Basic Study Open Access
Copyright ©The Author(s) 2018. Published by Baishideng Publishing Group Inc. All rights reserved.
World J Gastroenterol. Mar 21, 2018; 24(11): 1196-1205
Published online Mar 21, 2018. doi: 10.3748/wjg.v24.i11.1196
NOD2- and disease-specific gene expression profiles of peripheral blood mononuclear cells from Crohn’s disease patients
Holger Schäffler, Maria Rohde, Sarah Rohde, Astrid Huth, Nicole Gittel, Hannes Hollborn, Georg Lamprecht, Robert Jaster, Department of Medicine II, Division of Gastroenterology, Rostock University Medical Center, Rostock 18057, Germany
Dirk Koczan, Institute of Immunology, Rostock University Medical Center, Rostock 18057, Germany
Änne Glass, Institute for Biostatistics and Informatics in Medicine and Ageing Research, Rostock 18057, Germany
ORCID number: Holger Schaffler (0000-0002-8475-3741); Maria Rohde (0000-0001-8924-3292); Sarah Rohde (0000-0002-6892-1565); Astrid Huth (0000-0002-0858-7948); Hannes Hollborn (0000-0002-2366-4390); Nicole Gittel (0000-0002-7134-069X); Dirk Koczan (0000-0002-3183-998X); Änne Glass (0000-0002-7715-9058); Georg Lamprecht (0000-0003-0997-3135); Robert Jaster (0000-0002-8220-4570).
Author contributions: Schäffler H, Rohde S and Jaster R designed the study; Huth A, Schäffler H and Lamprecht G took responsibility for patient care and follow-up; Rohde M, Rohde S, Hollborn H, Jaster R and Koczan D (microarray studies) performed the experiments; Gittel N, Huth A and Schäffler H collected the samples and performed the clinical characterization of the patients; Glass Ä performed the biostatistics; all authors analyzed the data; and Schäffler H and Jaster R wrote the manuscript.
Supported by a grant from the Damp-Foundation (2016-04) to Schäffler H and Rohde S.
Institutional review board statement: The study was approved by the ethics board of the Medical Faculty of the University of Rostock (A 2015-0042). Written informed consent was obtained from each participant prior to enrollment.
Conflict-of-interest statement: The authors declare that there is no conflict of interest.
Data sharing statement: No additional data are available.
Open-Access: This article is an open-access article which was selected by an in-house editor and fully peer-reviewed by external reviewers. It is distributed in accordance with the Creative Commons Attribution Non Commercial (CC BY-NC 4.0) license, which permits others to distribute, remix, adapt, build upon this work non-commercially, and license their derivative works on different terms, provided the original work is properly cited and the use is non-commercial. See: http://creativecommons.org/licenses/by-nc/4.0/
Correspondence to: Robert Jaster, MD, Academic Research, Professor, Senior Scientist, Department of Medicine II, Division of Gastroenterology, Rostock University Medical Center, E.-Heydemann-Str. 6, Rostock 18057, Germany. robert.jaster@med.uni-rostock.de
Telephone: +49-381-4947349 Fax: +49-381-4947482
Received: December 21, 2017
Peer-review started: December 21, 2017
First decision: January 18, 2018
Revised: January 29, 2018
Accepted: February 1, 2018
Article in press: February 1, 2018
Published online: March 21, 2018

Abstract
AIM

To investigate disease-specific gene expression profiles of peripheral blood mononuclear cells (PBMCs) from Crohn’s disease (CD) patients in clinical remission.

METHODS

Patients with CD in clinical remission or with very low disease activity according to the Crohn’s disease activity index were genotyped regarding nucleotide-binding oligomerization domain 2 (NOD2), and PBMCs from wild-type (WT)-NOD2 patients, patients with homozygous or heterozygous NOD2 mutations and healthy donors were isolated for further analysis. The cells were cultured with vitamin D, peptidoglycan (PGN) and lipopolysaccharide (LPS) for defined periods of time before RNA was isolated and subjected to microarray analysis using Clariom S assays and quantitative real-time PCR. NOD2- and disease-specific gene expression profiles were evaluated with repeated measure ANOVA by a general linear model.

RESULTS

Employing microarray assays, a total of 267 genes were identified that were significantly up- or downregulated in PBMCs of WT-NOD2 patients, compared to healthy donors after challenge with vitamin D and/or a combination of LPS and PGN (P < 0.05; threshold: ≥ 2-fold change). For further analysis by real-time PCR, genes with known impact on inflammation and immunity were selected that fulfilled predefined expression criteria. In a larger cohort of patients and controls, a disease-associated expression pattern, with higher transcript levels in vitamin D-treated PBMCs from patients, was observed for three of these genes, CLEC5A (P < 0.030), lysozyme (LYZ; P < 0.047) and TREM1 (P < 0.023). Six genes were found to be expressed in a NOD2-dependent manner (CD101, P < 0.002; CLEC5A, P < 0.020; CXCL5, P < 0.009; IL-24, P < 0.044; ITGB2, P < 0.041; LYZ, P < 0.042). Interestingly, the highest transcript levels were observed in patients with heterozygous NOD2 mutations.

CONCLUSION

Our data identify CLEC5A and LYZ as CD- and NOD2-associated genes of PBMCs and encourage further studies on their pathomechanistic roles.

Key Words: Peripheral blood mononuclear cells, Gene expression, NOD2, Lysozyme, Crohn’s disease, CLEC5A

Core tip: Peripheral blood mononuclear cells (PBMCs) are a useful tool to study peculiarities of the immune response in the context of Crohn’s disease (CD). Here, we investigated whether PBMCs from patients with CD, even at the stage of clinical remission, exhibit altered gene expression profiles after challenge with pathogen-associated molecular patterns and vitamin D. For TREM1, lysozyme and CLEC5A, disease-associated expression patterns, with higher transcript levels in patient-derived PBMCs, were observed. The two latter genes, along with four other transcripts, also showed NOD2-dependent expression profiles. TREM1 and CLEC5A may act with NOD2 in a regulatory network with a pathophysiological role in CD.



INTRODUCTION

Inflammatory bowel diseases (IBD) are chronic intestinal disorders and mainly consist of the two entities Crohn’s disease (CD) and ulcerative colitis (UC)[1,2]. The clinical course of IBD is characterized by intermittent periods of relapses and remission, which are unpredictable in clinical practice. The pathogenesis of IBD is multifactorial, including genetic and environmental factors, and involves an inappropriate activation of the mucosal immune system, which is triggered by the intestinal microbiota in genetically predisposed individuals[1-5]. In Caucasian populations, nucleotide-binding oligomerization domain 2 (NOD2) has emerged as one of the main susceptibility genes for CD[6-8]. NOD2 is an intracellular pattern recognition receptor sensing muramyl dipeptide (MDP)[9,10], a fragment of peptidoglycan (PGN), but also PGN by itself[11,12] and, upon ligand binding, induces activation of the transcription factor NF-κB[13]. However, NOD2 activation via PGN is dependent on a TLR2 co-stimulatory signal[14].

In addition, the environment, e.g., vitamin D deficiency, also affects the development and clinical course of IBD[15-17]. Vitamin D deficiency has a high prevalence in IBD patients[17,18]. We have recently shown that clinical factors, e.g., the use of tumor necrosis factor (TNF)-α inhibitor, are associated with significant changes in vitamin D levels[19]. Vitamin D was originally mainly implicated in bone health, regulating calcium and phosphate metabolism[20,21], but recent evidence has shown that vitamin D also profoundly impacts the innate and adaptive immune system[22-24]. Underscoring its role in the pathogenesis of CD, vitamin D was shown to be an inductor of NOD2 gene expression[25]. Using peripheral blood mononuclear cells (PBMCs) and dendritic cells, Dionne et al[26] showed that 1, 25-vitamin D acts as a modulator of the innate immune system. However, little is known about the effects of vitamin D and the presence of NOD2 mutations on different gene expression levels in CD. The aim of our study was therefore to further characterize different gene expression profiles in CD patients and healthy controls correlating to NOD2 mutation status and vitamin D pretreatment. We identified different genes associated with the presence of CD and mutations in the NOD2 gene. Follow-up studies on these genes may provide novel insights into the pathogenesis of CD and could contribute to the establishment of biomarkers to better predict the clinical course of the disease.

MATERIALS AND METHODS
Patients and controls

Sixteen patients with CD were recruited from the Rostock University Medical Center. The disease activity was determined via the Crohn’s disease activity index (CDAI)[27]. Furthermore, all patients were classified according to the Montreal classification[28], and age, gender and disease-specific medication were recorded. Six healthy volunteers without immune-mediated gastrointestinal or other autoimmune disorders served as controls. EDTA blood samples were drawn from all participants for genotyping studies and isolation of PBMCs. Plasma levels of vitamin D and C-reactive protein (CRP) were determined using routine laboratory methods.

The study was approved by the ethics board of the University of Rostock (A-2015-0042). Written informed consent was obtained from each participant prior to enrollment.

Isolation, culture and treatment of PBMCs

PBMCs were isolated from EDTA venous blood using density-gradient centrifugation over Pancoll (PAN-Biotech, Aidenbach, Germany). Immediately after isolation, PBMCs were resuspended in cryopreservation medium [fetal calf serum (FCS) supplemented with 10% dimethyl sulfoxide (DMSO)] and stored at -150 °C until required. After thawing, the cells were cultured in RPMI-1640 medium supplemented with 10% FCS and 1% penicillin/streptomycin (all reagents from Biochrom/Merck, Berlin, Germany), and exposed to 1α,25-dihydroxyvitamin D3 (Santa Cruz Biotechnology, Dallas, TX, United States) at 40 nmol/L as indicated. After an incubation period of 20 h at 37 °C in a 5% CO2 humidified atmosphere, lipopolysaccharide (LPS; 1 μg/mL; Sigma-Aldrich, Deisenhofen, Germany) and peptidoglycan (PGN; 10 μg/mL; Sigma-Aldrich) were added to the cells as indicated, and incubation continued for another 6 h (Figure 1). Subsequently, the cells were lysed in RTL Plus buffer, which was included in the RNeasy Plus Kit (Qiagen, Hilden, Germany), and subjected to RNA isolation (see below).

Figure 1
Figure 1 Protocol of peripheral blood mononuclear cells treatment.
NOD2 genotyping

DNA was isolated from whole blood using a QIAamp DNA Blood Mini Kit (Qiagen) according to the manufacturer’s protocol. All patients and controls were genotyped with respect to the three major mutations in the NOD2 gene (SNP 8; R702W, NCBI reference SNP ID: rs2066844, SNP 12; G908R, NCBI reference SNP ID: rs2066845 and SNP 13; 1007fs, NCBI reference SNP ID: rs2066847). The corresponding regions of the NOD2 gene were amplified by PCR using a Taq PCR Master Mix Kit (Qiagen) and primers as specified in Table 1. The following PCR conditions were used: 5 min, 94 °C; 1 min, 94/60/72 °C (45 cycles); 7 min, 72 °C; 4 °C. After Sanger sequencing (Seqlab, Göttingen, Germany), the data were analyzed using the software Chromas, version 2.6. Individuals with no SNP mutations were considered wild-type (WT) for NOD2.

Table 1 Primer for NOD2-genotyping.
SNPPrimer
8Forward: 5’CCTCTTCAATTGTGGCAGGC-3’
Reverse: 5’-CTCCTGCATCTCGTACAGGC-3’
12Forward: 5’-ATGGAGGCAGGTCCACTTTG-3’
Reverse: 5’-TTACCTGAGCCACCTCAAGC-3’
13Forward: 5’-GATGGTACTGAGCCTTTGTTGA-3’
Reverse: 5’-CAGACTTCCAGGATGGTGTCAT -3’
Microarray analysis of RNA expression profiles

RNA was extracted employing an RNeasy Plus Kit according to the manufacturer’s protocol. Total RNA samples were quantified with a spectrophotometer (NanoDrop 1000, Thermo Fisher Scientific, Waltham, MA, United States), and their integrity was confirmed using the Agilent Bioanalyzer 2100 with an RNA Nano chip kit (both from Agilent Technologies, Waldbronn, Germany).

Expression profiling was performed using 200 ng RNA and the Affymetrix Human Clariom S Assay (Affymetrix/Thermo Fisher Scientific), which interrogates over 20000 well-annotated genes. Therefore, the so-called Whole Transcriptome protocol was employed. T7 promoter tags were introduced into all RNA molecules by using N6 3′-ends for DNA strand synthesis, before RNA strand replacement according to Eberwine[29] was conducted. Non-labeled aRNA was produced by in vitro transcription. All RNA molecules were amplified in a linear manner, avoiding a 3′ bias. Using purified aRNA as a template, a new strand-identical single-strand DNA was produced by adding random primers and dNTPs (including dUTP, which replaced a limited amount of dTTP). After digestion with RNase H, endpoint fragmentation was performed with uracil-DNA-glycosylase in combination with apurinic/apyrimidinic endonuclease 1, and biotinylated dNTPs were added to the 3′-ends of the single-stranded DNA fragments with deoxynucleotidyl transferase. Subsequently, hybridization of the microarrays was performed at 45 °C in a GeneChip® Hybridization Oven 645 (Affymetrix/Thermo Fisher Scientific). After overnight incubation, the microarrays were scanned using the GeneChip Scanner 3000 (Affymetrix/Thermo Fisher Scientific) at 0.7 μm resolution.

Primary data analysis was performed with the Affymetrix Transcriptome Analysis Console software version 3.1.0.5 including the Robust Multiarray Average module for normalization. Gene expression data were log-transformed. A change was considered significant when the ANOVA P-value met the criterion P < 0.05 at fold changes >|2|, i.e., expression increments or declines larger than two. Along with the publication of the manuscript, our complete microarray data will be available in the Gene Expression Omnibus database (GEO accession number: GSE110186).

Quantitative reverse transcriptase-PCR using real-time TaqManTM technology

Unless indicated otherwise, reagents from Thermo Fisher Scientific were used in all subsequent steps. Cellular RNA prepared as described above was treated with a DNA-free kit to remove traces of genomic DNA, and 250 ng of RNA per sample was reverse transcribed into cDNA using TaqManTM Reverse Transcription Reagents and random priming. Using a ViiA 7 sequence detection system (Thermo Fisher Scientific), target cDNA levels were quantified by real-time PCR. Therefore, qPCR MasterMix (Eurogentec, Seraing, Liège, Belgium) and the following human-specific TaqManTM gene expression assays with fluorescently labeled MGB probes were used: Hs00355476_m1 (CCL20), Hs00188627_m1 (CD101), Hs00370621_m1 (CLEC12A), Hs04398399_m1 (CLEC5A), Hs01902549_s1 (CLEC7A), Hs01099660_g1 (CXCL5), Hs01114274_m1 (IL24), Hs00167304_m1 (ITGAM), Hs00164957_m1 (ITGB2), Hs00426232_m1 (LYZ), Hs00234007_m1 (MSR1), Hs01065279_m1 (PECAM1), Hs00218624_m1 (TREM1), and Hs99999905_m1 (GAPDH). PCR conditions were as follows: 95 °C for 10 min, followed by 40 cycles of 15 s at 95 °C/ 1 min at 60 °C. Relative amounts of target mRNA in PBMCs were expressed as 2-(∆Ct) values.

Statistical analysis

Real-time PCR data were analyzed with repeated-measures ANOVA. Mean group differences were compared for “disease” (patients with CD vs controls) and “NOD2-status” (WT, heterozygote, homozygote), as well as for (within-subject factors) “vitamin D application” (yes vs no) and “stimulation” (LPS, PGN, LPS + PGN, controls), employing a general linear model for repeated measurements. Age was considered a covariate in the disease model because disease groups were not balanced by age, and NOD2 groups were tested post hoc by LSD. Normal distribution of measurements was assessed using the Kolmogorov-Smirnov test. P < 0.05 was considered statistically significant. All data were processed using IBM® SPSS® Advanced Statistics 22.0.

RESULTS

PBMCs provide an easily accessible tool to investigate disease-associated peculiarities of the antipathogenic immune response of patients with CD. To study transcripts in an unbiased manner, we initially chose a microarray approach. Therefore, PBMCs from healthy individuals and patients with CD in remission (n = 3 each; all of them NOD2-WT) were pretreated with vitamin D3 for 20 h before they were challenged simultaneously with LPS and PGN for 6 h. Subsequently, global gene expression was analyzed employing Clariom S assays, and data were compared with those of untreated controls. Table 2 gives an overview of the significant differences between patients with CD and controls under identical conditions of PBMC treatment.

Table 2 Numbers and maximum changes of up- and downregulated genes in peripheral blood mononuclear cells from Crohn’s disease patient vs identically treated controls.
Treatment of PBMCsUpregulated genesDownregulated genes
Untreated85 (59-fold)39 (39-fold)
Vitamin D325 (21-fold)12 (9-fold)
LPS/PGN54 (6-fold)15 (5-fold)
Vitamin D3 + LPS/PGN29 (15-fold)8 (5-fold)

Under basal conditions and any treatment regimen, genes upregulated in patients with CD exceeded downregulated genes both in number and maximum change. Complete lists of the 267 genes are presented as Supplementary Table 1.

Many of the differentially expressed genes are well-known modulators of immune cell functions in the context of innate and adaptive immunity, and unsurprisingly, some of them have previously been implicated in the pathogenesis of CD, including several immune cell receptors, cytokines/chemokines and their cognate receptors and the antimicrobial peptide lysozyme[30-36]. The latter transcript was found to be upregulated in PBMCs of patients with CD in response to LPS/PGN treatment, independent of the presence or absence of vitamin D3. Intriguingly, expression of various genes was synchronously up- or downregulated under different conditions, suggesting a robustness of the expression profile against external perturbations.

For in-depth analysis, we selected a panel of 11 genes from the list of candidates shown in Table 2 that fulfilled the following criteria: (1) differential expression in PBMCs from patients with CD and controls under basal conditions and/or under at least two treatment regimens (vitamin D3, LPS+PGN and their combination, respectively) and (2) an established or potential role in inflammation and/or regulation of the immune response. Table 3 shows details regarding all selected genes as well as a twelfth gene, TREM1, that was included as a control as an established vitamin D-responsive gene with immunomodulatory function[37]. Interestingly, three pro-inflammatory mediators, CCL20, CXCL5 and IL-24, displayed lower expression levels in patients with CD, which might be a consequence of their disease-specific medication (see below).

Table 3 Genes selected for real-time PCR studies.
TranscriptFold changes: patients vs controls
Details on function/ reasons to study
Basal+D, -L/P-D, +L/P+D, +L/P
MSR19.5613.1914.72macrophage scavenger receptor[49]; differentially expressed in 3 of 4 groups
CD1012.66expressed on various immune cells; inhibits expansion of colitogenic T cells[30]
CLEC5A2.823.11C-type lectin member 5A, pattern recognition receptor; involved in antibacterial/antiviral defense[43]
CLEC7A4.533.94C-type lectin member 7A, pattern recognition receptor; control of fungal infections[50]
CLEC12A6.044.52C-type lectin member 12A, pattern recognition receptor, inhibits cell death-induced inflammation[51]
ITGAM2.18CD11b; integrin αM; expressed by many immune cells; polymorphisms linked to autoimmunity[52]
LYZ3.512.03antimicrobial enzyme; essential role in innate immunity; increased production linked to CD[31]
PECAM13.15CD31; implicated in transendothelial leukocyte migration in experimental colitis[32]
CCL20-2.33-5.08chemokine expressed by neutrophils, enterocytes, Bcells and dendritic cell; IBD predilection gene[33]
CXCL5-38.89-5.32regulates neutrophil homeostasis and chemotaxis; increased serum levels in IBD patients reported[34]
IL-24-3.49-4.82Involved in host defence against bacteria and fungi; increased expression in patients with active IBD[35]
TREM1amplifier of antimicrobial immune responses and inflammation in experimental colitis and IBD[36]

The expression profiles of the selected genes were subsequently studied by real-time PCR. In addition to WT-NOD2 patients and healthy controls (n = 6 each, including the samples previously analyzed by microarray technology), we also included patients with heterozygous and homozygous mutations of NOD2 (n = 5 each). Furthermore, we refined the protocol of PBMC treatment using LPS and PGN both in combination and as individual factors (Supplementary Table 2).

The clinical characteristics, laboratory findings and the medication of all 16 patients are shown in Table 4. Except for one person with a slightly increased CDAI of 166, all patients presented with a CDAI of < 150, indicating disease remission[27]. The CRP-values of 13 patients were in the normal range (below 5 mg/L). In the remaining three patients, modestly elevated CRP-values (all below 14 mg/L) were detected. Disease activity in all patients could still be considered low. All but two patients presented with vitamin D levels below 75 nmol/L, suggesting an insufficiency or even deficiency (levels below 50 nmol/L). This finding was not unexpected, as all of the samples were collected during the European winter season. As a consequence, a vitamin D substitution therapy was initiated, if appropriate. Three of the patients were on steroids (> 10 mg prednisolone/d) at the time of the study, 7 received azathioprine, and 12 were treated with anti-TFN-α antibodies. Healthy controls consisted of 3 males and 3 females with an age range from 25 to 53 years.

Table 4 Characteristics of the patients -nucleotide-binding oligomerization domain 2 status, classification and activity of the disease, C-reactive protein and vitamin D levels, and medication at the time of the study.
No.SexAgeNOD2Montreal classificationCDAICRP (mg/L)Vitamin D (nmol/L)Prednisolon (> 10 mg/d)Azathio-prineAnti-TNF-α
1M24WT1A2 L3 L4 B11212.58104.0YesNoYes
2M28WT1A2 L3 B3p46< 1.062.9NoNoYes
3M64WT1A3 L3 B2p1114.5827.0YesNoYes
4F60WTA2 L3 B21002.2772.5NoNoYes
5F62WTA3 L3 B3p82< 1.058.2YesYesYes
6M46WTA2 L3 L4 B2p1104.9732.9NoNoYes
7M38HOA2 L3 B3p9213.3040.2NoYesNo
8M64HOA3 L3 B134< 1.053.0NoNoYes
9F48HO2A2 L3 B254< 1.067.6NoNoYes
10M54HOA2 L3 B2p1031.6351.9NoYesNo
11M26HOA1 L3 B2p1207.8631.0NoYesYes
12M27HTA2 L3 B11061.4147.2NoNoYes
13M37HTA2 L3 B2p1664.5637.9NoYesNo
14F48HTA2 L1 B3p1151.0155.2NoYesYes
15M57HTA1 L3 B2p13212.3052.3NoNoYes
16M47HTA2 L3 B3602.6292.0NoYesNo

For statistical data analysis, a general linear model repeated measure was chosen to assess mean differences between groups. Comparing controls to CD patients, no significant differences were detected when samples with and without vitamin D incubation were considered conjointly, due to the strong overlaying effect of vitamin D. However, focusing on the subgroup of vitamin D treated samples, disease-related differences were observed (no. 5-8 in Supplementary Table 2). Here, with age adjustment, three genes displayed a disease-dependent expression pattern (Table 5): consistent with the microarray data, significantly higher CLEC5A and LYZ transcript levels were observed in PBMCs of patients with CD. The same finding was observed for TREM1, which was included due to its vitamin D-dependent expression pattern but not because of the microarray results. Together, these findings suggest that CD-associated changes in gene expression could be attributed to treatment protocols that include vitamin D, as analyses in the remaining subgroup (w/o vitamin D) did not show any differences (all P > 0.20).

Table 5 differentially expressed transcripts in vitamin D-pretreated peripheral blood mononuclear cells from Crohn’s disease patients and controls (with and without additional stimulation with lipopolysaccharide and peptidoglycan, respectively, age adjusted).
GeneP valueUpregulated in
CLEC5A0.030CD patients
LYZ0.047CD patients
TREM10.023CD patients

We also analyzed the influence of NOD2 mutations on the expression of the gene panel described above. Considering all treated and untreated samples (Supplementary Table 2), a significant effect of the NOD2 status was observed for 5 of these genes, including CLEC5A and LYZ, and one additional gene from Table 2, integrin subunit beta 2 (ITGB2) (Table 6). With a P-value of 0.053, TREM1 just missed statistical significance.

Table 6 Transcripts with a NOD2-dependent expression pattern.
GeneP valueHighest levels
CD1010.002Heterozygotes
CLEC5A0.020Heterozygotes
CXCL50.009Heterozygotes
IL-240.044Heterozygotes
ITGB20.041Heterozygotes
LYZ0.042Heterozygotes

Unexpectedly, heterozygous CD patients displayed the highest expression levels for all of the genes, whereas no statistically significant differences between persons with WT-NOD2 (patients and controls) and homozygous NOD2 mutations were detected. The phenomenon is apparently unrelated to medication, which was very similar in the groups of patients with heterozygous and homozygous NOD2 mutations (Table 4). This conclusion is also supported by statistical evaluations, which did not show any significant association between treatment with prednisolone, azathioprine or anti-TNF-α and expression of CLEC5A, LYZ and TREM1 in PBMCs of patients with CD.

DISCUSSION

Many studies have shown that numerous risk genes of CD code for molecules involved in host defense against pathogens, such as nucleotide-binding oligomerization domain 2 (NOD2), ATG16L1, and those implicated in the T helper type 17 (Th17) pathway[38-43]. Here, we tested the hypothesis that PBMCs of patients with CD, even at the stage of clinical remission, exhibit an altered gene expression profile upon challenge with pathogen-associated molecular patterns (PAMPs) and/or the immunomodulatory hormone vitamin D, which has previously been shown to exert differential effects on the expression of NOD2- and TLR-induced cytokines in the context of CD[26].

Initial microarray experiments identified more than 200 genes with different expression patterns among patients with CD and controls. Based on predefined expression criteria, genes with roles in inflammation and immunity were selected for in-depth analysis by real-time PCR. A disease-associated expression pattern was identified for CLEC5A, lysozyme and TREM1. Six genes, including CLEC5A and lysozyme, displayed a NOD2-dependent expression pattern. With respect to lysozyme and TREM1, our findings are consistent with previous reports, which found that increased levels of both proteins in serum were implicated in the pathophysiology of IBD[44,45]. To the best of our knowledge, however, this is the first report of an association between CLEC5A expression and CD. CLEC5A has most recently been identified as an important receptor in innate immunity by neutrophil trap formation and secretion of different proinflammatory cytokines after stimulation with Listeria monocytogenes[43]. This finding is especially interesting, as defective bacterial clearance was shown to play a crucial role in the pathogenesis of CD[46,47]. Of note, both CLEC5A and TREM1 proteins can be linked to the product of the best-established CD risk gene, NOD2, by the STRING database[48] (Figure 2).

Figure 2
Figure 2 Network analysis using the STRING database[48]. The network was derived employing human TREM1 as the search term (https://string-db.org/cgi/network.pl?taskId=PmXpOD7RMwaM).

In conclusion, we found that PBMCs of patients with CD display alterations in their response to vitamin D and PAMPs. Disease-associated and NOD2-dependent gene expression profiles are preserved even at the stage of clinical remission. Our data identify CLEC5A, LYZ and TREM1 as genes of particular interest for follow-up studies. We hypothesize that these genes may act in a common network relevant to CD pathogenesis. Establishment of biomarkers to better predict the clinical course of the disease remains a long-term goal of our studies.

ARTICLE HIGHLIGHTS
Research background

In Crohn’s disease (CD), the interplay of genetic and environmental factors converges at the level of an altered antipathogenic immune response, which is incompletely understood. Peripheral blood mononuclear cells (PBMCs) provide a useful tool to study elements of the immunopathogenesis of the disease in vitro.

Research motivation

Currently, there is a lack of biomarkers to predict the clinical course of CD. Furthermore, the development of specific therapies would benefit from an improved mechanistic understanding of the pathogenesis of the disease.

Research objectives

The aim of this study was to identify disease-specific gene expression profiles of PBMCs from patients with CD in clinical remission. Specifically, we were interested in alterations of the gene expression profile after challenging PBMCs with pathogen-associated molecular patterns (PAMPs) and the immunomodulatory hormone vitamin D.

Research methods

PBMCs from patients with CD and healthy donors were cultured with vitamin D, peptidoglycan (PGN) and lipopolysaccharide (LPS), before RNA was isolated and subjected to microarray analysis and quantitative real-time PCR. Disease-specific gene expression profiles were evaluated by general linear model repeated measure analysis, paying particular attention to the well-established CD risk gene NOD2.

Research results

Microarray experiments yielded a total of 267 genes that were significantly up- or downregulated in PBMCs of patients with CD, compared to healthy donors, after challenge with vitamin D and/or a combination of LPS and PGN. For further analysis by real-time PCR, genes with roles in inflammation and immunity were selected. For three of these genes, CLEC5A, lysozyme and TREM1, a disease-associated expression pattern was validated. Six genes, including CLEC5A and lysozyme, were found to be expressed in a NOD2-dependent manner.

Research conclusions

PBMCs of patients with CD display alterations of their response to vitamin D and PAMPs that are preserved even at the stage of clinical remission. CLEC5A, TREM1 and NOD2 may act in a common network relevant to CD pathogenesis.

Research perspectives

Follow-up studies on alterations of the antipathogenic immune response may provide novel insights into the pathogenesis of CD and may also help to establish biomarkers to better predict the clinical course of the disease.

ACKNOWLEDGMENTS

We thank Mrs. Katja Bergmann for expert technical assistance.

Footnotes

Manuscript source: Unsolicited manuscript

Specialty type: Gastroenterology and hepatology

Country of origin: Germany

Peer-review report classification

Grade A (Excellent): A

Grade B (Very good): B

Grade C (Good): 0

Grade D (Fair): D

Grade E (Poor): 0

P- Reviewer: Day AS, Liu F, Macedo G S- Editor: Wang XJ L- Editor: A E- Editor: Huang Y

References
1.  Abraham C, Cho JH. Inflammatory bowel disease. N Engl J Med. 2009;361:2066-2078.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1967]  [Cited by in F6Publishing: 2038]  [Article Influence: 135.9]  [Reference Citation Analysis (5)]
2.  Baumgart DC, Sandborn WJ. Crohn‘s disease. Lancet. 2012;380:1590-1605.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1347]  [Cited by in F6Publishing: 1376]  [Article Influence: 114.7]  [Reference Citation Analysis (0)]
3.  Mayer L. Evolving paradigms in the pathogenesis of IBD. J Gastroenterol. 2010;45:9-16.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 121]  [Cited by in F6Publishing: 125]  [Article Influence: 8.9]  [Reference Citation Analysis (0)]
4.  Sartor RB. Microbial influences in inflammatory bowel diseases. Gastroenterology. 2008;134:577-594.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1339]  [Cited by in F6Publishing: 1319]  [Article Influence: 82.4]  [Reference Citation Analysis (0)]
5.  Naser SA, Arce M, Khaja A, Fernandez M, Naser N, Elwasila S, Thanigachalam S. Role of ATG16L, NOD2 and IL23R in Crohn‘s disease pathogenesis. World J Gastroenterol. 2012;18:412-424.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in CrossRef: 62]  [Cited by in F6Publishing: 71]  [Article Influence: 5.9]  [Reference Citation Analysis (1)]
6.  Hugot JP, Chamaillard M, Zouali H, Lesage S, Cézard JP, Belaiche J, Almer S, Tysk C, O‘Morain CA, Gassull M. Association of NOD2 leucine-rich repeat variants with susceptibility to Crohn‘s disease. Nature. 2001;411:599-603.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3986]  [Cited by in F6Publishing: 3811]  [Article Influence: 165.7]  [Reference Citation Analysis (0)]
7.  Ogura Y, Bonen DK, Inohara N, Nicolae DL, Chen FF, Ramos R, Britton H, Moran T, Karaliuskas R, Duerr RH. A frameshift mutation in NOD2 associated with susceptibility to Crohn‘s disease. Nature. 2001;411:603-606.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3555]  [Cited by in F6Publishing: 3393]  [Article Influence: 147.5]  [Reference Citation Analysis (0)]
8.  Cho JH. The genetics and immunopathogenesis of inflammatory bowel disease. Nat Rev Immunol. 2008;8:458-466.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 666]  [Cited by in F6Publishing: 663]  [Article Influence: 41.4]  [Reference Citation Analysis (0)]
9.  Girardin SE, Boneca IG, Viala J, Chamaillard M, Labigne A, Thomas G, Philpott DJ, Sansonetti PJ. Nod2 is a general sensor of peptidoglycan through muramyl dipeptide (MDP) detection. J Biol Chem. 2003;278:8869-8872.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1755]  [Cited by in F6Publishing: 1701]  [Article Influence: 81.0]  [Reference Citation Analysis (0)]
10.  Inohara N, Ogura Y, Fontalba A, Gutierrez O, Pons F, Crespo J, Fukase K, Inamura S, Kusumoto S, Hashimoto M. Host recognition of bacterial muramyl dipeptide mediated through NOD2. Implications for Crohn‘s disease. J Biol Chem. 2003;278:5509-5512.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1255]  [Cited by in F6Publishing: 1219]  [Article Influence: 58.0]  [Reference Citation Analysis (0)]
11.  Natsuka M, Uehara A, Yang S, Echigo S, Takada H. A polymer-type water-soluble peptidoglycan exhibited both Toll-like receptor 2- and NOD2-agonistic activities, resulting in synergistic activation of human monocytic cells. Innate Immun. 2008;14:298-308.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 24]  [Cited by in F6Publishing: 31]  [Article Influence: 2.1]  [Reference Citation Analysis (0)]
12.  Iyer JK, Coggeshall KM. Cutting edge: primary innate immune cells respond efficiently to polymeric peptidoglycan, but not to peptidoglycan monomers. J Immunol. 2011;186:3841-3845.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 34]  [Cited by in F6Publishing: 34]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
13.  Strober W, Murray PJ, Kitani A, Watanabe T. Signalling pathways and molecular interactions of NOD1 and NOD2. Nat Rev Immunol. 2006;6:9-20.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 584]  [Cited by in F6Publishing: 573]  [Article Influence: 31.8]  [Reference Citation Analysis (0)]
14.  Schäffler H, Demircioglu DD, Kühner D, Menz S, Bender A, Autenrieth IB, Bodammer P, Lamprecht G, Götz F, Frick JS. NOD2 stimulation by Staphylococcus aureus-derived peptidoglycan is boosted by Toll-like receptor 2 costimulation with lipoproteins in dendritic cells. Infect Immun. 2014;82:4681-4688.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 33]  [Cited by in F6Publishing: 34]  [Article Influence: 3.4]  [Reference Citation Analysis (0)]
15.  Rebouças PC, Netinho JG, Cunrath GS, Ronchi LS, de Melo MM, Gonçalves Filho Fde A, Muniz RC, Martins AT, de Oliveira RA, Costa Junior RM. Association between vitamin D serum levels and disease activity markers in patients with Crohn‘s Disease. Int J Colorectal Dis. 2016;31:1495-1496.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 4]  [Cited by in F6Publishing: 4]  [Article Influence: 0.5]  [Reference Citation Analysis (0)]
16.  Ananthakrishnan AN, Khalili H, Higuchi LM, Bao Y, Korzenik JR, Giovannucci EL, Richter JM, Fuchs CS, Chan AT. Higher predicted vitamin D status is associated with reduced risk of Crohn‘s disease. Gastroenterology. 2012;142:482-489.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 286]  [Cited by in F6Publishing: 299]  [Article Influence: 24.9]  [Reference Citation Analysis (0)]
17.  Ulitsky A, Ananthakrishnan AN, Naik A, Skaros S, Zadvornova Y, Binion DG, Issa M. Vitamin D deficiency in patients with inflammatory bowel disease: association with disease activity and quality of life. JPEN J Parenter Enteral Nutr. 2011;35:308-316.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 216]  [Cited by in F6Publishing: 228]  [Article Influence: 17.5]  [Reference Citation Analysis (0)]
18.  Leslie WD, Miller N, Rogala L, Bernstein CN. Vitamin D status and bone density in recently diagnosed inflammatory bowel disease: the Manitoba IBD Cohort Study. Am J Gastroenterol. 2008;103:1451-1459.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 102]  [Cited by in F6Publishing: 99]  [Article Influence: 6.2]  [Reference Citation Analysis (0)]
19.  Schäffler H, Schmidt M, Huth A, Reiner J, Glass Ä, Lamprecht G. Clinical factors are associated with vitamin D levels in IBD patients: A retrospective analysis. J Dig Dis. 2018;19:24-32.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 28]  [Cited by in F6Publishing: 28]  [Article Influence: 4.7]  [Reference Citation Analysis (0)]
20.  DeLuca HF. Overview of general physiologic features and functions of vitamin D. Am J Clin Nutr. 2004;80:1689S-1696S.  [PubMed]  [DOI]  [Cited in This Article: ]
21.  Holick MF. Optimal vitamin D status for the prevention and treatment of osteoporosis. Drugs Aging. 2007;24:1017-1029.  [PubMed]  [DOI]  [Cited in This Article: ]
22.  Cantorna MT, Zhu Y, Froicu M, Wittke A. Vitamin D status, 1,25-dihydroxyvitamin D3, and the immune system. Am J Clin Nutr. 2004;80:1717S-1720S.  [PubMed]  [DOI]  [Cited in This Article: ]
23.  Olliver M, Spelmink L, Hiew J, Meyer-Hoffert U, Henriques-Normark B, Bergman P. Immunomodulatory effects of vitamin D on innate and adaptive immune responses to Streptococcus pneumoniae. J Infect Dis. 2013;208:1474-1481.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 61]  [Cited by in F6Publishing: 67]  [Article Influence: 6.1]  [Reference Citation Analysis (0)]
24.  Zhao H, Zhang H, Wu H, Li H, Liu L, Guo J, Li C, Shih DQ, Zhang X. Protective role of 1,25(OH)2 vitamin D3 in the mucosal injury and epithelial barrier disruption in DSS-induced acute colitis in mice. BMC Gastroenterol. 2012;12:57.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 171]  [Cited by in F6Publishing: 215]  [Article Influence: 17.9]  [Reference Citation Analysis (0)]
25.  Wang TT, Dabbas B, Laperriere D, Bitton AJ, Soualhine H, Tavera-Mendoza LE, Dionne S, Servant MJ, Bitton A, Seidman EG. Direct and indirect induction by 1,25-dihydroxyvitamin D3 of the NOD2/CARD15-defensin beta2 innate immune pathway defective in Crohn disease. J Biol Chem. 2010;285:2227-2231.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 269]  [Cited by in F6Publishing: 282]  [Article Influence: 18.8]  [Reference Citation Analysis (0)]
26.  Dionne S, Calderon MR, White JH, Memari B, Elimrani I, Adelson B, Piccirillo C, Seidman EG. Differential effect of vitamin D on NOD2- and TLR-induced cytokines in Crohn‘s disease. Mucosal Immunol. 2014;7:1405-1415.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 18]  [Cited by in F6Publishing: 19]  [Article Influence: 1.9]  [Reference Citation Analysis (0)]
27.  Best WR, Becktel JM, Singleton JW, Kern F Jr. Development of a Crohn‘s disease activity index. National Cooperative Crohn‘s Disease Study. Gastroenterology. 1976;70:439-444.  [PubMed]  [DOI]  [Cited in This Article: ]
28.  Silverberg MS, Satsangi J, Ahmad T, Arnott ID, Bernstein CN, Brant SR, Caprilli R, Colombel JF, Gasche C, Geboes K. Toward an integrated clinical, molecular and serological classification of inflammatory bowel disease: report of a Working Party of the 2005 Montreal World Congress of Gastroenterology. Can J Gastroenterol. 2005;19 Suppl A:5A-36A.  [PubMed]  [DOI]  [Cited in This Article: ]
29.  Van Gelder RN, von Zastrow ME, Yool A, Dement WC, Barchas JD, Eberwine JH. Amplified RNA synthesized from limited quantities of heterogeneous cDNA. Proc Natl Acad Sci USA. 1990;87:1663-1667.  [PubMed]  [DOI]  [Cited in This Article: ]
30.  Schey R, Dornhoff H, Baier JL, Purtak M, Opoka R, Koller AK, Atreya R, Rau TT, Daniel C, Amann K. CD101 inhibits the expansion of colitogenic T cells. Mucosal Immunol. 2016;9:1205-1217.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 22]  [Cited by in F6Publishing: 24]  [Article Influence: 3.0]  [Reference Citation Analysis (0)]
31.  Rubio CA. Increased Production of Lysozyme Associated with Bacterial Proliferation in Barrett‘s Esophagitis, Chronic Gastritis, Gluten-induced Atrophic Duodenitis (Celiac Disease), Lymphocytic Colitis, Collagenous Colitis, Ulcerative Colitis and Crohn‘s Colitis. Anticancer Res. 2015;35:6365-6372.  [PubMed]  [DOI]  [Cited in This Article: ]
32.  Rijcken E, Mennigen RB, Schaefer SD, Laukoetter MG, Anthoni C, Spiegel HU, Bruewer M, Senninger N, Krieglstein CF. PECAM-1 (CD 31) mediates transendothelial leukocyte migration in experimental colitis. Am J Physiol Gastrointest Liver Physiol. 2007;293:G446-G452.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 23]  [Cited by in F6Publishing: 26]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
33.  Liu JZ, van Sommeren S, Huang H, Ng SC, Alberts R, Takahashi A, Ripke S, Lee JC, Jostins L, Shah T. Association analyses identify 38 susceptibility loci for inflammatory bowel disease and highlight shared genetic risk across populations. Nat Genet. 2015;47:979-986.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1401]  [Cited by in F6Publishing: 1540]  [Article Influence: 171.1]  [Reference Citation Analysis (0)]
34.  Singh UP, Singh NP, Murphy EA, Price RL, Fayad R, Nagarkatti M, Nagarkatti PS. Chemokine and cytokine levels in inflammatory bowel disease patients. Cytokine. 2016;77:44-49.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 162]  [Cited by in F6Publishing: 195]  [Article Influence: 21.7]  [Reference Citation Analysis (0)]
35.  Fonseca-Camarillo G, Furuzawa-Carballeda J, Granados J, Yamamoto-Furusho JK. Expression of interleukin (IL)-19 and IL-24 in inflammatory bowel disease patients: a cross-sectional study. Clin Exp Immunol. 2014;177:64-75.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 50]  [Cited by in F6Publishing: 52]  [Article Influence: 5.2]  [Reference Citation Analysis (0)]
36.  Schenk M, Bouchon A, Seibold F, Mueller C. TREM-1--expressing intestinal macrophages crucially amplify chronic inflammation in experimental colitis and inflammatory bowel diseases. J Clin Invest. 2007;117:3097-3106.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 256]  [Cited by in F6Publishing: 275]  [Article Influence: 16.2]  [Reference Citation Analysis (0)]
37.  Kim TH, Lee B, Kwon E, Choi SJ, Lee YH, Song GG, Sohn J, Ji JD. Regulation of TREM-1 expression by 1,25-dihydroxyvitamin D3 in human monocytes/macrophages. Immunol Lett. 2013;154:80-85.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 23]  [Cited by in F6Publishing: 26]  [Article Influence: 2.4]  [Reference Citation Analysis (0)]
38.  Saleh M, Elson CO. Experimental inflammatory bowel disease: insights into the host-microbiota dialog. Immunity. 2011;34:293-302.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 117]  [Cited by in F6Publishing: 118]  [Article Influence: 9.1]  [Reference Citation Analysis (0)]
39.  Rai E, Wakeland EK. Genetic predisposition to autoimmunity--what have we learned? Semin Immunol. 2011;23:67-83.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 51]  [Cited by in F6Publishing: 57]  [Article Influence: 4.4]  [Reference Citation Analysis (0)]
40.  Jostins L, Ripke S, Weersma RK, Duerr RH, McGovern DP, Hui KY, Lee JC, Schumm LP, Sharma Y, Anderson CA, Essers J, Mitrovic M, Ning K, Cleynen I, Theatre E, Spain SL, Raychaudhuri S, Goyette P, Wei Z, Abraham C, Achkar JP, Ahmad T, Amininejad L, Ananthakrishnan AN, Andersen V, Andrews JM, Baidoo L, Balschun T, Bampton PA, Bitton A, Boucher G, Brand S, Büning C, Cohain A, Cichon S, D‘Amato M, De Jong D, Devaney KL, Dubinsky M, Edwards C, Ellinghaus D, Ferguson LR, Franchimont D, Fransen K, Gearry R, Georges M, Gieger C, Glas J, Haritunians T, Hart A, Hawkey C, Hedl M, Hu X, Karlsen TH, Kupcinskas L, Kugathasan S, Latiano A, Laukens D, Lawrance IC, Lees CW, Louis E, Mahy G, Mansfield J, Morgan AR, Mowat C, Newman W, Palmieri O, Ponsioen CY, Potocnik U, Prescott NJ, Regueiro M, Rotter JI, Russell RK, Sanderson JD, Sans M, Satsangi J, Schreiber S, Simms LA, Sventoraityte J, Targan SR, Taylor KD, Tremelling M, Verspaget HW, De Vos M, Wijmenga C, Wilson DC, Winkelmann J, Xavier RJ, Zeissig S, Zhang B, Zhang CK, Zhao H; International IBD Genetics Consortium (IIBDGC), Silverberg MS, Annese V, Hakonarson H, Brant SR, Radford-Smith G, Mathew CG, Rioux JD, Schadt EE, Daly MJ, Franke A, Parkes M, Vermeire S, Barrett JC, Cho JH. Host-microbe interactions have shaped the genetic architecture of inflammatory bowel disease. Nature. 2012;491:119-124.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3465]  [Cited by in F6Publishing: 3322]  [Article Influence: 276.8]  [Reference Citation Analysis (0)]
41.  Franke A, McGovern DP, Barrett JC, Wang K, Radford-Smith GL, Ahmad T, Lees CW, Balschun T, Lee J, Roberts R. Genome-wide meta-analysis increases to 71 the number of confirmed Crohn‘s disease susceptibility loci. Nat Genet. 2010;42:1118-1125.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1894]  [Cited by in F6Publishing: 1924]  [Article Influence: 137.4]  [Reference Citation Analysis (0)]
42.  Khor B, Gardet A, Xavier RJ. Genetics and pathogenesis of inflammatory bowel disease. Nature. 2011;474:307-317.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1693]  [Cited by in F6Publishing: 1713]  [Article Influence: 131.8]  [Reference Citation Analysis (1)]
43.  Chen ST, Li FJ, Hsu TY, Liang SM, Yeh YC, Liao WY, Chou TY, Chen NJ, Hsiao M, Yang WB. CLEC5A is a critical receptor in innate immunity against Listeria infection. Nat Commun. 2017;8:299.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 40]  [Cited by in F6Publishing: 56]  [Article Influence: 8.0]  [Reference Citation Analysis (0)]
44.  Park JJ, Cheon JH, Kim BY, Kim DH, Kim ES, Kim TI, Lee KR, Kim WH. Correlation of serum-soluble triggering receptor expressed on myeloid cells-1 with clinical disease activity in inflammatory bowel disease. Dig Dis Sci. 2009;54:1525-1531.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 51]  [Cited by in F6Publishing: 51]  [Article Influence: 3.4]  [Reference Citation Analysis (0)]
45.  Fixa B, Komárková O, Procházková J. Serum lysozyme in inflammatory gastric and enteric diseases and in functional dyspepsia. Scand J Gastroenterol. 1983;18:349-352.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3]  [Cited by in F6Publishing: 3]  [Article Influence: 0.1]  [Reference Citation Analysis (0)]
46.  Cooney R, Baker J, Brain O, Danis B, Pichulik T, Allan P, Ferguson DJ, Campbell BJ, Jewell D, Simmons A. NOD2 stimulation induces autophagy in dendritic cells influencing bacterial handling and antigen presentation. Nat Med. 2010;16:90-97.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 760]  [Cited by in F6Publishing: 766]  [Article Influence: 51.1]  [Reference Citation Analysis (0)]
47.  Travassos LH, Carneiro LA, Ramjeet M, Hussey S, Kim YG, Magalhães JG, Yuan L, Soares F, Chea E, Le Bourhis L. Nod1 and Nod2 direct autophagy by recruiting ATG16L1 to the plasma membrane at the site of bacterial entry. Nat Immunol. 2010;11:55-62.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 916]  [Cited by in F6Publishing: 967]  [Article Influence: 64.5]  [Reference Citation Analysis (0)]
48.  Szklarczyk D, Morris JH, Cook H, Kuhn M, Wyder S, Simonovic M, Santos A, Doncheva NT, Roth A, Bork P. The STRING database in 2017: quality-controlled protein-protein association networks, made broadly accessible. Nucleic Acids Res. 2017;45:D362-D368.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 4345]  [Cited by in F6Publishing: 4684]  [Article Influence: 585.5]  [Reference Citation Analysis (0)]
49.  Platt N, Haworth R, Darley L, Gordon S. The many roles of the class A macrophage scavenger receptor. Int Rev Cytol. 2002;212:1-40.  [PubMed]  [DOI]  [Cited in This Article: ]
50.  Taylor PR, Tsoni SV, Willment JA, Dennehy KM, Rosas M, Findon H, Haynes K, Steele C, Botto M, Gordon S. Dectin-1 is required for beta-glucan recognition and control of fungal infection. Nat Immunol. 2007;8:31-38.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 841]  [Cited by in F6Publishing: 869]  [Article Influence: 48.3]  [Reference Citation Analysis (0)]
51.  Neumann K, Castiñeiras-Vilariño M, Höckendorf U, Hannesschläger N, Lemeer S, Kupka D, Meyermann S, Lech M, Anders HJ, Kuster B. Clec12a is an inhibitory receptor for uric acid crystals that regulates inflammation in response to cell death. Immunity. 2014;40:389-399.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 119]  [Cited by in F6Publishing: 130]  [Article Influence: 13.0]  [Reference Citation Analysis (0)]
52.  Rosetti F, Mayadas TN. The many faces of Mac-1 in autoimmune disease. Immunol Rev. 2016;269:175-193.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 68]  [Cited by in F6Publishing: 84]  [Article Influence: 10.5]  [Reference Citation Analysis (0)]