maintenance of intestinal homeostasis, which is regulated by cytokines.
T helper (Th) 17 cells and type 3 innate lymphocytes (ILC3) maintain intestinal epithelial function by producing IL-6, IL-17, and IL-22. IL-6 and IL-22 promote IEC survival and proliferation through activation of signal transducer and activator of transcription (STAT) 3 signaling [23]. IL-17 promotes antimicrobial peptide secretion and IEC tight junction formation [24,25]. Gut microbiota stimulation triggers mononuclear phagocytes to produce IL-1β, IL-6, and IL-23, which in turn promotes IL-22 and IL-17 production by Th17 cells [26,27]. Dietary metabolites like retinoic acid (RA) and aryl hydrocarbon receptor (AHR) ligands also stimulate IL-22 production by ILC3 [28,29].
Type 1 interferon (IFN) produced by mononuclear phagocytes enhances barrier integrity through STAT1 and STAT2 signaling and promotes anti-inflammatory cytokine production by T regulatory (Treg) cells [30].
When IECs are damaged, macrophages are activated to promote the repair of the injured tissue. Classically activated M1 macrophages differentiate into M2 macrophages through STAT6 activation. M2 macrophages directly activate Wnt signaling pathway by expressing Wnt ligands, facilitating tissue repair [31,32].
Treg cells are essential for regulating excessive immune responses. IL-10 is a crucial anti-inflammatory cytokine secreted by Treg cells [33,34]. IL-10 inhibits both antigen presentation and subsequent release of pro-inflammatory cytokines. IL-10 gene polymorphisms are associated with early-onset and severe intestinal inflammation [33]. Recently, the transcription factor cellular musculoaponeurotic fibrosarcoma (c-Maf) was shown to promote IL-10 production in forkhead box P3 (Foxp3)+ Treg and T regulatory type 1 (Tr1) cells [35]. Foxp3+ Treg cells express numerous IL-2 receptors, and IL-2 promotes Treg cell differentiation in the thymus and maintenance in the periphery [36]. IL-33 directly acts on Foxp3+ Treg cells, promoting their accumulation [37].
The intestinal epithelium is coated by secreted immunoglobulin (Ig) A to prevent pathogenic bacteria from interacting directly with the intestinal epithelium [38]. B cell activating factor (BAFF), IL-10, IL-6, and transforming growth factor (TGF)-β promote IgA secretion from plasma cells [39,40].
The interaction between mononuclear phagocytes, ILCs, T cells, and B cells is an important host adaptation to microbial stimuli, and its deficiency could be linked to the onset of IBD (Figure 1).

2.2. Pathophysiology of UC

The immune system is divided into two main systems: innate and adaptive immunity. Innate immune cells respond rapidly and non-specifically to pathogens or other foreign entities. Once activated, they cause inflammation by releasing cytokines and chemokines and phagocytizing pathogens and cellular debris. Adaptive immunity takes longer to activate than innate immunity due to its dependence on antigen presentation and the cytokines produced by the innate immune response, but its response is specific. Both innate and adaptive immune cells have been shown to be significantly involved in the pathophysiology of UC [41].

2.2.1. Innate Immunity

Disruption of the epithelial barrier function and microbial dysbiosis are important triggers in the pathophysiology of UC. This process involves the activation of the innate immune system by damage-associated molecular patterns (DAMPs) and pathogen-associated molecular patterns (PAMPs), which interact with pattern recognition receptors (PRRs) on IECs and immune cells. Immune cells such as neutrophils and mononuclear phagocytes and inflammatory cytokines like IL-1β, TNF-α, and IL-6 play crucial roles in the downstream of these responses.
Neutrophils are recruited from circulating blood to inflamed tissues by cytokines such as IL-1β, IL-6, and TNF-α and chemokines like C-X-C motif ligand (CXCL) 5 and IL-8 [42,43,44]. Abnormal activation of neutrophils has been observed in patients with UC [45]. Neutrophils directly promote tissue damage by releasing proteases such as matrix metalloproteinases (MMPs) and neutrophil elastases, and reactive oxygen species (ROS) [46]. The accumulation of activated neutrophils promotes changes in cryptic structures and the formation of crypt abscesses, mediated by an unbalanced enzymatic response, generation of TNF-α and IL-1β, and secretion of calprotectin [47,48,49].
In UC, lymphocyte antigen 6 complex (Ly6C)high monocytes are increased at inflammatory sites. Their migration is regulated by the C-C motif chemokine receptor (CCR)2, IL-8 and TGF-β, which are constitutively produced by IECs [50,51,52]. Ly6Chigh monocytes are highly susceptible to bacteria due to elevated expression of Toll-like receptor (TLR)2 and nod-like receptor (NOD)2 [53]. Additionally, C-X3-C motif chemokine receptor (CX3CR)int macrophages with upregulated TLR2 also accumulate at inflammatory sites in UC [50,54]. These macrophages have an increased ability to secrete inflammatory cytokines such as TNF-α, IL-1β, and IL-6 [55]. Abnormally activated macrophages promote myofibroblast-mediated fibrosis by producing TGF-β1, connective tissue growth factor (CTGF), and fibroblast activating protein (FAP) [56].
Dendritic cells (DCs) link the innate and adaptive immune systems. Plasmacytoid DCs secrete type Ⅰ IFN, X-C motif chemokine receptor 1 (XCR1)+ myeloid DC1s are superior in antigen presentation to cytotoxic T cells, and signal regulatory protein α (SIPRα)+ myeloid DC2s polarize cluster of differentiation (CD)4+ T cell responses. Myeloid DCs are the dominant subtype in the intestinal lamina propria [57]. Activated DCs secrete inflammatory cytokines such as IL-1β, IL-18, IL-6, IL-12, and IL-23 to activate adaptive immunity. In UC, DCs exhibit an abnormal, immature phenotype with decreased expression of cutaneous lymphocyte antigen (CLA) and CCR4 and increased expression of CCR9 and β7 integrin [58,59,60]. Immature CD11c+CD11b+ myeloid DCs produce IL-23, which could sustain colitis [61]. The circulating cells in the peripheral blood of patients with active UC contain plasmacytoid DCs that migrate to secondary lymphoid organs, where they produce IL-6, IL-8, and TNF-α to perpetuate the disease [62].
ILCs are derived from common lymphoid progenitor (CLP) but are classified under the innate immune system. ILCs respond rapidly to signals and cytokines and act early in immune responses [63,64]. Retinoic acid receptor-related orphan receptor (ROR)γt+ ILC3 produces IL-17A and IFN-γ, promoting the pathogenesis of T cell-independent colitis [65]. Conversely, the frequency of natural cytotoxicity receptor (NCR)+ ILC3 producing IL-22 is reduced at inflammatory sites in patients with UC [66,67,68]. RORγt ILC1 promotes colitis through the production of IFN-γ and TNF-α [69,70]. ILC2 produces type 2 cytokines like IL-5 and IL-13, which contribute to epithelial barrier function and antigen clearance in the lumen but also play a role in chronic inflammation and tissue fibrosis [71].

2.2.2. Adaptive Immunity

Under steady-state conditions, there are only a small number of CD4+ T cells in the intestinal epithelium. However, in patients with active UC, there is a significant increase in the number of CD4+ T cells in the intestinal epithelium [72]. Infiltration of T cells and accumulation of cytokines associated with T cells at inflammatory sites are common features of patients with UC [73,74,75]. Naïve CD4+ T cells are activated by antigen-presenting cells (APCs) in lymphoid tissues. Upon activation, these cells upregulate homing receptors, such as α4β7 integrins. Naïve CD4+ T cells differentiate into different subsets, including Th1, Th2, Th9, Th17, Th22, Treg, and Tr1 cells, depending on the microenvironment of various cytokines and transcription factors [72]. The cytokines produced by these subsets have multifaceted functions and complicatedly influence the pathophysiology of UC.
Several murine colitis models are characterized by excessive IL-12 and IL-23 production and decreased IL-10 and TGF-β production [76,77]. Both IL-12 and IL-23 are secreted by DCs and macrophages in response to early innate signals [78,79]. IL-12 induces Th1 cell differentiation and promotes IFN-γ and TNF-α production [80]. The identification of the IL-23 receptor (IL-23R) as a susceptibility locus for IBD and the clinical utility of anti-IL-23 antibodies confirm IL-23′s involvement in the pathophysiology of UC [81,82]. Prolonged IL-23 production in both humans and murine models converts barrier-promoting Th17 cells into pathogenic Th17 cells, leading to increased secretion of multiple cytokines, including IFN-γ, TNF-α, IL-17A, and granulocyte–macrophage colony-stimulating factor (GM-CSF) [83,84,85]. IL-23 also promotes intestinal inflammation by suppressing Foxp3+ Treg cells [86]. IL-1β and IL-23 collaborate to induce IL-17 production by Th17 cells and ILCs, promoting pathogenic responses [87]. The efficacy of anti-IL-12/23p40 antibodies against UC indicates that Th1 and Th17 cells are involved in the pathophysiology of UC [88].
On the other hand, Th2 and Th9 cells are abnormally activated at inflammatory sites in UC, leading to increased expression of their major cytokines, including IL-5, IL-13, and IL-9 [89,90]. Th2 cells are involved in extracellular microorganism elimination and support IgE-mediated B cell responses by secreting IL-4, IL-5, and IL-13, but their overactivation can contribute to chronic inflammation [72]. IL-5 is involved in the differentiation of eosinophils [91]. IL-33, a Th2 cytokine, is elevated at inflammatory sites in patients with UC compared to healthy controls [90,92,93,94]. Furthermore, blocking IL-33/suppression of tumorigenicity 2 (ST2) signaling has been shown to decrease disease activity, suggesting a pathogenic role for IL-33 [95].
IL-9 is a pro-inflammatory cytokine that activates JAK1 and JAK3 signaling upon binding to its receptor. Excessive IL-9 production in the intestinal tract can impair resistance to commensal bacteria by compromising the integrity of the epithelial barrier, leading to inflammation [96]. In addition, IL-9 promotes early population expansion of memory B cells and the production of IgG and IgE by B cells [97,98,99].
B cells can be classified into regulatory B cells secreting IL-10 and effector B cells secreting antibodies and various cytokines [100]. Regulatory B cells are decreased in the blood and intestinal tissues of patients with UC. Effector B cells cause inflammation by presenting antigens to T cells and secreting IL-2, IL-4, IFN-γ, TGF-β, and GM-CSF [101]. In contrast to the predominance of IgA in the intestinal mucosa of healthy individuals, IgG is predominant in the inflammatory mucosa of patients with UC [102,103]. The decreased IgA and increased IgG may be involved in the pathophysiology of UC [98]. IgG has a high affinity for antigens and activates the complement system. By binding to the fragment crystallizable (Fc)γ receptor, IgG promotes immune cell migration and maturation [101].
Indigenous bacterial-specific IgG antibodies are increased in the inflamed mucosa of patients with UC. In murine models, the induction of these antibodies induces intestinal inflammation by activating macrophage and Th17 cells and promoting neutrophil migration [104]. Fc receptor neonatal (FcRn) maintains IgG concentration by recycling IgG extracellularly. In a mouse model of dextran sodium sulfate (DSS)-induced colitis, specific inhibition of FcRn has been demonstrated to lower IgG concentrations and ameliorate colitis [105].
Both innate immune cells, including neutrophils, monocytes, macrophages, DCs, and ILCs, and adaptive immune cells, including T cells and B cells, are intricately involved in the pathophysiology of UC (Figure 2).

3. Mechanisms of Biologics and Small Molecule Compounds

3.1. Anti-TNF-α Antibodies

TNF-α is synthesized as a transmembrane TNF (mTNF), from which soluble TNF (sTNF) is released. sTNF binds preferentially to TNF receptor (TNFR) 1, while mTNF binds preferentially to TNFR2 [106]. TNFR1 is ubiquitously expressed and has a cell death domain, while TNFR2 is primarily expressed on lymphocytes and endothelial cells and lacks a cell death domain [107]. TNF-α primarily causes receptor-interacting serine/threonine-protein kinase (RIPK)1/3-dependent cell death via TNFR1 [108,109]. The TNFR2 pathway leads to the production of inflammatory cytokines such as IL-1β and IL-6, which have anti-apoptotic effects [110]. Anti-TNF-α antibodies inhibit TNFR1 signaling by neutralizing sTNF and TNF-α production by binding mTNF, which induces cell apoptosis through antibody-dependent cellular cytotoxicity, complement-dependent cytotoxicity, and outer-to-inner signaling. Anti-TNF-α antibodies also indirectly increase regulatory macrophages [111] and Foxp3+ Treg cells suppressing Th17 cells [112].

3.2. Anti-IL-12/23p40 Antibodies and Anti-IL-23p19 Antibodies

IL-12 and IL-23 are primarily produced by APCs. IL-12 acts on naïve T cells and promotes their differentiation into Th1 cells that produce TNF-α and IFN-γ [113]. IL-12 also induces the release of IFN-γ and TNF-α from CD8+ T cells, natural killer (NK) cells, and ILC1. Furthermore, IL-12 activates antimicrobial responses by DCs and macrophages [114]. IL-23 contributes to Th17 proliferation and stabilization and promotes the secretion of IL-17, TNF-α, and IL-22 by Th17 cells [115,116]. IL-23 also induces the secretion of IL-17 and IL-22 by γδ T cells and ILC3 [114] and the secretion of inflammatory cytokines such as TNF-α and IL-1β by binding to IL-23Rs on macrophages [117,118]. In addition, IL-23 suppresses IL-10 production from Treg cells, thereby reducing intestinal barrier and defense functions. Furthermore, experiments in mice have shown that IL-23 contributes to the differentiation, proliferation, and maintenance of Th2 cells [6].
Anti-IL-12/23p40 antibodies target p40, a common subunit of IL-12 and IL-23, and exert their anti-inflammatory effect by inhibiting both IL-12 and IL-23 signaling. Conversely, anti-IL-23p19 antibodies target p19, a subunit specific to IL-23, inhibiting only IL-23 without affecting other IL-12 family members [114].
Anti-IL-23p19 antibodies may be safer than anti-IL-12/23p40 antibodies due to the preservation of Th1 immune responses against infection and malignancies [119]. Although both anti-IL-12/23p40 and anti-IL-23p19 antibodies have been shown to alleviate experimental colitis to the same extent [25], there is no evidence suggesting a difference in the effects of these two antibodies in clinical practice. Basic research in mouse colitis models and clinical studies in patients with UC have shown that cytokine profiles differ depending on the disease stage of UC [6]. Th1- and Th17-cytokines are primarily involved in the pathogenesis of UC in the early phase, while Th2- and Th17-cytokines are primarily involved in the late phase. These differences could influence the choice of antibody treatment.
In addition, differences in binding affinity to target cytokines and specific effects via the Fc region of antibodies have also been reported. For example, guselkumab more effectively suppresses IL-23 secretion from CD64+ monocytes by binding to CD64 via the Fc region [120]. These differences in the molecular properties of antibodies may guide clinicians in selecting the most appropriate antibody for treatment.

3.3. Anti-α4β7 Integrin Antibodies

Anti-α4β7 integrin antibodies recognize a conformational epitope of the α4β7 integrin heterodimer. The transmembrane cell adhesion protein α4β7 is expressed on immune cells, including T cells, B cells, eosinophils, and ILC [121,122]. The ligand for α4β7, mucosal addressin cell adhesion molecule 1 (MAdCAM-1), is primarily found on endothelial cells within the gastrointestinal track and gut-associated lymph tissue [123,124]. Naïve T cells migrate to lymphoid tissues via high endothelial venules (HEVs) [125]. T cells activated by DCs or antigens in lymph nodes are transported from efferent lymphatic vessels through the S1P-S1P receptor (S1PR) signaling pathway [126]. Activated immune cells expressing α4β7 migrate from HEV-like vessels to intestinal tissues via the α4β7 and MAdCAM-1 signaling pathways [127,128]. Anti-α4β7 integrin antibodies suppress mucosal inflammation by inhibiting the migration of immune cells to the mucosal lamina propria.

3.4. JAK Inhibitors

Cytokines bind to specific receptors, causing activation and initiation of intracellular signaling pathways in various cell types [129]. JAK is a tyrosine kinase that binds to the intracellular domain of cytokine receptors [130]. When extracellular ligands bind to cytokine receptors, the JAK-STAT pathway is activated, and signals are transduced. The JAK family includes JAK1, JAK2, JAK3, and tyrosine kinase 2 (TYK2), with different combinations for each cytokine receptor [131,132,133]. In IBD field, currently available JAK inhibitors include tofacitinib, filgotinib, and upadacitinib, each with selectivity for different JAK families [5]. Compared to approaches that use antibodies to inhibit a single cytokine (e.g., TNF-α), JAK inhibitors have the potential to affect multiple cytokine-dependent pathways involved in the pathophysiology of UC (Table 1).

4. Molecules Predicting Therapeutic Efficacy

Previous reports on cytokine profiles predicting the therapeutic efficacy of each biologic in patients with UC are summarized in Table 2.

4.1. Anti-TNF-α Antibodies

Recent studies have demonstrated several mechanisms of resistance to treatment with anti-TNF-α antibodies. The mechanisms are briefly summarized below (Figure 3).

4.1.1. Oncostatin M

OSM is a member of the IL-6 cytokine family. Oncostatin M receptor (OSMR) is expressed primarily on stromal cells, which produce IL-6, leukocyte adhesion factor, and chemokines in response to OSM [134].
West et al. reported that OSM mRNA expression is elevated in the inflamed mucosa of patients with UC who are resistant to anti-TNF-α antibodies compared to those who responded to anti-TNF-α antibodies [134]. Mucosal healing was achieved in 69–85% of patients with low OSM gene expression but only in 10–15% of patients with high expression. In addition, they reported that tissues with high OSM mRNA and OSMR mRNA expression showed enrichment in genes related to leukocyte chemotaxis, extracellular matrix organization, and mesenchymal development. Smillie et al. conducted single-cell RNA sequencing of colonic mucosa from patients with UC and healthy controls [141], revealing an increase in inflammation-associated fibroblasts (IAFs), a unique subset of fibroblasts in inflammatory tissues of patients with UC who are resistant to anti-TNF-α antibodies. The most enriched gene in IAF was OSMR, suggesting that downstream signaling activation of fibroblasts by OSM contributes to anti-TNF-α antibody resistance. Friedrich et al. also reported an increase in chemokine-high expressing fibroblasts in tissues from patients with IBD who were resistant to treatment with anti-TNF-α antibodies [142]. These results suggest that not only immune cells but also stromal cells, such as fibroblasts, may be involved in the therapeutic effect of anti-TNF-α antibodies [135].
The JAK-STAT pathway, activated by OSM, contributed to intestinal inflammation in a mouse model of DSS-induced colitis [143]. Clinical study in patients with IBD demonstrated a significant reduction in serum OSM levels in patients who respond to treatment with JAK inhibitors [144].

4.1.2. IFN Signature

Type I IFNs, secreted by various cells, including DCs and fibroblasts, play an anti-inflammatory role by promoting epithelial repair and Treg cell differentiation. However, they also play a pro-inflammatory role by inducing IFN-stimulated gene, which has potent antimicrobial activity, and by promoting the production of inflammatory cytokines [145,146]. Type II IFNs, secreted by various cells including Th1 cells, ILC1, B cells, macrophages, and epithelial cells, provide broad protection against intracellular microbes [147]. Excess type II IFN exacerbates inflammation by activating macrophages, augmenting antigen processing, and inducing epithelial cell death [148].
IFN mRNA expressions in the blood and mucosa of patients with UC are heterogeneous [136,149]. High type I and type II IFN signatures in blood and colonic mucosa are associated with resistance to treatment with anti-TNF-α antibodies [136,150]. Mavragani et al. calculated type I IFN scores using interferon-induced protein with tetratricopeptide repeat 1 (IFIT-1) and interferon-induced protein 44 (IFI-44) mRNA expression. Similarly, type II IFN scores were calculated using guanylate-binding protein 1 (GBP-1) and CXCL9 mRNA expression [136]. They reported that the group with lower baseline values for both type I and type II IFN scores had a better response to treatment with anti-TNF-α antibodies. Combining the type I and type II IFN scores predicted the response to treatment with anti-TNF-α antibodies with an area under the curve (AUC) of 0.98, sensitivity of 1.0, specificity of 0.88, positive predictive value (PPV) of 0.88, and negative predictive value (NPV) of 1.0.
Cell death signaling by TNF-α is mainly dependent on RIPK1/3 signaling [105,106], while IFN-induced cell death depends on JAK signaling [151]. It has also been suggested that the combination of TNF-α and IFN promotes cell death in colon epithelial cell lines through a TNFR1-independent synergistic effect [151,152]. It has been suggested that these differences in the mechanism of cell death may contribute to resistance to anti-TNF-α antibodies [152]. Flood et al. demonstrated that cell death induced by the cooperative effects of IFN-β and TNF-α is inhibited by JAK inhibitors using a colon organoid model [151]. Woznicki et al. showed that JAK inhibitors also inhibit IFN-γ-induced cell death using a colon cancer cell line [152].

4.1.3. TREM-1

The triggering receptor expressed on myeloid cells 1 (TREM-1) belongs to the PRR family. TREM-1 is present on the surface of immune and non-immune cells, playing an important role in the host immune system [153]. TREM-1 recognizes PAMPs and DAMPs, collaborating with TLRs to amplify inflammatory innate immune responses [154].
Verstockt et al. reported that IBD patients with lower levels of TREM-1 mRNA expression in their blood and mucosa were more likely to respond to treatment with anti-TNF antibodies and achieve endoscopic remission [137]. Czarnewski et al., using human homologs of differentially expressed genes identified by mRNA sequencing of the colon from mouse models of DSS-induced colitis, stratified patients with UC into two major transcriptome profiles (UC1 and UC2) [155]. Approximately 70% of UC2 patients responded to infliximab, compared to less than 10% of UC1 patients. Furthermore, approximately 60% of UC2 patients responded to vedolizumab (VED), while only 13% of UC1 patients did. TREM-1 was identified as the most accurate biomarker for classifying UC1 and UC2.
Prins et al. reported that the differentiation of monocytes into regulatory macrophages induced by anti-TNF-α antibodies was inhibited in monocytes expressing high levels of TREM-1 [156]. In addition, they reported a decrease in the expression of genes related to autophagy in these monocytes. Kökten et al. reported that inhibiting TREM-1 decreased the expression of mammalian target of rapamycin (mTOR) and increased the expression of autophagy related protein (ATG)1/UNC-51-like kinase (ULK)1 and ATG13 (which are involved in the initiation of autophagosome formation) and ATG5, ATG16L1, and microtubule-associated protein 1 light chain 3 (MAP1LC3)-I/II (which are involved in membrane elongation and expansion of formed autophagosomes) [157]. Because autophagy is indirectly involved in the differentiation and maintenance of regulatory macrophages [158,159], an increase in monocytes expressing high levels of TREM-1 may contribute to resistance to treatment with anti-TNF-α antibodies [160]. In addition, high TREM-1 expression may impair autophagy in intestinal epithelial cells, which promotes endoplasmic reticulum stress and exacerbates inflammation resistant to anti-TNF-α antibodies [157,161,162,163,164]. In a mouse model of DSS-induced colitis, pharmacological or genetic inhibition of TREM-1 has been shown to enhance macroautophagy and chaperone-mediated autophagy via mTOR dysregulation, thereby reducing endoplasmic reticulum stress and suppressing colitis [157].

4.1.4. IL-23

Regarding a resistance factor to T cell apoptosis induced by anti-TNF-α antibodies, Schmitt et al. confirmed that IL-23 is upregulated in the mucosa of patients with IBD who do not respond to treatment with anti-TNF-α antibodies [165]. They found that a unique T cell subset, TNFR2+IL-23R+ T cells, was increased in the mucosa of these patients. TNFR2+IL-23R+ T cells were resistant to apoptosis induced by anti-TNF-α antibodies in an IL-23-dependent manner. They also reported that apoptosis-resistant TNFR2+IL23R+ T cells expressed integrin α4β7 and showed increased expression of IFN-γ, T-box protein expressed in T cells (T-bet), IL-17A, and RORγt compared to TNFR2+IL23R cells.

4.1.5. IL-1β

Obraztsov et al. calculated Pearson correlation coefficients to analyze the relationship between 17 serum cytokine levels and the response to treatment with anti-TNF-α antibodies and developed a cytokine score using seven cytokine subsets including IL-1β, TNF-α, IL-12, IL-8, IL-2, IL-5, and IFN-γ [16]. In the study, patients received treatment with 5 mg/kg of anti-TNF-α antibody at weeks 0, 2, and 6, and their response to the treatment was evaluated at week 12. The score predicted clinical remission (Mayo score < 3) after three courses of anti-TNF-α antibody treatment with a sensitivity of 84.2%, specificity of 93.3%, and accuracy of 89.8% (44/49). Among several cytokine expressions, a high level of IL-1β is mostly associated with resistance to anti-TNF-α antibodies. Another report also suggests that high expression of IL-1β is associated with resistance to anti-TNF-α antibodies [16].
Bouwman et al. evaluated the association between activation of signaling pathways in the inflamed mucosa of patients with UC and responsiveness to treatment with anti-TNF-α antibodies [166]. They reported that the nuclear factor κ-light-chain-enhancer of activated B cells (NFκB), TGF-β, and JAK-STAT3 signaling pathways were activated in the mucosa of patients with UC who resistant to anti-TNF-α antibodies [166]. Candidate cytokines upstream of these signals were IL-1β and IL-17.
Friedrich et al. conducted RNA sequencing on surgically resected specimens from patients with IBD who did not respond to medical therapy [142]. They used weighted gene correlation network analysis to cluster co-expressed genes and identified 38 modules. Among these modules, two modules were associated with resistance to treatment with anti-TNF-α antibodies, and strongly linked to neutrophils and stromal cells. In addition, these modules were enriched with genes associated with inflammasomes. Furthermore, they identified inflammatory fibroblasts with high levels of chemokines that promote neutrophil recruitment using single-cell RNA sequencing. They reported that anti-TNF-α antibodies failed to suppress chemokine expression in these inflammatory fibroblasts, whereas IL-1 receptor inhibitors did. Thus, the activation of fibroblasts by IL-1β could contribute to resistance to treatment with anti-TNF-α antibody.

4.2. Anti-IL-12/23p40 Antibodies and Anti-IL-23p19 Antibodies

IL-22

In two randomized controlled trials, the administration of an anti-IL-23 antibody significantly reduced IL-22 levels in the blood and IL-22 mRNA expression in the mucosa compared to the placebo group [81,167].
Pavlidis et al. assessed the IL-22 enrichment scores of colonic mucosa collected before ustekinumab initiation in patients with UC and observed substantial variation among patients [138]. Patients with low IL-22 enrichment scores had nearly twice the remission rates compared to all non-stratified patients, including clinical remission (25% vs. 13%) and mucosal healing (26% vs. 16%). Conversely, patients with high IL-22 enrichment scores showed outcomes similar to those in the placebo groups. The study confirmed that in patients resistant to anti-IL-12/23p40 antibodies, IL-22 expression depends on IL-1β rather than IL-23. Additionally, it demonstrated that IL-22 might induce the production of CXCL1 and CXCL5 from IECs via STAT3 activation and recruit C-X-C chemokine receptor 2 (CXCR2)-expressing neutrophils to the intestinal mucosa.

4.3. Anti-Integrin α4β7 Antibodies

4.3.1. α4β7 Integrin

α4β7 integrin expression in immune cells contributes to the therapeutic effect of VED. Rath et al. reported that responders to treatment with VED had a higher mean number of α4β7+ immune cells per high power field in the intestinal lamina propria compared to non-responders (13.4 vs. 5.8, p = 0.0003) [168]. In a prospective clinical trial, responders to VED had lower baseline α4β7 expression of CD3+ and CD4+ T cells in peripheral blood compared to non-responders [169]. Furthermore, in responders, α4β7+ immune cells increased in peripheral blood and decreased in intestinal mucosa after VED administration [170], suggesting that the α4β7-MAdCAM-1 axis is the dominant recruitment pathway in VED responders.
Inflammatory cytokines like TNF-α promote MAdCAM-1 expression [171] and also increase vascular cell adhesion molecule 1 (VCAM-1), E-cadherin and L-selectin [172,173]. Baseline expression of α4β1 and αEβ7 in peripheral blood was not associated with the response to treatment with VED [169], but CD4+ T cells expressing αEβ7 and α4β1 in the peripheral blood increased after VED administration in non-responders [170,174]. Thus, redundancy in compensatory homing pathways may contribute to VED resistance. In a mouse model of enteritis unresponsive to α4β7 blockade, simultaneous inhibition of compensatory pathways such as L-selectin [175] and αEβ7 [176] has been shown to improve enteritis.

4.3.2. IL-6 and IL-8

Bertani et al. reported that higher baseline serum IL-6 and IL-8 concentrations and decreased IL-6 and IL-8 after VED administration predict mucosal healing at week 54 [139]. However, the mechanism by which IL-6 and IL-8 predict the efficacy of treatment with VED is unknown. Several studies, including the above study, have shown that patients resistant to treatment with VED have higher disease activity, including longer disease duration, higher endoscopy scores, higher c-reactive protein (CRP), and lower albumin [140,177,178,179]. Elevated IL-6 and IL-8 levels may reflect disease severity. VED prevents the recruitment of new effector immune cells but has a limited effect on resident effector immune cells [180], making VED less effective in patients with severe disease activity. Combination therapy with other advanced therapies may overcome this limitation [181].

4.4. JAK Inhibitor

There are few reports on cytokine biomarkers that can predict the therapeutic effect of JAK inhibitors. While cytokines, such as serum IL-4 and IL-10, have been reported as potential biomarkers, their utility and underlying mechanisms have not been fully investigated [182]. Roblin et al. examined serum cytokine concentrations in patients with Crohn’s disease treated with filgotinib [144]. Although their study did not examine in detail whether baseline serum cytokine concentrations could predict treatment response, they found that significant reductions in IL-6 and OSM after treatment with filgotinib were significantly correlated with an endoscopic response (>50% reduction in simple endoscopic score for Crohn’s disease at week 10). As noted in the previous section, JAK inhibitors have been shown to alleviate inflammation induced by OSM and IFN in experimental mouse colitis models [143,151], suggesting the potential for JAK inhibitors to be effective in patients with UC exhibiting IL-6, OSM, and IFN-dominant cytokine profiles.

5. Prospects

Treatment options for UC have increased rapidly in recent years, leading to increasingly complex management algorithms, including which drugs to use, in what order, when to start, when to change doses, and when to discontinue. One of the obstacles to promoting personalized medicine for UC is the absence of a definitive biomarker for guiding treatment selection. The clinical application of cytokine profiles as a biomarker may be a feasible approach, because the underlying inflammatory cytokine profiles vary among patients with UC [183]. However, there is currently no evidence supporting the use of cytokine profiles to select optimal therapeutic agents. The complexity of differences in cytokine profiles between individuals, the variety of cytokine roles, and differences in racial and genetic backgrounds may make the use of cytokine profiles as biomarkers difficult [184].
This review summarizes the mechanisms underlying the therapeutic efficacy of advanced therapies for UC. Clinical trials in large cohorts are needed to determine the utility of cytokine profiles in selecting optimal therapeutic agents.

Author Contributions

Conceptualization, H.K., Y.Y., T.H., K.A., Y.H., T.K., H.I. and H.N.; methodology, H.K., Y.Y. and H.N.; formal analysis, H.K., Y.Y. and H.N.; data curation, H.K., Y.Y. and H.N.; writing—original draft preparation, H.K.; writing—review and editing, Y.Y. and H.N.; supervision, H.N. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Health and Labour Sciences Research Grants for research on intractable diseases from the Ministry of Health, Labour and Welfare (MHLW) of Japan (Investigation and Research for Intractable Inflammatory Bowel Disease) (Grant Number 20316729).

Data Availability Statement

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

Conflicts of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Nakase, H.; Uchino, M.; Shinzaki, S.; Matsuura, M.; Matsuoka, K.; Kobayashi, T.; Saruta, M.; Hirai, F.; Hata, K.; Hiraoka, S.; et al. Evidence-based clinical practice guidelines for inflammatory bowel disease 2020. J. Gastroenterol. 2021, 56, 489–526. [Google Scholar] [CrossRef]
  2. Jostins, L.; Ripke, S.; Weersma, R.K.; Duerr, R.H.; McGovern, D.P.; Hui, K.Y.; Lee, J.C.; Schumm, L.P.; Sharma, Y.; Anderson, C.A.; et al. Host-microbe interactions have shaped the genetic architecture of inflammatory bowel disease. Nature 2012, 491, 119–124. [Google Scholar] [CrossRef] [PubMed]
  3. Liu, J.Z.; van Sommeren, S.; Huang, H.; Ng, S.C.; Alberts, R.; Takahashi, A.; Ripke, S.; Lee, J.C.; Jostins, L.; Shah, T.; et al. Association analyses identify 38 susceptibility loci for inflammatory bowel disease and highlight shared genetic risk across populations. Nat. Genet. 2015, 47, 979–986. [Google Scholar] [CrossRef] [PubMed]
  4. de Lange, K.M.; Moutsianas, L.; Lee, J.C.; Lamb, C.A.; Luo, Y.; Kennedy, N.A.; Jostins, L.; Rice, D.L.; Gutierrez-Achury, J.; Ji, S.G.; et al. Genome-wide association study implicates immune activation of multiple integrin genes in inflammatory bowel disease. Nat. Genet. 2017, 49, 256–261. [Google Scholar] [CrossRef] [PubMed]
  5. Nakase, H. Understanding the efficacy of individual Janus kinase inhibitors in the treatment of ulcerative colitis for future positioning in inflammatory bowel disease treatment. Immunol. Med. 2023, 46, 121–130. [Google Scholar] [CrossRef] [PubMed]
  6. Nakase, H.; Sato, N.; Mizuno, N.; Ikawa, Y. The influence of cytokines on the complex pathology of ulcerative colitis. Autoimmun. Rev. 2022, 21, 103017. [Google Scholar] [CrossRef]
  7. Baumgart, D.C.; Le Berre, C. Newer Biologic and Small-Molecule Therapies for Inflammatory Bowel Disease. N. Engl. J. Med. 2021, 385, 1302–1315. [Google Scholar] [CrossRef]
  8. Luo, H.; Cao, G.; Luo, C.; Tan, D.; Vong, C.T.; Xu, Y.; Wang, S.; Lu, H.; Wang, Y.; Jing, W. Emerging pharmacotherapy for inflammatory bowel diseases. Pharmacol. Res. 2022, 178, 106146. [Google Scholar] [CrossRef]
  9. Chang, J.T. Pathophysiology of Inflammatory Bowel Diseases. N. Engl. J. Med. 2020, 383, 2652–2664. [Google Scholar] [CrossRef]
  10. Schmidt, C.; Grunert, P.C.; Stallmach, A. An Update for Pharmacologists on New Treatment Options for Inflammatory Bowel Disease: The Clinicians’ Perspective. Front. Pharmacol. 2021, 12, 655054. [Google Scholar] [CrossRef]
  11. Geem, D.; Kugathasan, S. It Takes Two to Make It Right: Dual Biologic and Small Molecule Therapy for Treatment-Refractory Pediatric Inflammatory Bowel Disease. Inflamm. Bowel Dis. 2021, 27, 1361–1362. [Google Scholar] [CrossRef]
  12. Guagnozzi, D.; Caprilli, R. Natalizumab in the treatment of Crohn’s disease. Biologics 2008, 2, 275–284. [Google Scholar]
  13. Nakase, H. Treatment of inflammatory bowel disease from the immunological perspective. Immunol. Med. 2020, 43, 79–86. [Google Scholar] [CrossRef]
  14. Roda, G.; Jharap, B.; Neeraj, N.; Colombel, J.F. Loss of Response to Anti-TNFs: Definition, Epidemiology, and Management. Clin. Transl. Gastroenterol 2016, 7, e135. [Google Scholar] [CrossRef]
  15. Sands, B.E.; Chen, J.; Feagan, B.G.; Penney, M.; Rees, W.A.; Danese, S.; Higgins, P.D.R.; Newbold, P.; Faggioni, R.; Patra, K.; et al. Efficacy and Safety of MEDI2070, an Antibody Against Interleukin 23, in Patients with Moderate to Severe Crohn’s Disease: A Phase 2a Study. Gastroenterology 2017, 153, 77–86.e6. [Google Scholar] [CrossRef] [PubMed]
  16. Obraztsov, I.V.; Shirokikh, K.E.; Obraztsova, O.I.; Shapina, M.V.; Wang, M.-H.; Khalif, I.L. Multiple Cytokine Profiling: A New Model to Predict Response to Tumor Necrosis Factor Antagonists in Ulcerative Colitis Patients. Inflamm. Bowel Dis. 2019, 25, 524–531. [Google Scholar] [CrossRef] [PubMed]
  17. Dahlén, R.; Magnusson, M.K.; Bajor, A.; Lasson, A.; Ung, K.-A.; Strid, H.; Öhman, L. Global mucosal and serum cytokine profile in patients with ulcerative colitis undergoing anti-TNF therapy. Scand. J. Gastroenterol. 2015, 50, 1118–1126. [Google Scholar] [CrossRef] [PubMed]
  18. Verstockt, B.; Bressler, B.; Martinez-Lozano, H.; McGovern, D.; Silverberg, M.S. Time to Revisit Disease Classification in Inflammatory Bowel Disease: Is the Current Classification of Inflammatory Bowel Disease Good Enough for Optimal Clinical Management? Gastroenterology 2022, 162, 1370–1382. [Google Scholar] [CrossRef]
  19. Ananthakrishnan, A.N.; Kaplan, G.G.; Ng, S.C. Changing Global Epidemiology of Inflammatory Bowel Diseases: Sustaining Health Care Delivery Into the 21st Century. Clin. Gastroenterol. Hepatol. 2020, 18, 1252–1260. [Google Scholar] [CrossRef]
  20. Plichta, D.R.; Graham, D.B.; Subramanian, S.; Xavier, R.J. Therapeutic Opportunities in Inflammatory Bowel Disease: Mechanistic Dissection of Host-Microbiome Relationships. Cell 2019, 178, 1041–1056. [Google Scholar] [CrossRef]
  21. Yang, Y.; Fu, K.Z.; Pan, G. Role of Oncostatin M in the prognosis of inflammatory bowel disease: A meta-analysis. World J. Gastrointest. Surg. 2024, 16, 228–238. [Google Scholar] [CrossRef] [PubMed]
  22. Guo, A.; Ross, C.; Chande, N.; Gregor, J.; Ponich, T.; Khanna, R.; Sey, M.; Beaton, M.; Yan, B.; Kim, R.B.; et al. High oncostatin M predicts lack of clinical remission for patients with inflammatory bowel disease on tumor necrosis factor α antagonists. Sci. Rep. 2022, 12, 1185. [Google Scholar] [CrossRef] [PubMed]
  23. Pickert, G.; Neufert, C.; Leppkes, M.; Zheng, Y.; Wittkopf, N.; Warntjen, M.; Lehr, H.A.; Hirth, S.; Weigmann, B.; Wirtz, S.; et al. STAT3 links IL-22 signaling in intestinal epithelial cells to mucosal wound healing. J. Exp. Med. 2009, 206, 1465–1472. [Google Scholar] [CrossRef] [PubMed]
  24. Lee, J.S.; Tato, C.M.; Joyce-Shaikh, B.; Gulan, F.; Cayatte, C.; Chen, Y.; Blumenschein, W.M.; Judo, M.; Ayanoglu, G.; McClanahan, T.K.; et al. Interleukin-23-Independent IL-17 Production Regulates Intestinal Epithelial Permeability. Immunity 2015, 43, 727–738. [Google Scholar] [CrossRef] [PubMed]
  25. Maxwell, J.R.; Zhang, Y.; Brown, W.A.; Smith, C.L.; Byrne, F.R.; Fiorino, M.; Stevens, E.; Bigler, J.; Davis, J.A.; Rottman, J.B.; et al. Differential Roles for Interleukin-23 and Interleukin-17 in Intestinal Immunoregulation. Immunity 2015, 43, 739–750. [Google Scholar] [CrossRef]
  26. Shih, V.F.-S.; Cox, J.; Kljavin, N.M.; Dengler, H.S.; Reichelt, M.; Kumar, P.; Rangell, L.; Kolls, J.K.; Diehl, L.; Ouyang, W.; et al. Homeostatic IL-23 receptor signaling limits Th17 response through IL-22-mediated containment of commensal microbiota. Proc. Natl. Acad. Sci. USA 2014, 111, 13942–13947. [Google Scholar] [CrossRef]
  27. Ivanov, I.I.; Atarashi, K.; Manel, N.; Brodie, E.L.; Shima, T.; Karaoz, U.; Wei, D.; Goldfarb, K.C.; Santee, C.A.; Lynch, S.V.; et al. Induction of Intestinal Th17 Cells by Segmented Filamentous Bacteria. Cell 2009, 139, 485–498. [Google Scholar] [CrossRef]
  28. Grizotte-Lake, M.; Zhong, G.; Duncan, K.; Kirkwood, J.; Iyer, N.; Smolenski, I.; Isoherranen, N.; Vaishnava, S. Commensals Suppress Intestinal Epithelial Cell Retinoic Acid Synthesis to Regulate Interleukin-22 Activity and Prevent Microbial Dysbiosis. Immunity 2018, 49, 1103–1115.e6. [Google Scholar] [CrossRef]
  29. Schiering, C.; Wincent, E.; Metidji, A.; Iseppon, A.; Li, Y.; Potocnik, A.J.; Omenetti, S.; Henderson, C.J.; Wolf, C.R.; Nebert, D.W.; et al. Feedback control of AHR signalling regulates intestinal immunity. Nature 2017, 542, 242–245. [Google Scholar] [CrossRef]
  30. Kole, A.; He, J.; Rivollier, A.; Silveira, D.D.; Kitamura, K.; Maloy, K.J.; Kelsall, B.L. Type I IFNs regulate effector and regulatory T cell accumulation and anti-inflammatory cytokine production during T cell-mediated colitis. J. Immunol. 2013, 191, 2771–2779. [Google Scholar] [CrossRef]
  31. Cosín-Roger, J.; Ortiz-Masiá, D.; Calatayud, S.; Hernández, C.; Esplugues, J.V.; Barrachina, M.D. The activation of Wnt signaling by a STAT6-dependent macrophage phenotype promotes mucosal repair in murine IBD. Mucosal Immunol. 2016, 9, 986–998. [Google Scholar] [CrossRef] [PubMed]
  32. Stein, M.; Keshav, S.; Harris, N.; Gordon, S. Interleukin 4 Potently Enhances Murine Macrophage Mannose Receptor Activity: A Marker of Alternative Immunologic Macrophage Activation. J. Exp. Med. 1992, 176, 287–292. [Google Scholar] [CrossRef] [PubMed]
  33. Krawiec, P.; Pawłowska-Kamieniak, A.; Pac-Kożuchowska, E. Interleukin 10 and interleukin 10 receptor in paediatric inflammatory bowel disease: From bench to bedside lesson. J. Inflamm. 2021, 18, 13. [Google Scholar] [CrossRef]
  34. Saraiva, M.; O’Garra, A. The regulation of IL-10 production by immune cells. Nat. Rev. Immunol. 2010, 10, 170–181. [Google Scholar] [CrossRef] [PubMed]
  35. Gabryšová, L.; Alvarez-Martinez, M.; Luisier, R.; Cox, L.S.; Sodenkamp, J.; Hosking, C.; Pérez-Mazliah, D.; Whicher, C.; Kannan, Y.; Potempa, K.; et al. c-Maf controls immune responses by regulating disease-specific gene networks and repressing IL-2 in CD4+ T cells. Nat. Immunol. 2018, 19, 497–507. [Google Scholar] [CrossRef] [PubMed]
  36. Chinen, T.; Kannan, A.K.; Levine, A.G.; Fan, X.; Klein, U.; Zheng, Y.; Gasteiger, G.; Feng, Y.; Fontenot, J.D.; Rudensky, A.Y. An essential role for the IL-2 receptor in T reg cell function. Nat. Immunol. 2016, 17, 1322–1333. [Google Scholar] [CrossRef] [PubMed]
  37. Schiering, C.; Krausgruber, T.; Chomka, A.; Fröhlich, A.; Adelmann, K.; Wohlfert, E.A.; Pott, J.; Griseri, T.; Bollrath, J.; Hegazy, A.N.; et al. The alarmin IL-33 promotes regulatory T-cell function in the intestine. Nature 2014, 513, 564–568. [Google Scholar] [CrossRef] [PubMed]
  38. Macpherson, A.J.; Yilmaz, B.; Limenitakis, J.P.; Ganal-Vonarburg, S.C. IgA Function in Relation to the Intestinal Microbiota. Annu. Rev. Immunol. 2018, 36, 359–381. [Google Scholar] [CrossRef]
  39. Cerutti, A. The regulation of IgA class switching. Nat. Rev. Immunol. 2008, 8, 421–434. [Google Scholar] [CrossRef]
  40. Wang, L.; Ray, A.; Jiang, X.; Wang, J.; Basu, S.; Liu, X.; Qian, T.; He, R.; Dittel, B.N.; Chu, Y. T regulatory cells and B cells cooperate to form a regulatory loop that maintains gut homeostasis and suppresses dextran sulfate sodium-induced colitis. Mucosal Immunol. 2015, 8, 1297–1312. [Google Scholar] [CrossRef]
  41. Tan, Z.; Liu, C.; He, P.; Wu, Y.; Li, J.; Zhang, J.; Dong, W. Based on weighted gene co-expression network analysis reveals the hub immune infiltration-related genes associated with ulcerative colitis. J. Inflamm. Res. 2024, 17, 357–370. [Google Scholar] [CrossRef] [PubMed]
  42. Mateer, S.W.; Mathe, A.; Bruce, J.; Liu, G.; Maltby, S.; Fricker, M.; Goggins, B.J.; Tay, H.L.; Marks, E.; Burns, G.; et al. IL-6 Drives Neutrophil-Mediated Pulmonary Inflammation Associated with Bacteremia in Murine Models of Colitis. Am. J. Pathol. 2018, 188, 1625–1639. [Google Scholar] [CrossRef] [PubMed]
  43. Okabe, S.; Ishikawa, T.; Tanami, S.; Kuwabara, H.; Hukahara, T.; Udagawa, M.; Ootsukasa, S.; Arai, T.; Maruyama, S.; Murase, N.; et al. Abdominosacral repair for a rectovaginal fistula with anastomotic stenosis after low anterior resection: Report of a case. J. Med. Dent. Sci. 2001, 48, 41–44. [Google Scholar] [PubMed]
  44. Salem, M.; El Azreq, M.A.; Pelletier, J.; Robaye, B.; Aoudjit, F.; Sévigny, J. Exacerbated intestinal inflammation in P2Y6 deficient mice is associated with Th17 activation. Biochim. Biophys. Acta-Mol. Basis Dis. 2019, 1865, 2595–2605. [Google Scholar] [CrossRef]
  45. Hansberry, D.R.; Shah, K.; Agarwal, P.; Agarwal, N. Fecal Myeloperoxidase as a Biomarker for Inflammatory Bowel Disease. Cureus 2017, 9, e1004. [Google Scholar] [CrossRef] [PubMed]
  46. Drury, B.; Hardisty, G.; Gray, R.D.; Ho, G.-T. Neutrophil Extracellular Traps in Inflammatory Bowel Disease: Pathogenic Mechanisms and Clinical Translation. Cell. Mol. Gastroenterol. Hepatol. 2021, 12, 321–333. [Google Scholar] [CrossRef] [PubMed]
  47. Yui, S.; Nakatani, Y.; Mikami, M. Calprotectin (S100A8/S100A9), an inflammatory protein complex from neutrophils with a broad apoptosis-inducing activity. Biol. Pharm. Bull. 2003, 26, 753–760. [Google Scholar] [CrossRef] [PubMed]
  48. Cunliffe, R.N.; Kamal, M.; Rose, F.R.; James, P.D.; Mahida, Y.R. Expression of antimicrobial neutrophil defensins in epithelial cells of active inflammatory bowel disease mucosa. J. Clin. Pathol. 2002, 55, 298–304. [Google Scholar] [CrossRef]
  49. Karmakar, M.; Minns, M.; Greenberg, E.N.; Diaz-Aponte, J.; Pestonjamasp, K.; Johnson, J.L.; Rathkey, J.K.; Abbott, D.W.; Wang, K.; Shao, F.; et al. N-GSDMD trafficking to neutrophil organelles facilitates IL-1β release independently of plasma membrane pores and pyroptosis. Nat. Commun. 2020, 11, 2212. [Google Scholar] [CrossRef]
  50. Bain, C.C.; Scott, C.L.; Uronen-Hansson, H.; Gudjonsson, S.; Jansson, O.; Grip, O.; Guilliams, M.; Malissen, B.; Agace, W.W.; Mowat, A.M. Resident and pro-inflammatory macrophages in the colon represent alternative context-dependent fates of the same Ly6C hi monocyte precursors. Mucosal Immunol. 2013, 6, 498–510. [Google Scholar] [CrossRef]
  51. Bain, C.C.; Bravo-Blas, A.; Scott, C.L.; Perdiguero, E.G.; Geissmann, F.; Henri, S.; Malissen, B.; Osborne, L.C.; Artis, D.; Mowat, A.M. Constant replenishment from circulating monocytes maintains the macrophage pool in the intestine of adult mice. Nat. Immunol. 2014, 15, 929–937. [Google Scholar] [CrossRef]
  52. Serbina, N.V.; Pamer, E.G. Monocyte emigration from bone marrow during bacterial infection requires signals mediated by chemokine receptor CCR2. Nat. Immunol. 2006, 7, 311–317. [Google Scholar] [CrossRef] [PubMed]
  53. Zigmond, E.; Varol, C.; Farache, J.; Elmaliah, E.; Satpathy, A.T.; Friedlander, G.; Mack, M.; Shpigel, N.; Boneca, I.G.; Murphy, K.M.; et al. Ly6C hi Monocytes in the Inflamed Colon Give Rise to Proinflammatory Effector Cells and Migratory Antigen-Presenting Cells. Immunity 2012, 37, 1076–1090. [Google Scholar] [CrossRef] [PubMed]
  54. Redhu, N.S.; Bakthavatchalu, V.; Conaway, E.A.; Shouval, D.S.; Tsou, A.; Goettel, J.A.; Biswas, A.; Wang, C.; Field, M.; Muller, W.; et al. Macrophage dysfunction initiates colitis during weaning of infant mice lacking the interleukin-10 receptor. Elife 2017, 6, e27652. [Google Scholar] [CrossRef] [PubMed]
  55. Grimm, M.C.; Pullman, W.E.; Bennett, G.M.; Sullivan, P.J.; Pavli, P.; Doe, W.F. Direct evidence of monocyte recruitment to inflammatory bowel disease mucosa. J. Gastroenterol. Hepatol. 1995, 10, 387–395. [Google Scholar] [CrossRef] [PubMed]
  56. Yao, H.; Tang, G. Macrophages in intestinal fibrosis and regression. Cell. Immunol. 2022, 381, 104614. [Google Scholar] [CrossRef] [PubMed]
  57. Chieppa, M.; Rescigno, M.; Huang, A.Y.C.; Germain, R.N. Dynamic imaging of dendritic cell extension into the small bowel lumen in response to epithelial cell TLR engagement. J. Exp. Med. 2006, 203, 2841–2852. [Google Scholar] [CrossRef] [PubMed]
  58. Hart, A.L.; Al-Hassi, H.O.; Rigby, R.J.; Bell, S.J.; Emmanuel, A.V.; Knight, S.C.; Kamm, M.A.; Stagg, A.J. Characteristics of intestinal dendritic cells in inflammatory bowel diseases. Gastroenterology 2005, 129, 50–65. [Google Scholar] [CrossRef] [PubMed]
  59. Al-Hassi, H.O.; Mann, E.R.; Sanchez, B.; English, N.R.; Peake, S.T.C.; Landy, J.; Man, R.; Urdaci, M.; Hart, A.L.; Fernandez-Salazar, L.; et al. Altered human gut dendritic cell properties in ulcerative colitis are reversed by Lactobacillus plantarum extracellular encrypted peptide STp. Mol. Nutr. Food Res. 2014, 58, 1132–1143. [Google Scholar] [CrossRef]
  60. Magnusson, M.K.; Brynjólfsson, S.F.; Dige, A.; Uronen-Hansson, H.; Börjesson, L.G.; Bengtsson, J.L.; Gudjonsson, S.; Öhman, L.; Agnholt, J.; Sjövall, H.; et al. Macrophage and dendritic cell subsets in IBD: ALDH+ cells are reduced in colon tissue of patients with ulcerative colitis regardless of inflammation. Mucosal. Immunol. 2016, 9, 171–182. [Google Scholar] [CrossRef]
  61. Hue, S.; Ahern, P.; Buonocore, S.; Kullberg, M.C.; Cua, D.J.; McKenzie, B.S.; Powrie, F.; Maloy, K.J. Interleukin-23 drives innate and T cell-mediated intestinal inflammation. J. Exp. Med. 2006, 203, 2473–2483. [Google Scholar] [CrossRef]
  62. Xavier, R.J.; Podolsky, D.K. Unravelling the pathogenesis of inflammatory bowel disease. Nature 2007, 448, 427–434. [Google Scholar] [CrossRef] [PubMed]
  63. Ignacio, A.; Breda, C.N.S.; Camara, N.O.S. Innate lymphoid cells in tissue homeostasis and diseases. World J. Hepatol. 2017, 9, 979–989. [Google Scholar] [CrossRef] [PubMed]
  64. Spits, H.; Artis, D.; Colonna, M.; Diefenbach, A.; Di Santo, J.P.; Eberl, G.; Koyasu, S.; Locksley, R.M.; McKenzie, A.N.; Mebius, R.E.; et al. Innate lymphoid cells-a proposal for uniform nomenclature. Nat. Rev. Immunol. 2013, 13, 145–149. [Google Scholar] [CrossRef]
  65. Buonocore, S.; Ahern, P.P.; Uhlig, H.H.; Ivanov, I.I.; Littman, D.R.; Maloy, K.J.; Powrie, F. Innate lymphoid cells drive interleukin-23-dependent innate intestinal pathology. Nature 2010, 464, 1371–1375. [Google Scholar] [CrossRef]
  66. Bernink, J.H.; Peters, C.P.; Munneke, M.; te Velde, A.A.; Meijer, S.L.; Weijer, K.; Hreggvidsdottir, H.S.; Heinsbroek, S.E.; Legrand, N.; Buskens, C.J.; et al. Human type 1 innate lymphoid cells accumulate in inflamed mucosal tissues. Nat. Immunol. 2013, 14, 221–229. [Google Scholar] [CrossRef] [PubMed]
  67. Fuchs, A.; Vermi, W.; Lee, J.S.; Lonardi, S.; Gilfillan, S.; Newberry, R.D.; Cella, M.; Colonna, M. Intraepithelial type 1 innate lymphoid cells are a unique subset of IL-12- and IL-15-responsive IFN-γ-producing cells. Immunity 2013, 38, 769–781. [Google Scholar] [CrossRef]
  68. Forkel, M.; van Tol, S.; Höög, C.; Michaëlsson, J.; Almer, S.; Mjösberg, J. Distinct Alterations in the Composition of Mucosal Innate Lymphoid Cells in Newly Diagnosed and Established Crohn’s Disease and Ulcerative Colitis. J. Crohns Colitis 2019, 13, 67–78. [Google Scholar] [CrossRef]
  69. Gwela, A.; Siddhanathi, P.; Chapman, R.W.; Travis, S.; Powrie, F.; Arancibia-Cárcamo, C.V.; Geremia, A. Th1 and Innate Lymphoid Cells Accumulate in Primary Sclerosing Cholangitis-associated Inflammatory Bowel Disease. J. Crohns Colitis 2017, 11, 1124–1134. [Google Scholar] [CrossRef]
  70. Powell, N.; Walker, A.W.; Stolarczyk, E.; Canavan, J.B.; Gökmen, M.R.; Marks, E.; Jackson, I.; Hashim, A.; Curtis, M.A.; Jenner, R.G.; et al. The transcription factor T-bet regulates intestinal inflammation mediated by interleukin-7 receptor+ innate lymphoid cells. Immunity 2012, 37, 674–684. [Google Scholar] [CrossRef]
  71. Sunaga, S.; Tsunoda, J.; Teratani, T.; Mikami, Y.; Kanai, T. Heterogeneity of ILC2s in the Intestine; Homeostasis and Pathology. Front. Immunol. 2022, 13, 867351. [Google Scholar] [CrossRef] [PubMed]
  72. Gomez-Bris, R.; Saez, A.; Herrero-Fernandez, B.; Rius, C.; Sanchez-Martinez, H.; Gonzalez-Granado, J.M. CD4 T-Cell Subsets and the Pathophysiology of Inflammatory Bowel Disease. Int. J. Mol. Sci. 2023, 24, 2696. [Google Scholar] [CrossRef] [PubMed]
  73. Tindemans, I.; Joosse, M.E.; Samsom, J.N. Dissecting the Heterogeneity in T-Cell Mediated Inflammation in IBD. Cells 2020, 9, 110. [Google Scholar] [CrossRef] [PubMed]
  74. Friedrich, M.; Pohin, M.; Powrie, F. Cytokine Networks in the Pathophysiology of Inflammatory Bowel Disease. Immunity 2019, 50, 992–1006. [Google Scholar] [CrossRef]
  75. Cao, H.; Diao, J.; Liu, H.; Liu, S.; Liu, J.; Yuan, J.; Lin, J. The Pathogenicity and Synergistic Action of Th1 and Th17 Cells in Inflammatory Bowel Diseases. Inflamm. Bowel Dis. 2023, 29, 818–829. [Google Scholar] [CrossRef] [PubMed]
  76. Siddiqui, K.R.R.; Laffont, S.; Powrie, F. E-cadherin marks a subset of inflammatory dendritic cells that promote T cell-mediated colitis. Immunity 2010, 32, 557–567. [Google Scholar] [CrossRef] [PubMed]
  77. Arnold, I.C.; Mathisen, S.; Schulthess, J.; Danne, C.; Hegazy, A.N.; Powrie, F. CD11c+ monocyte/macrophages promote chronic Helicobacter hepaticus-induced intestinal inflammation through the production of IL-23. Mucosal Immunol. 2016, 9, 352–363. [Google Scholar] [CrossRef] [PubMed]
  78. Pirhonen, J.; Matikainen, S.; Julkunen, I. Regulation of virus-induced IL-12 and IL-23 expression in human macrophages. J. Immunol. 2002, 169, 5673–5678. [Google Scholar] [CrossRef]
  79. Verreck, F.A.W.; de Boer, T.; Langenberg, D.M.L.; Hoeve, M.A.; Kramer, M.; Vaisberg, E.; Kastelein, R.; Kolk, A.; de Waal-Malefyt, R.; Ottenhoff, T.H. Human IL-23-producing type 1 macrophages promote but IL-10-producing type 2 macrophages subvert immunity to (myco)bacteria. Proc. Natl. Acad. Sci. USA 2004, 101, 4560–4565. [Google Scholar] [CrossRef]
  80. Teng, M.W.L.; Bowman, E.P.; McElwee, J.J.; Smyth, M.J.; Casanova, J.-L.; Cooper, A.M.; Cua, D.J. IL-12 and IL-23 cytokines: From discovery to targeted therapies for immune-mediated inflammatory diseases. Nat. Med. 2015, 21, 719–729. [Google Scholar] [CrossRef]
  81. Sandborn, W.J.; Ferrante, M.; Bhandari, B.R.; Berliba, E.; Feagan, B.G.; Hibi, T.; D’Haens, G.R.; Tuttle, J.L.; Krueger, K.; Friedrich, S.; et al. Efficacy and Safety of Continued Treatment with Mirikizumab in a Phase 2 Trial of Patients with Ulcerative Colitis. Gastroenterology 2020, 158, 537–549.e10. [Google Scholar] [CrossRef] [PubMed]
  82. Duerr, R.H.; Taylor, K.D.; Brant, S.R.; Rioux, J.D.; Silverberg, M.S.; Daly, M.J.; Steinhart, A.H.; Abraham, C.; Regueiro, M.; Griffiths, A.; et al. A genome-wide association study identifies IL23R as an inflammatory bowel disease gene. Science 2006, 314, 1461–1463. [Google Scholar] [CrossRef] [PubMed]
  83. Stockinger, B.; Omenetti, S. The dichotomous nature of T helper 17 cells. Nat. Rev. Immunol. 2017, 17, 535–544. [Google Scholar] [CrossRef] [PubMed]
  84. Ahern, P.P.; Schiering, C.; Buonocore, S.; McGeachy, M.J.; Cua, D.J.; Maloy, K.J.; Powrie, F. Interleukin-23 drives intestinal inflammation through direct activity on T cells. Immunity 2010, 33, 279–288. [Google Scholar] [CrossRef] [PubMed]
  85. Griseri, T.; Arnold, I.C.; Pearson, C.; Krausgruber, T.; Schiering, C.; Franchini, F.; Schulthess, J.; McKenzie, B.S.; Crocker, P.R.; Powrie, F. Granulocyte Macrophage Colony-Stimulating Factor-Activated Eosinophils Promote Interleukin-23 Driven Chronic Colitis. Immunity 2015, 43, 187–199. [Google Scholar] [CrossRef] [PubMed]
  86. Izcue, A.; Hue, S.; Buonocore, S.; Arancibia-Cárcamo, C.V.; Ahern, P.P.; Iwakura, Y.; Maloy, K.J.; Powrie, F. Interleukin-23 restrains regulatory T cell activity to drive T cell-dependent colitis. Immunity 2008, 28, 559–570. [Google Scholar] [CrossRef] [PubMed]
  87. Coccia, M.; Harrison, O.J.; Schiering, C.; Asquith, M.J.; Becher, B.; Powrie, F.; Maloy, K.J. IL-1β mediates chronic intestinal inflammation by promoting the accumulation of IL-17A secreting innate lymphoid cells and CD4+ Th17 cells. J. Exp. Med. 2012, 209, 1595–1609. [Google Scholar] [CrossRef]
  88. Sands, B.E.; Sandborn, W.J.; Panaccione, R.; O’Brien, C.D.; Zhang, H.; Johanns, J.; Adedokun, O.J.; Li, K.; Peyrin-Biroulet, L.; Van Assche, G.; et al. Ustekinumab as Induction and Maintenance Therapy for Ulcerative Colitis. N. Engl. J. Med. 2019, 381, 1201–1214. [Google Scholar] [CrossRef]
  89. de Souza, H.S.P.; Fiocchi, C. Immunopathogenesis of IBD: Current state of the art. Nat. Rev. Gastroenterol. Hepatol. 2016, 13, 13–27. [Google Scholar] [CrossRef]
  90. Mahapatro, M.; Erkert, L.; Becker, C. Cytokine-Mediated Crosstalk between Immune Cells and Epithelial Cells in the Gut. Cells 2021, 10, 111. [Google Scholar] [CrossRef]
  91. Bharti, S.; Bharti, M. The Business of T Cell Subsets and Cytokines in the Immunopathogenesis of Inflammatory Bowel Disease. Cureus 2022, 14, e27290. [Google Scholar] [CrossRef]
  92. Pushparaj, P.N.; Li, D.; Komai-Koma, M.; Guabiraba, R.; Alexander, J.; McSharry, C.; Xu, D. Interleukin-33 exacerbates acute colitis via interleukin-4 in mice. Immunology 2013, 140, 70–77. [Google Scholar] [CrossRef] [PubMed]
  93. Sedhom, M.A.K.; Pichery, M.; Murdoch, J.R.; Foligné, B.; Ortega, N.; Normand, S.; Mertz, K.; Sanmugalingam, D.; Brault, L.; Grandjean, T.; et al. Neutralisation of the interleukin-33/ST2 pathway ameliorates experimental colitis through enhancement of mucosal healing in mice. Gut 2013, 62, 1714–1723. [Google Scholar] [CrossRef]
  94. Latiano, A.; Palmieri, O.; Pastorelli, L.; Vecchi, M.; Pizarro, T.T.; Bossa, F.; Merla, G.; Augello, B.; Latiano, T.; Corritore, G.; et al. Associations between genetic polymorphisms in IL-33, IL1R1 and risk for inflammatory bowel disease. PLoS ONE 2013, 8, e62144. [Google Scholar] [CrossRef]
  95. Duan, L.; Chen, J.; Zhang, H.; Yang, H.; Zhu, P.; Xiong, A.; Xia, Q.; Zheng, F.; Tan, Z.; Gong, F.; et al. Interleukin-33 ameliorates experimental colitis through promoting Th2/Foxp3+ regulatory T-cell responses in mice. Mol. Med. 2012, 18, 753–761. [Google Scholar] [CrossRef] [PubMed]
  96. Shohan, M.; Sabzevary-Ghahfarokhi, M.; Bagheri, N.; Shirzad, H.; Rahimian, G.; Soltani, A.; Ghatreh-Samani, M.; Deris, F.; Tahmasbi, K.; Shahverdi, E.; et al. Intensified Th9 Response is Associated with the Immunopathogenesis of Active Ulcerative Colitis. Immunol. Investig. 2018, 47, 700–711. [Google Scholar] [CrossRef] [PubMed]
  97. Petit-Frere, C.; Dugas, B.; Braquet, P.; Mencia-Huerta, J.M. Interleukin-9 potentiates the interleukin-4-induced IgE and IgG1 release from murine B lymphocytes. Immunology 1993, 79, 146–151. [Google Scholar]
  98. Dugas, B.; Renauld, J.C.; Pène, J.; Bonnefoy, J.Y.; Peti-Frère, C.; Braquet, P.; Bousquet, J.; Van Snick, J.; Mencia-Huerta, J.M. Interleukin-9 potentiates the interleukin-4-induced immunoglobulin (IgG, IgM and IgE) production by normal human B lymphocytes. Eur. J. Immunol. 1993, 23, 1687–1692. [Google Scholar] [CrossRef]
  99. Takatsuka, S.; Yamada, H.; Haniuda, K.; Saruwatari, H.; Ichihashi, M.; Renauld, J.-C.; Kitamura, D. IL-9 receptor signaling in memory B cells regulates humoral recall responses. Nat. Immunol. 2018, 19, 1025–1034. [Google Scholar] [CrossRef]
  100. Wang, X.; Zhu, Y.; Zhang, M.; Wang, H.; Jiang, Y.; Gao, P. Ulcerative Colitis Is Characterized by a Decrease in Regulatory B Cells. J. Crohns Colitis 2016, 10, 1212–1223. [Google Scholar] [CrossRef]
  101. Kałużna, A.; Olczyk, P.; Komosińska-Vassev, K. The Role of Innate and Adaptive Immune Cells in the Pathogenesis and Development of the Inflammatory Response in Ulcerative Colitis. J. Clin. Med. 2022, 11, 400. [Google Scholar] [CrossRef] [PubMed]
  102. Brandtzaeg, P.; Baklien, K.; Fausa, O.; Hoel, P.S. Immunohistochemical characterization of local immunoglobulin formation in ulcerative colitis. Gastroenterology 1974, 66, 1123–1136. [Google Scholar] [CrossRef] [PubMed]
  103. Boland, B.S.; He, Z.; Tsai, M.S.; Olvera, J.G.; Omilusik, K.D.; Duong, H.G.; Kim, E.S.; Limary, A.E.; Jin, W.; Milner, J.J.; et al. Heterogeneity and clonal relationships of adaptive immune cells in ulcerative colitis revealed by single-cell analyses. Sci. Immunol. 2020, 5, eabb4432. [Google Scholar] [CrossRef] [PubMed]
  104. Castro-Dopico, T.; Dennison, T.W.; Ferdinand, J.R.; Mathews, R.J.; Fleming, A.; Clift, D.; Stewart, B.J.; Jing, C.; Strongili, K.; Labzin, L.I.; et al. Anti-commensal IgG Drives Intestinal Inflammation and Type 17 Immunity in Ulcerative Colitis. Immunity 2019, 50, 1099–1114.e10. [Google Scholar] [CrossRef] [PubMed]
  105. Wen, C.; Hu, H.; Yang, W.; Zhao, Y.; Zheng, L.; Jiang, X.; Wang, L. Targeted inhibition of FcRn reduces NET formation to ameliorate experimental ulcerative colitis by accelerating ANCA clearance. Int. Immunopharmacol. 2022, 113, 109474. [Google Scholar] [CrossRef] [PubMed]
  106. Wajant, H.; Pfizenmaier, K.; Scheurich, P. Tumor necrosis factor signaling. Cell Death Differ. 2003, 10, 45–65. [Google Scholar] [CrossRef] [PubMed]
  107. So, T.; Ishii, N. The TNF-TNFR Family of Co-signal Molecules. Adv. Exp. Med. Biol. 2019, 1189, 53–84. [Google Scholar] [CrossRef] [PubMed]
  108. Zhang, Y.; Huang, K.; Zhang, Y.; Han, T.; Li, L.; Ruan, C.; Sun, Y.H.; Shi, W.; Han, W.; Wu, S.Q.; et al. A unique death pathway keeps RIPK1 D325A mutant mice in check at embryonic day 10.5. PLoS Biol. 2021, 19, e3001304. [Google Scholar] [CrossRef] [PubMed]
  109. Mocarski, E.S.; Mandal, P. TNF-dependent hyperactivation of RIPK1-dependent cytotoxic signaling during embryogenesis and inflammation. PLoS Biol. 2021, 19, e3001371. [Google Scholar] [CrossRef]
  110. Eissner, G.; Kolch, W.; Scheurich, P. Ligands working as receptors: Reverse signaling by members of the TNF superfamily enhance the plasticity of the immune system. Cytokine Growth Factor Rev. 2004, 15, 353–366. [Google Scholar] [CrossRef]
  111. Vos, A.C.W.; Wildenberg, M.E.; Duijvestein, M.; Verhaar, A.P.; van den Brink, G.R.; Hommes, D.W. Anti-tumor necrosis factor-α antibodies induce regulatory macrophages in an Fc region-dependent manner. Gastroenterology 2011, 140, 221–230. [Google Scholar] [CrossRef] [PubMed]
  112. Nguyen, D.X.; Ehrenstein, M.R. Anti-TNF drives regulatory T cell expansion by paradoxically promoting membrane TNF-TNF-RII binding in rheumatoid arthritis. J. Exp. Med. 2016, 213, 1241–1253. [Google Scholar] [CrossRef] [PubMed]
  113. Szabo, S.J.; Kim, S.T.; Costa, G.L.; Zhang, X.; Fathman, C.G.; Glimcher, L.H. A novel transcription factor, T-bet, directs Th1 lineage commitment. Cell 2000, 100, 655–669. [Google Scholar] [CrossRef] [PubMed]
  114. Verstockt, B.; Salas, A.; Sands, B.E.; Abraham, C.; Leibovitzh, H.; Neurath, M.F.; Vande Casteele, N.; Alimentiv Translational Research Consortium (ATRC). IL-12 and IL-23 pathway inhibition in inflammatory bowel disease. Nat. Rev. Gastroenterol. Hepatol. 2023, 20, 433–446. [Google Scholar] [CrossRef] [PubMed]
  115. Moschen, A.R.; Tilg, H.; Raine, T. IL-12, IL-23 and IL-17 in IBD: Immunobiology and therapeutic targeting. Nat. Rev. Gastroenterol. Hepatol. 2019, 16, 185–196. [Google Scholar] [CrossRef]
  116. Liu, Z.; Yadav, P.K.; Xu, X.; Su, J.; Chen, C.; Tang, M.; Lin, H.; Yu, J.; Qian, J.; Yang, P.C.; et al. The increased expression of IL-23 in inflammatory bowel disease promotes intraepithelial and lamina propria lymphocyte inflammatory responses and cytotoxicity. J. Leukoc. Biol. 2011, 89, 597–606. [Google Scholar] [CrossRef] [PubMed]
  117. Cua, D.J.; Sherlock, J.; Chen, Y.; Murphy, C.A.; Joyce, B.; Seymour, B.; Lucian, L.; To, W.; Kwan, S.; Churakova, T.; et al. Interleukin-23 rather than interleukin-12 is the critical cytokine for autoimmune inflammation of the brain. Nature 2003, 421, 744–748. [Google Scholar] [CrossRef] [PubMed]
  118. Duvallet, E.; Semerano, L.; Assier, E.; Falgarone, G.; Boissier, M.-C. Interleukin-23: A key cytokine in inflammatory diseases. Ann. Med. 2011, 43, 503–511. [Google Scholar] [CrossRef]
  119. Deepak, P.; Sandborn, W.J. Ustekinumab and Anti-Interleukin-23 Agents in Crohn’s Disease. Gastroenterol. Clin. N. Am. 2017, 46, 603–626. [Google Scholar] [CrossRef]
  120. Atreya, R.; Abreu, M.T.; Krueger, J.G.; Eyerich, K.; Sachen, K.; Greving, C.; Hammaker, D.; Bao, P.; Lacy, E.; Sarabia, I.; et al. P504 Guselkumab, an IL-23p19 subunit–specific monoclonal antibody, binds CD64+ myeloid cells and potently neutralises IL-23 produced from the same cells. J. Crohns Colitis 2023, 17, i634–i635. [Google Scholar] [CrossRef]
  121. Walsh, G.M.; Symon, F.A.; Lazarovils, A.L.; Wardlaw, A.J. Integrin α4β7 mediates human eosinophil interaction with MAdCAM-1, VCAM-1, and fibronectin. Immunology 1996, 89, 112–119. [Google Scholar] [CrossRef] [PubMed]
  122. Zhou, W.; Sonnenberg, G.F. Activation and Suppression of Group 3 Innate Lymphoid Cells in the Gut. Trends Immunol. 2020, 41, 721–733. [Google Scholar] [CrossRef] [PubMed]
  123. von Andrian, U.H.; Engelhardt, B. α4 integrins as therapeutic targets in autoimmune disease. N. Engl. J. Med. 2003, 348, 68–72. [Google Scholar] [CrossRef] [PubMed]
  124. Hamann, A.; Andrew, D.P.; Jablonski-Westrich, D.; Holzmann, B.; Butcher, E.C. Role of alpha 4-integrins in lymphocyte homing to mucosal tissues in vivo. J. Immunol. 1994, 152, 3282–3293. [Google Scholar] [CrossRef] [PubMed]
  125. Girard, J.-P.; Moussion, C.; Förster, R. HEVs, lymphatics and homeostatic immune cell trafficking in lymph nodes. Nat. Rev. Immunol. 2012, 12, 762–773. [Google Scholar] [CrossRef] [PubMed]
  126. Matloubian, M.; Lo, C.G.; Cinamon, G.; Lesneski, M.J.; Xu, Y.; Brinkmann, V.; Allende, M.L.; Proia, R.L.; Cyster, J.G. Lymphocyte egress from thymus and peripheral lymphoid organs is dependent on S1P receptor 1. Nature 2004, 427, 355–360. [Google Scholar] [CrossRef] [PubMed]
  127. Rosen, S.D. Ligands for L-selectin: Homing, inflammation, and beyond. Annu. Rev. Immunol. 2004, 22, 129–156. [Google Scholar] [CrossRef] [PubMed]
  128. Berlin, C.; Berg, E.L.; Briskin, M.J.; Andrew, D.P.; Kilshaw, P.J.; Holzmann, B.; Weissman, I.L.; Hamann, A.; Butcher, E.C. α4β7 integrin mediates lymphocyte binding to the mucosal vascular addressin MAdCAM-1. Cell 1993, 74, 185–195. [Google Scholar] [CrossRef]
  129. Pérez-Jeldres, T.; Tyler, C.J.; Boyer, J.D.; Karuppuchamy, T.; Yarur, A.; Giles, D.A.; Yeasmin, S.; Lundborg, L.; Sandborn, W.J.; Patel, D.R.; et al. Targeting Cytokine Signaling and Lymphocyte Traffic via Small Molecules in Inflammatory Bowel Disease: JAK Inhibitors and S1PR Agonists. Front. Pharmacol. 2019, 10, 212. [Google Scholar] [CrossRef]
  130. Darnell, J.E., Jr.; Kerr, I.M.; Stark, G.R. Jak-STAT pathways and transcriptional activation in response to IFNs and other extracellular signaling proteins. Science 1994, 264, 1415–1421. [Google Scholar] [CrossRef]
  131. Levy, D.E.; Darnell, J.E., Jr. Stats: Transcriptional control and biological impact. Nat. Rev. Mol. Cell Biol. 2002, 3, 651–662. [Google Scholar] [CrossRef] [PubMed]
  132. Sekimoto, T.; Yoneda, Y. Nuclear import and export of proteins: The molecular basis for intracellular signaling. Cytokine Growth Factor Rev. 1998, 9, 205–211. [Google Scholar] [CrossRef] [PubMed]
  133. Villarino, A.V.; Kanno, Y.; O’Shea, J.J. Mechanisms and consequences of Jak-STAT signaling in the immune system. Nat. Immunol. 2017, 18, 374–384. [Google Scholar] [CrossRef]
  134. West, N.R.; Hegazy, A.N.; Owens, B.M.J.; Bullers, S.J.; Linggi, B.; Buonocore, S.; Coccia, M.; Görtz, 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, 579–589. [Google Scholar] [CrossRef]
  135. Arijs, I.; Li, K.; Toedter, G.; Quintens, R.; Van Lommel, L.; Van Steen, K.; Leemans, P.; De Hertogh, G.; Lemaire, K.; Ferrante, M.; et al. Mucosal gene signatures to predict response to infliximab in patients with ulcerative colitis. Gut 2009, 58, 1612–1619. [Google Scholar] [CrossRef]
  136. Mavragani, C.P.; Nezos, A.; Dovrolis, N.; Andreou, N.-P.; Legaki, E.; Sechi, L.A.; Bamias, G.; Gazouli, M. Type I and II Interferon Signatures Can Predict the Response to Anti-TNF Agents in Inflammatory Bowel Disease Patients: Involvement of the Microbiota. Inflamm. Bowel Dis. 2020, 26, 1543–1553. [Google Scholar] [CrossRef]
  137. Verstockt, B.; Verstockt, S.; Dehairs, J.; Ballet, V.; Blevi, H.; Wollants, W.-J.; Breynaert, C.; Van Assche, G.; Vermeire, S.; Ferrante, M. Low TREM1 expression in whole blood predicts anti-TNF response in inflammatory bowel disease. EBioMedicine 2019, 40, 733–742. [Google Scholar] [CrossRef] [PubMed]
  138. Pavlidis, P.; Tsakmaki, A.; Pantazi, E.; Li, K.; Cozzetto, D.; Digby-Bell, J.; Yang, F.; Lo, J.W.; Alberts, E.; Sa, A.C.C.; et al. Interleukin-22 regulates neutrophil recruitment in ulcerative colitis and is associated with resistance to ustekinumab therapy. Nat. Commun. 2022, 13, 5820. [Google Scholar] [CrossRef]
  139. Bertani, L.; Baglietto, L.; Antonioli, L.; Fornai, M.; Tapete, G.; Albano, E.; Ceccarelli, L.; Mumolo, M.G.; Pellegrini, C.; Lucenteforte, E.; et al. Assessment of serum cytokines predicts clinical and endoscopic outcomes to vedolizumab in ulcerative colitis patients. Br. J. Clin. Pharmacol. 2020, 86, 1296–1305. [Google Scholar] [CrossRef]
  140. Soendergaard, C.; Seidelin, J.B.; Steenholdt, C.; Nielsen, O.H. Putative biomarkers of vedolizumab resistance and underlying inflammatory pathways involved in IBD. BMJ Open Gastroenterol. 2018, 5, e000208. [Google Scholar] [CrossRef]
  141. Smillie, C.S.; Biton, M.; Ordovas-Montanes, J.; Sullivan, K.M.; Burgin, G.; Graham, D.B.; Herbst, R.H.; Rogel, N.; Slyper, M.; Waldman, J.; et al. Intra- and Inter-cellular Rewiring of the Human Colon during Ulcerative Colitis. Cell 2019, 178, 714–730.e22. [Google Scholar] [CrossRef] [PubMed]
  142. Friedrich, M.; Pohin, M.; Jackson, M.A.; Korsunsky, I.; Bullers, S.J.; 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, 1970–1981. [Google Scholar] [CrossRef] [PubMed]
  143. Li, H.; Feng, C.; Fan, C.; Yang, Y.; Yang, X.; Lu, H.; Lu, Q.; Zhu, F.; Xiang, C.; Zhang, Z.; et al. Intervention of oncostatin M-driven mucosal inflammation by berberine exerts therapeutic property in chronic ulcerative colitis. Cell Death Dis. 2020, 11, 271. [Google Scholar] [CrossRef] [PubMed]
  144. Roblin, X.; Serone, A.; Yoon, O.K.; Zhuo, L.; Grant, E.; Woo, J.; Liu, J.; Galien, R.; D’Haens, G. Effects of JAK1-Preferential Inhibitor Filgotinib on Circulating Biomarkers and Whole Blood Genes/Pathways of Patients with Moderately to Severely Active Crohn’s Disease. Inflamm. Bowel Dis. 2022, 28, 1207–1218. [Google Scholar] [CrossRef] [PubMed]
  145. Cho, H.; Kelsall, B.L. The role of type I interferons in intestinal infection, homeostasis, and inflammation. Immunol. Rev. 2014, 260, 145–167. [Google Scholar] [CrossRef] [PubMed]
  146. Kotredes, K.P.; Thomas, B.; Gamero, A.M. The Protective Role of Type I Interferons in the Gastrointestinal Tract. Front. Immunol. 2017, 8, 410. [Google Scholar] [CrossRef] [PubMed]
  147. Wang, L.; Wang, Y.; Song, Z.; Chu, J.; Qu, X. Deficiency of interferon-gamma or its receptor promotes colorectal cancer development. J. Interf. Cytokine Res. 2015, 35, 273–280. [Google Scholar] [CrossRef]
  148. Padoan, A.; Musso, G.; Contran, N.; Basso, D. Inflammation, Autoinflammation and Autoimmunity in Inflammatory Bowel Diseases. Curr. Issues Mol. Biol. 2023, 45, 5534–5557. [Google Scholar] [CrossRef]
  149. Østvik, A.E.; Svendsen, T.D.; van Beelen Granlund, A.; Doseth, B.; Skovdahl, H.K.; Bakke, I.; Thorsvik, S.; Afroz, W.; Walaas, G.A.; Mollnes, T.E.; et al. Intestinal Epithelial Cells Express Immunomodulatory ISG15 During Active Ulcerative Colitis and Crohn’s Disease. J. Crohns Colitis 2020, 14, 920–934. [Google Scholar] [CrossRef]
  150. Samie, M.; Lim, J.; Verschueren, E.; Baughman, J.M.; Peng, I.; Wong, A.; Kwon, Y.; Senbabaoglu, Y.; Hackney, J.A.; Keir, M.; et al. Selective autophagy of the adaptor TRIF regulates innate inflammatory signaling. Nat. Immunol. 2018, 19, 246–254. [Google Scholar] [CrossRef]
  151. Flood, P.; Fanning, A.; Woznicki, J.A.; Crowley, T.; Christopher, A.; Vaccaro, A.; Houston, A.; McSweeney, S.; Ross, S.; Hogan, A.; et al. DNA sensor-associated type I interferon signaling is increased in ulcerative colitis and induces JAK-dependent inflammatory cell death in colonic organoids. Am. J. Physiol. Gastrointest. Liver Physiol. 2022, 323, G439–G460. [Google Scholar] [CrossRef] [PubMed]
  152. Woznicki, J.A.; Saini, N.; Flood, P.; Rajaram, S.; Lee, C.M.; Stamou, P.; Skowyra, A.; Bustamante-Garrido, M.; Regazzoni, K.; Crawford, N.; et al. TNF-α synergises with IFN-γ to induce caspase-8-JAK1/2-STAT1-dependent death of intestinal epithelial cells. Cell Death Dis. 2021, 12, 864. [Google Scholar] [CrossRef] [PubMed]
  153. Theobald, V.; Schmitt, F.C.F.; Middel, C.S.; Gaissmaier, L.; Brenner, T.; Weigand, M.A. Triggering receptor expressed on myeloid cells-1 in sepsis, and current insights into clinical studies. Crit. Care 2024, 28, 17. [Google Scholar] [CrossRef] [PubMed]
  154. Weber, B.; Schuster, S.; Zysset, D.; Rihs, S.; Dickgreber, N.; Schürch, C.; Riether, C.; Siegrist, M.; Schneider, C.; Pawelski, H.; et al. TREM-1 deficiency can attenuate disease severity without affecting pathogen clearance. PLoS Pathog. 2014, 10, e1003900. [Google Scholar] [CrossRef] [PubMed]
  155. Czarnewski, P.; Parigi, S.M.; Sorini, C.; Diaz, O.E.; Das, S.; Gagliani, N.; Villablanca, E.J. Conserved transcriptomic profile between mouse and human colitis allows unsupervised patient stratification. Nat. Commun. 2019, 10, 2892. [Google Scholar] [CrossRef] [PubMed]
  156. Prins, M.M.; Verstockt, B.; Ferrante, M.; Vermeire, S.; Wildenberg, M.E.; Koelink, P.J. Monocyte TREM-1 Levels Associate with Anti-TNF Responsiveness in IBD Through Autophagy and Fcγ-Receptor Signaling Pathways. Front. Immunol. 2021, 12, 627535. [Google Scholar] [CrossRef] [PubMed]
  157. Kökten, T.; Gibot, S.; Lepage, P.; D’Alessio, S.; Hablot, J.; Ndiaye, N.-C.; Busby-Venner, H.; Monot, C.; Garnier, B.; Moulin, D.; et al. TREM-1 Inhibition Restores Impaired Autophagy Activity and Reduces Colitis in Mice. J. Crohns Colitis 2018, 12, 230–244. [Google Scholar] [CrossRef] [PubMed]
  158. Roca, H.; Varsos, Z.S.; Sud, S.; Craig, M.J.; Ying, C.; Pienta, K.J. CCL2 and interleukin-6 promote survival of human CD11b+ peripheral blood mononuclear cells and induce M2-type macrophage polarization. J. Biol. Chem. 2009, 284, 34342–34354. [Google Scholar] [CrossRef]
  159. Chang, C.-P.; Su, Y.-C.; Hu, C.-W.; Lei, H.-Y. TLR2-dependent selective autophagy regulates NF-κB lysosomal degradation in hepatoma-derived M2 macrophage differentiation. Cell Death Differ. 2013, 20, 515–523. [Google Scholar] [CrossRef]
  160. Levine, B.; Klionsky, D.J. Development by self-digestion: Molecular mechanisms and biological functions of autophagy. Dev. Cell 2004, 6, 463–477. [Google Scholar] [CrossRef]
  161. Kaser, A.; Lee, A.-H.; Franke, A.; Glickman, J.N.; Zeissig, S.; Tilg, H.; Nieuwenhuis, E.E.; Higgins, D.E.; Schreiber, S.; Glimcher, L.H.; et al. XBP1 links ER stress to intestinal inflammation and confers genetic risk for human inflammatory bowel disease. Cell 2008, 134, 743–756. [Google Scholar] [CrossRef] [PubMed]
  162. Tréton, X.; Pédruzzi, E.; Cazals-Hatem, D.; Grodet, A.; Panis, Y.; Groyer, A.; Moreau, R.; Bouhnik, Y.; Daniel, F.; Ogier-Denis, E. Altered endoplasmic reticulum stress affects translation in inactive colon tissue from patients with ulcerative colitis. Gastroenterology 2011, 141, 1024–1035. [Google Scholar] [CrossRef]
  163. Hosomi, S.; Kaser, A.; Blumberg, R.S. Role of endoplasmic reticulum stress and autophagy as interlinking pathways in the pathogenesis of inflammatory bowel disease. Curr. Opin. Gastroenterol. 2015, 31, 81–88. [Google Scholar] [CrossRef] [PubMed]
  164. Luo, K.; Cao, S.S. Endoplasmic reticulum stress in intestinal epithelial cell function and inflammatory bowel disease. Gastroenterol. Res. Pract. 2015, 2015, 328791. [Google Scholar] [CrossRef] [PubMed]
  165. Schmitt, H.; Billmeier, U.; Dieterich, W.; Rath, T.; Sonnewald, S.; Reid, S.; Hirschmann, S.; Hildner, K.; Waldner, M.J.; Mudter, J.; et al. Expansion of IL-23 receptor bearing TNFR2+ T cells is associated with molecular resistance to anti-TNF therapy in Crohn’s disease. Gut 2019, 68, 814–828. [Google Scholar] [CrossRef] [PubMed]
  166. Bouwman, W.; Verhaegh, W.; van de Stolpe, A. Improved diagnosis of inflammatory bowel disease and prediction and monitoring of response to anti-TNF alpha treatment based on measurement of signal transduction pathway activity. Front. Pharmacol. 2022, 13, 1008976, Correction in Front. Pharmacol. 2022, 14, 1183639. [Google Scholar] [CrossRef] [PubMed]
  167. Feagan, B.G.; Sandborn, W.J.; D’Haens, G.; Panés, J.; Kaser, A.; Ferrante, M.; Louis, E.; Franchimont, D.; Dewit, O.; Seidler, U.; et al. Induction therapy with the selective interleukin-23 inhibitor risankizumab in patients with moderate-to-severe Crohn’s disease: A randomised, double-blind, placebo-controlled phase 2 study. Lancet 2017, 389, 1699–1709. [Google Scholar] [CrossRef] [PubMed]
  168. Rath, T.; Billmeier, U.; Ferrazzi, F.; Vieth, M.; Ekici, A.; Neurath, M.F.; Atreya, R. Effects of Anti-Integrin Treatment with Vedolizumab on Immune Pathways and Cytokines in Inflammatory Bowel Diseases. Front. Immunol. 2018, 9, 1700. [Google Scholar] [CrossRef] [PubMed]
  169. Schneider, I.; Allner, C.; Mühl, L.; Melde, M.; Lissner, D.; Mantzivi, E.; Glauben, R.; Vitali, F.; Becker, E.; Atreya, I.; et al. Expression and function of α4β7 integrin predict the success of vedolizumab treatment in inflammatory bowel disease. Transl. Res. 2023, 253, 8–15. [Google Scholar] [CrossRef]
  170. Fuchs, F.; Schillinger, D.; Atreya, R.; Hirschmann, S.; Fischer, S.; Neufert, C.; Atreya, I.; Neurath, M.F.; Zundler, S. Clinical Response to Vedolizumab in Ulcerative Colitis Patients Is Associated with Changes in Integrin Expression Profiles. Front. Immunol. 2017, 8, 764. [Google Scholar] [CrossRef]
  171. Oshima, T.; Jordan, P.; Grisham, M.B.; Alexander, J.S.; Jennings, M.; Sasaki, M.; Manas, K. TNF-α induced endothelial MAdCAM-1 expression is regulated by exogenous, not endogenous nitric oxide. BMC Gastroenterol. 2001, 1, 5. [Google Scholar] [CrossRef] [PubMed]
  172. Cromer, W.E.; Mathis, J.M.; Granger, D.N.; Chaitanya, G.V.; Alexander, J.S. Role of the endothelium in inflammatory bowel diseases. World J. Gastroenterol. 2011, 17, 578–593. [Google Scholar] [CrossRef] [PubMed]
  173. Bhatti, M.; Chapman, P.; Peters, M.; Haskard, D.; Hodgson, H.J. Visualising E-selectin in the detection and evaluation of inflammatory bowel disease. Gut 1998, 43, 40–47. [Google Scholar] [CrossRef] [PubMed]
  174. Allner, C.; Melde, M.; Becker, E.; Fuchs, F.; Mühl, L.; Klenske, E.; Müller, L.; Morgenstern, N.; Fietkau, K.; Hirschmann, S.; et al. Baseline levels of dynamic CD4+ T cell adhesion to MAdCAM-1 correlate with clinical response to vedolizumab treatment in ulcerative colitis: A cohort study. BMC Gastroenterol. 2020, 20, 103. [Google Scholar] [CrossRef] [PubMed]
  175. Rivera-Nieves, J.; Olson, T.; Bamias, G.; Bruce, A.; Solga, M.; Knight, R.F.; Hoang, S.; Cominelli, F.; Ley, K. L-selectin, α4β1, and α4β7 integrins participate in CD4+ T cell recruitment to chronically inflamed small intestine. J. Immunol. 2005, 174, 2343–2352. [Google Scholar] [CrossRef]
  176. Zundler, S.; Neurath, M.F. Pathogenic T cell subsets in allergic and chronic inflammatory bowel disorders. Immunol. Rev. 2017, 278, 263–276. [Google Scholar] [CrossRef] [PubMed]
  177. Boden, E.K.; Shows, D.M.; Chiorean, M.V.; Lord, J.D. Identification of Candidate Biomarkers Associated with Response to Vedolizumab in Inflammatory Bowel Disease. Dig Dis. Sci. 2018, 63, 2419–2429. [Google Scholar] [CrossRef] [PubMed]
  178. Dulai, P.S.; Boland, B.S.; Singh, S.; Chaudrey, K.; Koliani-Pace, J.L.; Kochhar, G.; Parikh, M.P.; Shmidt, E.; Hartke, J.; Chilukuri, P.; et al. Development and Validation of a Scoring System to Predict Outcomes of Vedolizumab Treatment in Patients with Crohn’s Disease. Gastroenterology 2018, 155, 687–695.e10. [Google Scholar] [CrossRef]
  179. Dulai, P.S.; Singh, S.; Vande Casteele, N.; Meserve, J.; Winters, A.; Chablaney, S.; Aniwan, S.; Shashi, P.; Kochhar, G.; Weiss, A.; et al. Development and Validation of Clinical Scoring Tool to Predict Outcomes of Treatment with Vedolizumab in Patients with Ulcerative Colitis. Clin. Gastroenterol. Hepatol. 2020, 18, 2952–2961.e8. [Google Scholar] [CrossRef]
  180. Gordon, H.; Rodger, B.; Lindsay, J.O.; Stagg, A.J. Recruitment and Residence of Intestinal T Cells—Lessons for Therapy in Inflammatory Bowel Disease. J. Crohns Colitis 2023, 17, 1326–1341. [Google Scholar] [CrossRef]
  181. Wetwittayakhlang, P.; Lakatos, P.L. Current Evidence for Combined Targeted Therapy for the Treatment of Inflammatory Bowel Disease. J. Can. Assoc. Gastroenterol. 2024, 7, 22–29. [Google Scholar] [CrossRef] [PubMed]
  182. Kolar, M.; Lukas, M.; Malickova, K.; Prochazkova, L.; Bortlik, M.; Duricova, D.; Hruba, V.; Machkova, N.; Mitrova, K.; Vasatko, M.; et al. P368 Tofacitinib induction efficiency and intracellular cytokine dynamics in ulcerative colitis: Results from clinical practice. J. Crohns Colitis 2020, 14, S348. [Google Scholar] [CrossRef]
  183. Ananthakrishnan, A.N. Precision medicine in inflammatory bowel diseases. Intest. Res. 2024, 22, 8–14. [Google Scholar] [CrossRef] [PubMed]
  184. Song, E.M.; Yang, S.-K. Natural history of inflammatory bowel disease: A comparison between the East and the West. Intest. Res. 2022, 20, 418–430. [Google Scholar] [CrossRef]
Figure 1. The role of cytokines that contribute to intestinal homeostasis. Intestinal homeostasis is maintained via interactions between IECs, immune cells, gut microbiota, and mesenchymal cells, and is tightly regulated by cytokines. Th17 cells and ILC3 promote IEC proliferation and survival, as well as antimicrobial substances and mucus secretion through the production of IL-22, IL-17, and IL-6. Mononuclear phagocytes enhance barrier integrity through the production of type I IFN. IL-10 produced by Tregs and Tr1 regulates excessive immune responses and maintains intestinal homeostasis. In addition, IgA secreted by B cells promotes effective defense by coating pathogenic bacteria from interacting directly with the epithelium. Abbreviations: AHR, aryl hydrocarbon receptor; BAFF, B cell activating factor; DC, dendric cell; IEC, intestinal epithelial cell; IFN, interferon; Ig, immunoglobulin; IL, interleukin; ILC, innate lymphoid cell; MAIT, mucosal-associated invariant T cells; RA, retinoic acid; SAA, serum amyloid A; TGF-β, transforming growth factor-β; Th, T helper; Tr1, Type 1 regulatory; Treg, T regulatory.
Figure 1. The role of cytokines that contribute to intestinal homeostasis. Intestinal homeostasis is maintained via interactions between IECs, immune cells, gut microbiota, and mesenchymal cells, and is tightly regulated by cytokines. Th17 cells and ILC3 promote IEC proliferation and survival, as well as antimicrobial substances and mucus secretion through the production of IL-22, IL-17, and IL-6. Mononuclear phagocytes enhance barrier integrity through the production of type I IFN. IL-10 produced by Tregs and Tr1 regulates excessive immune responses and maintains intestinal homeostasis. In addition, IgA secreted by B cells promotes effective defense by coating pathogenic bacteria from interacting directly with the epithelium. Abbreviations: AHR, aryl hydrocarbon receptor; BAFF, B cell activating factor; DC, dendric cell; IEC, intestinal epithelial cell; IFN, interferon; Ig, immunoglobulin; IL, interleukin; ILC, innate lymphoid cell; MAIT, mucosal-associated invariant T cells; RA, retinoic acid; SAA, serum amyloid A; TGF-β, transforming growth factor-β; Th, T helper; Tr1, Type 1 regulatory; Treg, T regulatory.
Biomedicines 12 00952 g001
Figure 2. The role of cytokines that contribute to the pathophysiology of UC. The interaction of DAMPs and PAMPs with PRRs on the IECs and immune cells activates the innate immune system. Inflammatory cytokines such as IL-1β, TNF-α, and IL-6 play a central role in this process. These cytokines exert multifaceted effects on immune and non-immune cells, exacerbating the pathophysiology of UC. In UC, abnormally activated neutrophils release MMPs, ROS, elastases, iNOS, and calprotectin, which damage IECs. In UC, monocytes and macrophages with upregulated PRRs proliferate and have increased susceptibility to bacteria. ILCs promote the pathophysiology of UC via the release of IL-17, TNF-α, and IFN-γ. Naïve CD4+ T cells are activated by APCs in lymphoid tissues to upregulate homing receptors such as chemokine receptors and α4β7 integrin for T cell distribution to mucosa. Naïve CD4+ T cells differentiate into distinct subsets through the microenvironment of various cytokines and activation and repression of transcription factors, which contribute to the pathophysiology of UC. B cells are also involved in the pathophysiology of UC, as IgG is increased in the inflamed mucosa of patients with UC. Furthermore, OSM, IL-6, and IL-17 activate downstream signaling of mesenchymal cells, which promote inflammation. Abbreviations: APC, antigen-presenting cell; CD, cluster of differentiation; CTGF, connective tissue growth factor; CXCL, C-X-C motif ligand; DAMP, damage-associated molecular pattern; DC, dendric cell; FAP, fibroblast activating protein; IEC, intestinal epithelial cell; IFN, interferon; IL, interleukin; ILC, innate lymphoid cell; Ig, immunoglobulin; iNOS, inducible nitric oxide synthase; MAdCAM-1, mucosal addressin cell adhesion molecule 1; MMP, matrix metalloproteinase; OSM, oncostatin M; PAMP, pathogen-associated molecular pattern; PRR, pattern recognition receptor; ROS, reactive oxygen species; TGF-β, transforming growth factor-β; Th, T helper; TNF, anti-tumor necrosis factor; UC, ulcerative colitis.
Figure 2. The role of cytokines that contribute to the pathophysiology of UC. The interaction of DAMPs and PAMPs with PRRs on the IECs and immune cells activates the innate immune system. Inflammatory cytokines such as IL-1β, TNF-α, and IL-6 play a central role in this process. These cytokines exert multifaceted effects on immune and non-immune cells, exacerbating the pathophysiology of UC. In UC, abnormally activated neutrophils release MMPs, ROS, elastases, iNOS, and calprotectin, which damage IECs. In UC, monocytes and macrophages with upregulated PRRs proliferate and have increased susceptibility to bacteria. ILCs promote the pathophysiology of UC via the release of IL-17, TNF-α, and IFN-γ. Naïve CD4+ T cells are activated by APCs in lymphoid tissues to upregulate homing receptors such as chemokine receptors and α4β7 integrin for T cell distribution to mucosa. Naïve CD4+ T cells differentiate into distinct subsets through the microenvironment of various cytokines and activation and repression of transcription factors, which contribute to the pathophysiology of UC. B cells are also involved in the pathophysiology of UC, as IgG is increased in the inflamed mucosa of patients with UC. Furthermore, OSM, IL-6, and IL-17 activate downstream signaling of mesenchymal cells, which promote inflammation. Abbreviations: APC, antigen-presenting cell; CD, cluster of differentiation; CTGF, connective tissue growth factor; CXCL, C-X-C motif ligand; DAMP, damage-associated molecular pattern; DC, dendric cell; FAP, fibroblast activating protein; IEC, intestinal epithelial cell; IFN, interferon; IL, interleukin; ILC, innate lymphoid cell; Ig, immunoglobulin; iNOS, inducible nitric oxide synthase; MAdCAM-1, mucosal addressin cell adhesion molecule 1; MMP, matrix metalloproteinase; OSM, oncostatin M; PAMP, pathogen-associated molecular pattern; PRR, pattern recognition receptor; ROS, reactive oxygen species; TGF-β, transforming growth factor-β; Th, T helper; TNF, anti-tumor necrosis factor; UC, ulcerative colitis.
Biomedicines 12 00952 g002
Figure 3. Factors contributing to resistance to treatment with anti-TNF-α antibodies. OSM and IL-1β induce chemokine production by inflammatory fibroblasts and promote neutrophil recruitment. This pathway may be resistant to treatment with anti-TNF-α antibodies. IFN induces epithelial cell death that is resistant to treatment with anti-TNF-α antibodies. Macrophages expressing high levels of TREM-1 are resistant to differentiation into the regulatory phenotype induced by anti-TNF-α antibodies. T cells expressing TNFR2 and IL-23R are resistant to apoptosis induced by anti-TNF-α antibodies. Abbreviation: DC, dendric cell; IFN, interferon; IL, interleukin; ILC, innate lymphoid cell; IL-23R, interleukin-23 receptor; OSM, oncostatin M; Th, T helper; TNF, anti-tumor necrosis factor; TNFR2, tumor necrosis factor receptor 2; TREM-1, triggering receptor expressed on myeloid cells 1.
Figure 3. Factors contributing to resistance to treatment with anti-TNF-α antibodies. OSM and IL-1β induce chemokine production by inflammatory fibroblasts and promote neutrophil recruitment. This pathway may be resistant to treatment with anti-TNF-α antibodies. IFN induces epithelial cell death that is resistant to treatment with anti-TNF-α antibodies. Macrophages expressing high levels of TREM-1 are resistant to differentiation into the regulatory phenotype induced by anti-TNF-α antibodies. T cells expressing TNFR2 and IL-23R are resistant to apoptosis induced by anti-TNF-α antibodies. Abbreviation: DC, dendric cell; IFN, interferon; IL, interleukin; ILC, innate lymphoid cell; IL-23R, interleukin-23 receptor; OSM, oncostatin M; Th, T helper; TNF, anti-tumor necrosis factor; TNFR2, tumor necrosis factor receptor 2; TREM-1, triggering receptor expressed on myeloid cells 1.
Biomedicines 12 00952 g003
Table 1. List of cytokines by JAK family.
Table 1. List of cytokines by JAK family.
JAK1 and
JAK3
JAK1 and
JAK2
JAK1 and
TYK2
JAK1 and
JAK2, TYK2
JAK2 and
TYK2
JAK2 and
JAK2
CytokineIL-2, IL-4, IL-7, IL-9, IL-15, IL-21IFN-γIFN-α, IFN-β, IL-22, IL-26, IL-10IL-11, IL-13, IL-6, OSM, LIFIL-12, IL-23, TPOEPO, GH, IL-3, GM-CSF, IL-5
JAK inhibitorTofa, Filgo, UpaTofa, Filgo, UpaTofa, Filgo, UpaTofa, Filgo, Upa(Tofa)(Tofa)
Describes the relationship between cytokine receptors and the corresponding JAK family members. Abbreviations: EPO, erythropoietin; Filgo, filgotinib; GH, growth hormone; GM-CSF, granulocyte–macrophage colony-stimulating factor; IFN, interferon; IL, interleukin; JAK, Janus kinase; LIF, leukemia inhibitory factor; OSM, oncostatin M; Tofa, tofacitinib; TPO, thrombopoietin; TYK, tyrosine kinase; Upa, upadacitinib.
Table 2. Cytokine profiles and therapeutic efficacy of biologics.
Table 2. Cytokine profiles and therapeutic efficacy of biologics.
MoleculeSampleMeasurementsOutcomePredicting Treatment EffectReference
Anti-TNF-α
antibodies
OSM, OSMRMucosamRNA (qPCR)Endoscopic and histologic remissionMucosal healing (based on endoscopic and histologic criteria) was achieved in 69–85% of patients with low OSM module expression, but only 10–15% of patients with high OSM module expression.[134]
Panel (IL-13Rα2, TNFRSF11B,
IL-11, STC1, PTGS2)
MucosamRNA (qPCR)Endoscopic and histologic remissionThe panel divided responders and non-responders, with a sensitivity of 0.95 and specificity of 0.85.[135]
IFNBloodmRNA (qPCR)Clinical and endoscopic remission and normalization of CRPA low type I IFN signature score predicted response to anti-TNF-α antibody with an AUC of 0.95, sensitivity of 0.93, specificity of 0.88, PPV of 0.87, and NPV of 0.93. A low type II IFN signature score predicted response to anti-TNF-α antibodies with an AUC of 0.87, sensitivity of 0.86, specificity of 0.75, PPV of 0.75, and NPV of 0.86.[136]
TREM-1Blood,
Mucosa
mRNA (qPCR)Endoscopic remission
(MES ≤ 1)
Low whole blood and mucosal TREM-1 mRNA levels predicted response to anti-TNF-α antibodies with AUCs of 0.78 (95% CI 0.65–0.90, p = 0.001) and 0.77 (95% CI 0.62–0.92, p = 0.003).[137]
Panel (TNF-α, IL-12, IL-8, IL-2, IL-5, IL-1β, IFN-γ)BloodConcentrationEndoscopic and histologic remissionThe cytokine score had a sensitivity of 0.84, specificity of 0.93, and accuracy rate of 0.90 (44/49) for predicting response to anti-TNF-α antibodies.[16]
Anti-IL-12/23
antibodies
IL-22MucosamRNA (qPCR)Clinical remission
(Mayo score of ≤2 and no subscore > 1) and
Mucosal healing
(Endoscopic and histologic remission)
Patients with low IL-22 enrichment scores had approximately twice as many clinical remissions (25% vs. 13%) and mucosal healing (26% vs. 16%) as all patients were not stratified.[138]
Anti-α4β7 integrin
antibodies
IL-6, IL-8BloodConcentrationClinical remission (partial Mayo score of <2) and
Endoscopic remission (MES of 0 or 1)
High serum IL-6 and IL-8 levels at baseline and decreased IL-6 and IL-8 levels 6 weeks after introduction of anti-α4β7 integrin antibodies predicted clinical remission with a sensitivity of 0.83 and specificity of 0.87, and endoscopic remission with a sensitivity of 0.82 and specificity of 0.90.[139]
IL-6BloodConcentrationNon-response (≤2 point decrease in Mayo score from baseline, 0 point decrease in rectal bleeding score or ≥1 point in rectal bleeding score)High baseline serum IL-6 levels predicted resistance to vedolizumab with an AUC of 0.77 (95% CI: 0.57–0.98), sensitivity of 0.79, and specificity of 0.88.[140]
Abbreviations: AUC, area under the curve; CI, confidence interval; CRP, C-reactive protein; IFN, interferon; IL, interleukin; MES, mayo endoscopic subscore; NPV, negative predictive value; OSM, oncostatin M; OSMR, oncostatin M receptor; PPV, positive predictive value; PTGS2, prostaglandin-endoperoxide synthase 2; qPCR, quantitative polymerase chain reaction; STC1, stanniocalcin 1; TNF, anti-tumor necrosis factor; TNFRSF11B, tumor necrosis factor receptor superfamily member 11B; TREM-1, triggering receptor expressed on myeloid cells 1.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kurumi, H.; Yokoyama, Y.; Hirano, T.; Akita, K.; Hayashi, Y.; Kazama, T.; Isomoto, H.; Nakase, H. Cytokine Profile in Predicting the Effectiveness of Advanced Therapy for Ulcerative Colitis: A Narrative Review. Biomedicines 2024, 12, 952. https://doi.org/10.3390/biomedicines12050952

AMA Style

Kurumi H, Yokoyama Y, Hirano T, Akita K, Hayashi Y, Kazama T, Isomoto H, Nakase H. Cytokine Profile in Predicting the Effectiveness of Advanced Therapy for Ulcerative Colitis: A Narrative Review. Biomedicines. 2024; 12(5):952. https://doi.org/10.3390/biomedicines12050952

Chicago/Turabian Style

Kurumi, Hiroki, Yoshihiro Yokoyama, Takehiro Hirano, Kotaro Akita, Yuki Hayashi, Tomoe Kazama, Hajime Isomoto, and Hiroshi Nakase. 2024. "Cytokine Profile in Predicting the Effectiveness of Advanced Therapy for Ulcerative Colitis: A Narrative Review" Biomedicines 12, no. 5: 952. https://doi.org/10.3390/biomedicines12050952

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

Article Metrics

Back to TopTop