Skip to main content
Advertisement
  • Loading metrics

A nod to the bond between NOD2 and mycobacteria

Abstract

Mycobacteria are responsible for several human and animal diseases. NOD2 is a pattern recognition receptor that has an important role in mycobacterial recognition. However, the mechanisms by which mutations in NOD2 alter the course of mycobacterial infection remain unclear. Herein, we aimed to review the totality of studies directly addressing the relationship between NOD2 and mycobacteria as a foundation for moving the field forward. NOD2 was linked to mycobacterial infection at 3 levels: (1) genetic, through association with mycobacterial diseases of humans; (2) chemical, through the distinct NOD2 ligand in the mycobacterial cell wall; and (3) immunologic, through heightened NOD2 signaling caused by the unique modification of the NOD2 ligand. The immune response to mycobacteria is shaped by NOD2 signaling, responsible for NF-κB and MAPK activation, and the production of various immune effectors like cytokines and nitric oxide, with some evidence linking this to bacteriologic control. Absence of NOD2 during mycobacterial infection of mice can be detrimental, but the mechanism remains unknown. Conversely, the success of immunization with mycobacteria has been linked to NOD2 signaling and NOD2 has been targeted as an avenue of immunotherapy for diseases even beyond mycobacteria. The mycobacteria–NOD2 interaction remains an important area of study, which may shed light on immune mechanisms in disease.

Introduction

Mycobacterium is a genus of bacteria responsible for epidemiologically important human and veterinary diseases. Tuberculosis (TB) is estimated to have killed over 1 billion people since 1882 when Robert Koch discovered the etiological agent, Mycobacterium tuberculosis (Mtb) [1,2], and TB continues killing over 1 million annually [3]. M. leprae and nontuberculous mycobacteria (NTM) like M. avium and M. abscessus cause suffering to hundreds of thousands of people each year [4]. Additionally, in the veterinary world, bovine TB (due to M. bovis and M. orygis) and bovine paratuberculosis (due to M. avium paratuberculosis, herein abbreviated as Map) present as health and economic challenges, both directly to livestock and due to their risk of zoonotic spread to humans [57]. Reliable vaccines are difficult to design for mycobacteria because the correlates of protection are unknown. To make better interventions, we need to understand the immune response to mycobacteria.

The immune response begins with the recognition of microbe-associated molecular patterns (MAMPs) by host pattern recognition receptors (PRRs). The overall importance of PRRs in mycobacterial infection seems to be limited by the redundancy of many PRRs, yet owing to their distinct signaling mechanisms, individual PRRs may alter the outcome of infection in specific, nonredundant ways with subtle, if not significant, consequences [8]. Two decades ago, a newly described gene called Nucleotide-binding Oligomerization Domain-containing 2 (NOD2, formerly known as Caspase Recruitment Domain 15 or CARD15), initially associated with susceptibility to Crohn’s disease [911] and Blau syndrome [12], was shown to mediate recognition of bacterial peptidoglycan (PGN) [13,14]. Not long thereafter, genetic, chemical, and immunological evidence specifically linking NOD2 to mycobacteria mounted. Despite this, there lacks a consensus on the mechanisms by which NOD2 affects the outcome of mycobacterial infections.

To provide a foundation for moving the topic forward, herein we have combined our knowledge of NOD2 and mycobacteria with a review of the literature directly examining these elements together. We have collected 60 papers experimentally testing a relationship between NOD2 and mycobacteria (Table 1). Reflecting the topics covered in these papers, we discuss NOD2 in the context of mycobacterial infection, including expression, elicited effector functions, and outcomes in animal models, followed by a nod to the potential interventions founded on the NOD2–mycobacteria interaction. While much work remains to understand the mechanisms behind the association between NOD2 and mycobacteria, the evidence supports that this relationship has important implications for the outcome of mycobacterial infection.

thumbnail
Table 1. Research articles with data directly addressing NOD2 and mycobacteriaa.

https://doi.org/10.1371/journal.ppat.1011389.t001

How are mycobacteria and NOD2 linked?

In 2003, NOD2 was demonstrated to be essential for the innate immune response to the common bacterial PGN motif muramyl dipeptide (MDP) [13,14]. NOD2 has 4 domains: The 2 caspase recruitment domains (CARDs) mediate interactions with other proteins; the nucleotide-binding domain (NBD, or NACHT) binds ATP, which was proposed to provide a conformational change that enhances the MDP interaction and oligomerization, while ATP hydrolysis to ADP has the opposite effect [75]; the leucine-rich repeats (LRRs) are thought to compose the ligand (MDP) binding domain [76,77]. A representation of human NOD2 is illustrated in Fig 1 [78,79]. During bacterial infection, PGN is digested in the phagosome by lysozyme to release muropeptides like MDPs [80], which traverse SLC15A transporters [8183] to enter the cytosol where NOD2 resides. NOD2 must also localize to cellular or phagosomal membranes via palmitoylation by ZDHHC5 for MDP recognition [8486]. Recently, the enzyme NAGK was identified to be essential for the response to MDP upstream of NOD2: NAGK phosphorylates MDP as it enters the cytoplasm, and only phospho-MDP can signal through NOD2 [87]. NOD2 binding to phospho-MDP has yet to be demonstrated. Earlier, surface plasmon resonance was used to show binding of unphosphorylated MDP to NOD2 [76,88], and tagged NOD2 was detected from a pull-down with biotin-MDP [75]. It remains to be clarified whether NOD2 binds to the phosphorylated and/or nonphosphorylated form of MDP. We refer the reader to [89,90] for more on NOD2 signaling.

thumbnail
Fig 1. Structure of human NOD2.

These structural representations were generated from UniProt entry Q9HC29 (“NOD2_HUMAN”) folded with AlphaFold. Approximately 180° of rotation separates the images from the top and bottom rows. (A) Ribbon structure of NOD2 with AlphaFold per-residue confidence score (pLDDT) in range from 0 to 100 (low to high confidence). (B) Ribbon and (C) space-filling models of AlphaFold NOD2 illustrated using UCSF ChimeraX based on UniProt domain annotation. NOD2 has 4 domains: 2 CARDs, 1 NACHT/NBD, and 1 LRR domain. In Crohn’s disease, some patients have a frameshift variant where the protein is foreshortened by 34 amino acids. This missing region is shown in mauve. CARD, caspase recruitment domain; LRR, leucine-rich repeat; NBD, nucleotide-binding domain; NOD2, Nucleotide-binding Oligomerization Domain-containing 2.

https://doi.org/10.1371/journal.ppat.1011389.g001

1. The genetic link

Mutations in NOD2 have been strongly associated with Crohn’s disease, especially the common LRR domain mutations 2104CT (R702W), 2722GC (G908R), and 3020insC (1007FS) [911,91,92]. These are recessive loss-of-function mutations (i.e., loss of MDP responsiveness) [13,14]. The 1001GA (R334Q) mutation in the NBD of NOD2 was associated with Blau syndrome [12]. Both Crohn’s disease and Blau syndrome are idiopathic, but the existence of a bacterial contributing agent is indirectly supported, given that NOD2 is a PGN sensor. Because of parallels to the intestinal pathology of Johne’s disease (caused by Map infection in ruminants), Map has and continues to be hypothesized to play a role in Crohn’s disease [93]. Map infection was reported in patients with NOD2 polymorphisms and/or Crohn’s disease [94,95], but a causative role for Map in Crohn’s disease is elusive. Other functions for NOD2 have been reported outside of being solely a PGN sensor, such as a role in viral RNA sensing [96,97] and the endoplasmic reticulum stress response [98101], which do not contradict but provide alternative hypotheses to a bacterial etiology for Crohn’s disease and Blau syndrome and, therefore, must also be considered.

NOD2 interacts with the adaptor protein kinase RIPK2 for signaling [102104], and mutations in both NOD2 and RIPK2 have been associated with leprosy and/or its endophenotypes [105109]. Additionally, NOD2 was recently shown to interact with LRRK2, a kinase genetically linked with Parkinson’s disease [74]. Reports have linked NOD2 to LRRK2-dependent intestinal homeostasis and RIPK2 phosphorylation [110112], and LRRK2 polymorphisms are also associated with Crohn’s disease and leprosy susceptibility [74,105,113117], indicating a common genetic network at play. Polymorphisms in the autophagy gene ATG16L1 are also linked to Crohn’s disease like NOD2 [118,119], suggesting another shared pathway, and the possibility of a contributing bacterial agent vis-à-vis autophagic control of the intruder.

In one study, the P268S and R702W mutations were linked with a decreased risk of TB, while the A725G mutation was associated with increased risk; all three are predicted to have an impact on the structure of NOD2, but functional data were not provided in this study [120]. Overexpressed NOD2 with R702W recently confirmed decreased signaling capacity [74]. In another study, intronic SNPs in NOD2 (rs6500328 and rs2111234) were associated with an Mtb resister phenotype (persistently negative for disease and immune conversion) but not TB per se [121,122]. Recently, NOD2 was associated with TB meningitis via SKAT-O analysis [123]. Intronic/synonymous SNPs were not evaluated for expression changes in NOD2 in these studies, which remains a possibility. Overall, the mechanism of the association between NOD2 and TB is unclear. Evaluation of the genetic link between NOD2 and TB may require a better definition of TB endophenotypes due to the nature of PRRs’ subtle modifications of immune responses, as has been done for leprosy [8].

NOD2 variants have been associated in one paper with Buruli ulcer in humans (due to M. ulcerans) [124] and in another to Map infection in cattle [125]. Case reports have been published identifying NOD2 mutations in patients with M. abscessus or MAC infection [126], but controlled genetic studies associating NOD2 with these infections have not been performed.

2. The chemical link

The mycobacterial cell envelope is a defining feature of the genus. Mycobacteria have an inner plasma membrane plus an outer “mycomembrane” (Fig 2A). Between these membranes are a poorly defined granular layer, the periplasmic space, PGN, and arabinogalactan [127,128]. The outermost surface of the cell is a carbohydrate capsule, which varies in composition across the genus [129,130]. Distinctly, mycobacterial PGN contains alternating units of N-glycolyl muramic acid (MurNGc) (or less frequently, N-acetyl muramic acid (MurNAc)) and N-acetyl glucosamine (GlcNAc) (Fig 2B). Other bacteria possess only MurNAc in the place of MurNGc. The muramic acid moiety is attached to peptides containing L-alanine (L-Ala), D-isoglutamine (D-isoGln), meso-diaminopimelic acid (meso-DAP), and occasionally 1 or 2 residues of D-alanine (D-Ala). Peptides cross-link glycan strands through DAP-DAP bonds or D-Ala-DAP bonds [131]. Muramyl dipeptide is the fragment containing muramic acid, L-Ala, and D-isoGln, and while most bacteria have N-acetyl MDP (Fig 2C), N-glycolyl MDP is the distinct NOD2 ligand of mycobacteria (Fig 2D). During PGN synthesis, uridine diphosphate-linked N-acetyl muramic acid (UDP-MurNAc) is a cytoplasmic intermediate. In mycobacteria, some of this intermediate becomes N-glycolylated by the enzyme N-acetyl muramic acid hydroxylase (NamH) using molecular oxygen to form UDP-MurNGc [132134]. The actinobacteria genera Nocardia and Rhodococcus have namH, while it is absent in the Streptococcus and Corynebacterium genera [25,127,133,135]. Mycobacterial cell wall synthesis is reviewed in [131].

thumbnail
Fig 2. The cell envelope of mycobacteria and mycobacterial peptidoglycan.

(A) Cross-sectional cartoon of the mycobacterial envelope, showing major regions in approximate relative size to each other. (B) Parallel strands of mycobacterial PGN containing MurNGc, MurNAc, and GlcNAc, with peptides partially cross-linked (red lines). (C) N-acetyl MDP, the minimal PGN fragment recognized by NOD2, common to most bacteria. The sugar and amino acids are separated by differently colored backgrounds. (D) N-glycolyl MDP, the distinct NOD2 ligand of mycobacteria, with the additional hydroxy group from N-glycolylation highlighted in red. The sugar and amino acids are separated by differently colored backgrounds. GlcNAc, N-acetyl glucosamine; MDP, muramyl dipeptide; MurNAc, N-acetyl muramic acid; MurNGc, N-glycolyl muramic acid; NOD2, Nucleotide-binding Oligomerization Domain-containing 2; PGN, peptidoglycan.

https://doi.org/10.1371/journal.ppat.1011389.g002

The presence of NamH in many nonpathogenic actinobacteria suggests that N-glycolylated PGN evolved in a host-free environment; supporting this, the presence of namH has been shown to confer resistance to beta-lactams and lysozyme [127,133]. N-glycolylation of PGN is nonessential for survival of mycobacteria, since it can be deleted from M. smegmatis and Mtb [25,46], and is degenerate in the M. leprae genome [134]. It is noteworthy that the PGN modification is retained in Mtb, despite thousands of years of Mtb–human interaction [136], yet lost in M. leprae. This divergent evolution provides indirect support that the NOD2 ligand plays important roles in the life cycles of pathogenic mycobacteria. The requirement for molecular oxygen for NamH-mediated N-glycolylation [132,133] is interesting, as Mtb experiences different concentrations of O2 in different zones of pulmonary lesions [137]. Indeed, Mtb cultured in vitro under hypoxic conditions up-regulates namH (Rv3818) expression upon returning to an aerobic environment [138]. The regulation and timing of namH expression may inform about the role of NOD2 sensing during Mtb infection and should be further studied.

The sensing of PGN can also occur via other PRRs. NOD1 senses D-glutamyl-meso-diaminopimelic acid, but its role is limited in mycobacterial infection relative to NOD2 [59]. NLRP3 inflammasome activation was shown to proceed through hexokinase sensing of N-acetyl glucosamine released from PGN [139], although this has not been verified in the context of mycobacterial infection. NLRP3 sensing of Mtb involves membrane damage and contributes significantly to the cellular immune response [140,141], but NLRP3 is not critical during Mtb infection of mice [142]. NLRP1 was attributed to MDP sensing [143,144] and PGN recognition proteins (PGLYRPs) contribute to detection of PGN [145148], but none are well examined in mycobacterial infection. While we must be mindful of other PPRs when discussing the immunology of mycobacterial PGN, only NOD2 sensing is known to be specifically affected by the N-glycolylation of PGN in mycobacteria, and the existence of these other PRR sensors does not per se detract from the importance of NOD2.

3. The immunologic link

Soon after NOD2 was described as a PGN sensor, NOD2 was demonstrated to contribute to the immune response to mycobacteria. Overexpression of NOD2 in HEK293T cells was sufficient to induce NF-κB signaling upon Mtb and Map infection, and various cytokine responses to Mtb and Map sonicates were reduced in human PBMCs homozygous for the NOD2 frameshift mutation [15,17]. NOD2 knockdown in human MDMs blunted immune responses to Mtb and BCG [30,53], and knockout of murine Nod2 decreased immune responses of murine macrophages and DCs to Mtb [15,18]. Mouse alveolar macrophages (AMs) required functional NOD2 for the full BCG immune response [22]. Likewise, cytokine production in response to the mycobacterial vaccine Mycobacterium indicus pranii (Mip) was reduced with Nod2 knockdown in murine peritoneal macrophages [40].

NOD2 was necessary for complete tumor necrosis factor (TNF) production in response to BCG, M. smegmatis, Map, and 2 other namH-possessing actinobacteria (Nocardia asteroides and Rhodococcus equi), in contrast to bacteria not having namH (Escherichia coli, Staphylococcus aureus, and Streptomyces sp.) [25]. In studies of complete Freund’s adjuvant (CFA, an emulsion containing dead mycobacteria), immune responses were dependent on both bacterial namH and murine Nod2 [24,64]. Collectively, these results support that the distinct chemical structure of mycobacterial peptidoglycan (i.e., the N-glycolylation of muramic acid by NamH) increases the NOD2-dependent immune response. The increased biological activity of N-glycolyl MDP to N-acetyl MDP was explained by increased NOD2 protein stability [149].

The M. leprae genome is degenerate [150], indicative of an obligate pathogen with a restricted lifestyle. M. leprae has pseudogenized namH, resulting in production of N-acetylated peptidoglycan only [134]. Nevertheless, NOD2 dependence was demonstrated in the immune response to M. leprae [28,32,60] and M. leprae-specific murodipeptides (containing glycine instead of alanine and with variable amidation of the glutamate residue) [60,134], providing grounds for the genetic associations with leprosy. Known M. leprae-associated NOD2 SNPs are noncoding or synonymous [105109]; if their effect is on the expression level of NOD2, which should be investigated, the system may be more vulnerable to a weaker (N-acetyl) MDP ligand. NOD2 polymorphisms were more strongly associated with multibacillary than paucibacillary infection, and with type II (antibody-mediated) than type I (cell-mediated) lepra reactions [105,106]. Mtb is more transmissible in those with cavitary disease requiring cell-mediated immunity [151,152]. This differential flavor of immune response required for efficient transmission may explain the evolutionary decay of namH in M. leprae.

We could find only 2 articles in the literature having data to suggest that N-acetylated MDP and PGN are more immuno-stimulatory. One study using PBMCs only compared MDPs and namH genotypes at single doses [54]. Another using HEK293T cells [153] is at odds with 3 different HEK studies [25,149,154]. There are at least 9 articles directly supporting greater potency with N-glycolylation: N-glycolyl MDP has shown greater potency in HEK cell reporter assays [25,149,154], RAW cells [25,27], primary mouse cells [25,64,155], primary human cells [53], and in vivo mouse models [25,156]. Bacteria possessing NamH elicited a greater immune response in mouse cells [25,46], in vivo mouse models [25,64], and human cells [46]. The conflicting results suggest that timing, MDP amount, or cell type may be critical factors in understanding the NOD2 response to mycobacteria, and future studies should test these factors to uncover the mechanisms of greater N-glycolyl MDP activity. An interaction between NOD2 and mycobacterial sulpholipid-1 was also reported [42], but this does not diminish the importance of the NOD2–mycobacterial MDP interaction.

Expression of NOD2 during mycobacterial infection

Monocytes, MDMs, and AMs express NOD2 at the mRNA and protein levels [30,37]. In one of the first studies on NOD2 expression during Mtb infection, humans with active TB compared to uninfected individuals had similar NOD2 transcript abundance relative to control transcripts in pulmonary leukocytes; 2 outliers with severe TB had exceptionally high NOD2 expression [19]. NOD2 expression increased 2.3-fold in PBMCs following antituberculous treatment but did not significantly change when experimentally infecting PBMCs of healthy controls with various mycobacteria [19]. Another study using BCG-infected mice showed that AMs down-regulated NOD2 expression within 2 weeks of infection [20]. BMDMs and J774A.1 cells infected with live Mtb up-regulated Nod2 transcripts up to 10.6-fold within 24 hours of infection [61]. NOD2 and RIPK2 mRNA were found to be up-regulated 3- to 4-fold in skin tissue samples from people with leprosy compared to controls [34], and the frequency of NOD2-positive cells was 3 times higher in tuberculoid than in lepromatous lesions [41]. Nod2 mRNA expression increased following M. leprae infection of mouse footpads and RAW264.7 cells [51]. Mouse peritoneal macrophages treated with Mip up-regulated Nod2 mRNA expression 3- to 4-fold [40], and goldfish macrophages up-regulated NOD2 mRNA expression 6-fold in response to the fish pathogen M. marinum [45]. Diverse mycobacterial infections appear to be associated with increased NOD2 levels, presumably increasing bacterial sensing, which requires experimental confirmation.

From the point of view that loss-of-function NOD2 variants are detrimental to the host’s immune response, having an increased abundance of functional NOD2 (and by conjecture increased NOD2 signaling) would be beneficial. NOD2 expression or signaling was shown to be dependent on NOD2 polymorphism [72], vitamin D levels [49], hypoxia [42], HIV preexposure prophylaxis [71], and (nonsignificantly) type 1 diabetes status [157], potentially explaining how such noninfectious factors have been associated with TB.

Cell signaling effects of NOD2 during mycobacterial infection

From our literature search and knowledge as stated above, in this section, we summarize what has been demonstrated in experiments testing the interaction between mycobacteria and NOD2 regarding intracellular signaling and intercellular (e.g., cytokine) signaling. NOD2 signaling during mycobacterial infection elicits a myriad of immune effectors from the cell, and the results of our literature analysis are summarized in Fig 3. The first studies to demonstrate that mycobacterial sensing was dependent on NOD2 used NF-κB–based reporter systems [15,17,32]. Live Mtb caused RIPK2 polyubiquitination in infected murine BMDMs independent of MyD88 and TLR2/4 [158], but dependent on NOD2 [27]. In human MDMs infected with Mtb, and PBMCs infected with Map, NOD2-dependent phosphorylation of RIPK2 was observed [30,54], which is required for signaling like ubiquitination [159,160]. During M. abscessus infection of BMDMs, IκBα phosphorylation and degradation were Nod2 independent, whereas phosphorylation of p38, JNK, and, possibly, ERK was Nod2 dependent [62]. N-glycolyl MDP was more potent than N-acetyl MDP in a HEK293 cell NF-κB assay, and in RIPK2 polyubiquitination and phosphorylation of IκBα and JNK but not p38 in RAW264.7 cells [25]. Sensing of mycobacteria and mycobacterial MDP through NOD2 significantly contributes to NF-κB and/or MAPK signaling.

thumbnail
Fig 3. NOD2 signaling during mycobacterial infection.

Cartoon depicting mycobacteria (fuchsia bacilli) being engulfed and sensed by a macrophage through NOD2. Muropeptides including the MDP motif exit the phagosome via a SLC15A family transporter. Upon entering the cytosol, MDP is phosphorylated by NAGK and subsequently sensed via NOD2, which has been recruited to the membrane by palmitoylation from ZDHHC5 (not shown). RIPK2 is phosphorylated and ubiquitinated, mediating signaling cascades and culminating in NF-κB and MAPK transcription factors translocating to the nucleus to regulate inflammatory gene transcription. Various immune mediators, both intracellular and extracellular in activity, result from NOD2 signaling. The degree of evidence for involvement of the depicted elements in the NOD2-dependent immune response to mycobacteria is indicated semiquantitatively (no data, limited/unclear, and many sources). ATG16L1, autophagy-related 16 like 1; IAP, inhibitor of apoptosis; iNOS, inducible nitric oxide synthase; IRF, interferon regulatory factor; IRGM, immunity-related GTPase M; KC, KC(CXCL1); LRRK2, leucine-rich repeat kinase 2; LUBAC, linear ubiquitin chain assembly complex; MAPK, mitogen-activated protein kinase; MDP, muramyl dipeptide; NAGK, N-acetylglucosamine kinase; NOD2, Nucleotide-binding Oligomerization Domain-containing 2; RIPK2, receptor-interacting protein kinase 2; TBK1, TANK-binding kinase 1; TNF, tumor necrosis factor; TRAF, tumor necrosis factor receptor-associated factor.

https://doi.org/10.1371/journal.ppat.1011389.g003

Numerous studies have reported significant NOD2 dependence of TNF production from various types of human and mouse cells in response to different mycobacteria [15,17,18,22,2527,30,32,46,48,58,59,62,66,74]. TNF is critical for host resistance to mycobacterial infection: Animal studies of Mtb or BCG infection using knockouts or anti-TNF demonstrated that this pathway is essential for survival [161164]. Additionally, humans receiving anti-TNF therapy are at heightened risk of developing TB [165]. Other Th1 cytokines are also important in mycobacterial immunity. IFN-γ is essential for the survival of mice infected with Mtb [166,167], restricting the extrapulmonary bacterial burden [168]. IL-12p40 is also critical during murine TB [169]. Moreover, humans with genetic mutations in the IFN-γ/IL-12 pathways are likely to develop disseminated BCG postvaccination, NTM infection, or later Mtb infection [170]. Multiple reports support that IL-12 responses require functional NOD2. Knockout of Nod2 in BMDMs reduced IL-12p40 release by about half during Mtb infection [18]. A contemporary study showed the same reduction in IL-12p40 production with Nod2-KO naive AMs infected with BCG ex vivo, as well as lung macrophages from BCG-infected mice restimulated ex vivo with BCG [22]. With M. smegmatis intraperitoneal immunization, Nod2 and namH were necessary for full splenocyte IL-12p40 secretion 14 days later upon rechallenge with the same bacteria (the number of IFN-γ–producing cells and total cytokine release were also Nod2 and namH dependent) [25]. During murine M. avium infection, Nod2-KO resulted in significant reductions in IL-12p40 and IFN-γ transcripts in splenocytes compared to wild-type (WT) controls at 100 days postinfection; IFN-γ in splenocyte supernatant after restimulation with M. avium was greatly reduced with Nod2-KO at the same time point, but not TNF [48]. M. avium-infected BMDM IL-12p70 production was also dependent on Nod2 in this study [48]. These results on Th1 cytokines are also reflected with mycobacterial adjuvants: The generation of antigen-specific CD4+ and CD8+ IFN-γ–producing T cells in spleens of mice immunized with CFA required Nod2 [24]. We demonstrated that both functional namH and Nod2 are required for CFA-induced antigen-specific CD4+ T cells, which secrete IFN-γ and/or IL-17A, and that about half of the adjuvancy of mycobacteria in CFA can be recapitulated with synthetic N-glycolyl MDP and a minimal motif of the mycobacterial glycolipid trehalose-6,6′-dimycolate (TDM), which signals through Mincle [64]. Thus, NOD2 contributes to promoting the Th1 response to mycobacteria.

Mycobacteria-induced IL-6 was significantly dependent on NOD2 in human and mouse cells in multiple reports [18,58,59,62,66,74], with corroborating nonsignificant trends in others [15,31]. In but one study investigating the Th1/17-inducing capacity of CFA, injection site IL-6 and IL-23a transcription were independent of Nod2, which may be specific to this model or tissue [43]. NOD2-dependent IL-6 could be contributing to the generation of Th17 cells. We have shown that the generation of IL-17A–producing T cells upon CFA immunization depends on Nod2 [64]. Mtb PGN plus TDM, or N-glycolyl MDP plus synthetic TDM, are sufficient to generate antigen-specific IL-17A–producing CD4+ T cells in mice [43,64]. Two reports of IL-17A production from human PBMCs infected with Mtb were inconsistent on NOD2 dependence, comparing functional NOD2 to 3020insC NOD2 [29,58].

Excess type I IFN drives susceptibility to Mtb in mice [171] and is associated with TB disease in humans [172]; it is thought to be detrimental for the Mtb-infected host in most (but not all) contexts [173]. Nod2-dependent type I IFN production in response to Mtb was demonstrated in 2 studies using murine BMDMs. NOD2 and RIPK2 were required for about half of the IFN-α and IFN-β transcription and secretion within 4 hours of infection with Mtb [23,27]. Moreover, the Nod2-dependent IFN-β transcriptional response appears to be mediated through IRF5, not IRF3, and may require Mtb to have cytosolic access via the ESX-1 secretion system, given that a ΔESX1 mutant was incapable of inducing IFN-β transcription [27]. However, a later study challenged this model, showing that IRF3 was absolutely essential for the type I IFN response to Mtb, while ESX-1 was still necessary, and that absence of Nod2 was not associated with a significant decrease of IFN-β transcription in BMDMs [39]. RIPK2 was also dispensable for IFN-β transcription in Mtb-infected BMDMs [21]. Apart from Mtb, TBK-1 phosphorylation and IFN-β transcription in M. abscessus-infected BMDMs was Nod2 dependent, and recombinant IFN-β promoted bacterial clearance in vivo [68]. Clearly, in some situations of mycobacterial infection, which are important to define, type I IFN is NOD2 dependent. Therefore, future studies must determine when (e.g., early/late) and where (e.g., lung/extrapulmonary) NOD2-dependent type I IFN exists, and whether it is beneficial or harmful in that context.

Multiple studies using IL-1 receptor knockout mice infected with Mtb have established the importance of IL-1 signaling [174176]. IL-1R signaling was critical to limit pathological type I IFN in murine TB [171]. IL-1α/β knockout mice were similarly or less susceptible than the IL-1 receptor knockout [164,177,178]. In PBMCs responding to Map or Mtb, full IL-1β secretion required functional NOD2 (without 3020insC) [17,26,58]; HEK293T cells infected with M. leprae required a NOD2 expression plasmid for IL1B transcription [32]. An siRNA-mediated NOD2 knockdown in MDMs reduced IL-1β secretion by over 80% upon Mtb and BCG infection [30]. However, an siRNA-mediated NOD2 knockdown (only 60% reduction in NOD2 transcript abundance) in THP-1 cells resulted in insignificant difference in IL-1β production following Mtb infection [44]. IL-1β data are less consistent in mouse cells. With murine macrophages (peritoneal and BMDMs), the release of IL-1β upon infection with Mtb, BCG [30] and M. abscessus [62], and Il1b transcription with Mip [40] was NOD2 dependent. However, NOD2 independence in IL-1β secretion upon Mtb infection was shown with BMDCs [35] and BMDMs [36,66], as well as Il1b transcription in CFA-immunized skin biopsy [43]. As both transcript and protein data are inconsistent across studies, a difference in caspase activity between cell types cannot fully explain the conflict. Other species/cell differences may explain the results, which require further experimentation. Nevertheless, NOD2 and IL-1β production are linked in certain contexts.

While IL-10 has the capacity to dampen immunopathological responses, its inhibition of protective immune system functions is thought to contribute to mycobacterial susceptibility [179]. NOD2-depdendent IL-10 was reported in human PBMCs treated with sonicated Mtb and Map [15,17]. A corroborative trend was observed in another study of Map infection of human monocytes [31]. Murine BMDCs also required functional Nod2 for maximum IL-10 production in response to Mtb infection [18]. However, AMs infected with BCG ex vivo and lung APCs from BCG-infected mice restimulated ex vivo with BCG or Mtb culture filtrate did not depend on Nod2 for IL-10 production [22]. Splenocytes from mice infected with M. avium also did not significantly depend on Nod2 for IL-10 production when reexposed ex vivo to M. avium [48]. Cell type–specific dependence on NOD2 could explain the inconsistencies.

Mtb and BCG can induce IL-32 production from PBMCs (mainly from monocytes) by a Caspase-1/IL-18/IFN-γ–dependent pathway [180]. IL-32 appears to be protective during Mtb infection [181,182]. NOD2 contributes to IL-32 production in response to M. leprae: MDP was able to drive IL-32–dependent differentiation of human DCs, promoting CD1b expression, and defects in this pathway correlated with M. leprae disease, providing biological grounds for the association of NOD2 polymorphisms with leprosy [41]. The dependence of the M. leprae immune response on NOD2 was subsequently demonstrated, where human monocytes’ up-regulation of IL-32 production and CD1b expression to M. leprae and M. leprae-derived muropeptides required NOD2 [60].

There are limited data on the NOD2 dependence of other immune effectors elicited by mycobacteria. Production of IL-22 by Mtb-infected PBMCs required functional NOD2 [58]. In CFA-immunized mice, antigen-specific splenic IL-4 plus serum IgG1 and IgG2c required Nod2 [24]. For chemokines, Nod2 dependence was shown for RANTES from Mtb-infected BMDMs, and KC (CXCL1) from Mtb lysate-stimulated BMDMs [58] or M. smegmatis-infected murine peritoneal cavity [25].

NOD2 has no mycobactericidal effect per se

In this section, we summarize the mechanisms of mycobacterial control/killing, which have been examined as a function of NOD2 status: nitric oxide (NO), bacterial burden during in vitro cellular infections, autophagy, and host cell death modality. Inducible nitric oxide synthase (iNOS), an important source of bactericidal NO, is critical for host resistance to Mtb infection [183185]. N-glycolyl MDP was able to induce expression of iNOS and NO production in human MDMs to levels similar to or greater than N-acetyl MDP or IFN-γ, and iNOS expression was NOD2 dependent during Mtb infection [53]. During Mtb infection of IFN-γ–stimulated BMDMs, generation of the NO product nitrite was Nod2 dependent, but bacterial control was affected only by IFN-γ and not Nod2 genotype. Nod2 was also required for nitrite production in response to heat-killed Mtb [18]. In a study of BCG-infected mice, lung macrophages examined ex vivo 4 weeks postinfection required Nod2 for maximum nitrite production [22]. IFN-γ–stimulated BMDMs infected with M. abscessus similarly required Nod2 for nitrite production [62].

Multiple studies have tested the NOD2 dependence of in vitro bacterial control. Human MDMs required NOD2 for control of Mtb and BCG, while Nod2 did not help murine BMDMs control Mtb, although it was necessary for BCG control [30]. Control of Map in monocytes from Crohn’s disease patients heterozygous for the common NOD2 mutations was not significantly hampered compared to patients homozygous for functional NOD2, suggesting NOD2 haplosufficiency or no role for NOD2 in Map control [31]. IFN-γ and iNOS were required for Nod2-dependent control of M. abscessus by BMDMs [62]. Other mouse studies did not support a role for Nod2 for in vitro bacterial control of Mtb, but they used BMDMs that were not stimulated with IFN-γ [18,66], as in Brooks and colleagues [30]. The requirement of IFN-γ for BMDMs to engage NOD2 in mycobacterial control should be further investigated. Macrophage ontogeny may be critical and informative about NOD2-dependent control of mycobacteria, but these elements have not yet been directly studied together.

Mtb can also be controlled through the regulation of cell death modality [186] and autophagy [187]. Autophagic control of Mtb in human AMs can be enhanced with supplementary MDP, associated with the recruitment of ATG16L1, IRGM, and LC3 to the Mtb-containing phagosome [37], and NOD2 is known to contribute to autophagy by RIPK2-dependent and/or independent mechanisms [188,189]. However, data directly demonstrating the importance of physiological NOD2 signaling to autophagic control of mycobacteria is lacking. During infection with avirulent Mtb (H37Ra), inclusion of a dominant negative RIPK2 construct in THP-1 cells did not significantly alter apoptosis, despite the fact that MDP was capable of suppressing apoptosis in the same uninfected cells [16]. Another study examining NOD2 during virulent Mtb (H37Rv) infection of murine BMDCs found no significant difference in apoptosis between WT and Nod2-KO cells [35]. HEK293T cells required overexpression of NOD2 during stimulation with live M. leprae to optimally up-regulate CASP1 expression [51]. Recently, an examination of LRRK2 and NOD2 polymorphisms in BCG-infected RAW264.7 cells revealed that while NOD2 and LRRK2 jointly contribute to ROS production, only LRRK2 genotype had a significant impact on apoptosis [74]. We could find no evidence directly linking NOD2 per se to apoptosis during mycobacterial infection.

Murine mycobacterial infection as a function of NOD2

In the first reported Mtb infections of Nod2-deficient animals (aerosol infections with 35 CFU of strain 1254 or 1,500 CFU of H37Rv), the burden of Mtb in the lungs of Nod2-KO C57BL/6 mice was indistinguishable from WT mice up to 8 weeks post aerosol infection, with no obvious histopathological differences and only a small increase in serum IL-12p40 in the knockout at earlier time points [18]. In a subsequent study using aerosol infection with 100 CFU of H37Rv, the lack of difference in bacterial burden up to 8 weeks postinfection was confirmed, but Nod2-KO mice (C57BL/6) had less pulmonary leukocyte infiltration than WT during Mtb and BCG infections [22]. Upon further examination of BCG infection, Nod2-KO mice had a pulmonary reduction in CD4+ and CD8+ T cell frequency compared to WT, plus significantly reduced TNF and IFN-γ in BAL fluid reflected by defective cytokine responses in the spleen and lymph nodes at 4 weeks postinfection. Pulmonary Mtb burdens at 24 weeks postinfection with 400 CFU of H37Rv were significantly higher in Nod2-KO animals compared to WT, and survival was reduced. We recently reported again on Mtb-infected Nod2-KO mice (C57BL/6) while investigating dual (NOD2 and Mincle) PRR knockout: After aerosol infection with 15 CFU H37Rv, there was no significant effect on bacterial burden and pulmonary leukocyte infiltration by 6 weeks postinfection [73]. After aerosol infection with 31 CFU H37Rv, the survival of Nod2-KO and Nod2-KO Mincle-KO (double knockout) animals was similar (291 and 281 days, respectively) and significantly reduced compared to WT (over 363 days), while Mincle-KO survival was intermediate (324 days), suggesting that Nod2 deficiency was the driver of hastened mortality in the double knockout. Interestingly, increased pulmonary cell death with distinct necrotic foci occurred in the lungs of only some of the Nod2-deficient animals upon TB-prescribed euthanasia [73]. These studies confirm that Nod2 plays an important, albeit delayed, role in host resistance to Mtb. The pulmonary necrosis phenotype should be further investigated as mechanism of increased susceptibility to Mtb in Nod2-deficient animals.

C57BL/6 murine infection with a H37Rv namH mutant of Mtb showed no differences in pulmonary nor splenic bacterial burden up to 12 weeks post aerosol infection with approximately 30 CFU, and pulmonary inflammation was indistinguishable from WT Mtb-infected animals at week 3 [46]. Comparing survival as a function of namH in susceptible C57BL/6-DBA2 F1 hybrid mice, no significant differences were obtained. In an experiment with C57BL/6 Rag1-KO mice (emphasizing the conditions of early infection where innate immunity predominates), mice aerosol infected with 51 CFU of the namH mutant Mtb survived significantly longer than those with 34 CFU of WT Mtb [46]. In this case, a less inflammatory pathogen permitted slightly prolonged survival in a mouse model where mortality occurs as inflammation compromises lung function.

A recent study described an N-deacetylase in Mtb, encoded by Rv1096, which is expected to remove N-acetyl groups from the PGN polymer at GlcNAc and/or MurNAc. When Rv1096 was ectopically expressed in M. smegmatis, NOD2- and TLR-dependent reduced inflammatory responses plus prolonged bacterial persistence were observed during cellular and C57BL/6 mouse infections. The corresponding knockout in Mtb H37Rv had corroboratory effects, although only data demonstrating overall partial deacetylation of peptidoglycan with the M. smegmatis mutant were presented using Fourier-transform infrared spectroscopy; no detailed chemical structure was confirmed [66]. This exemplifies the trade-off of reduced NOD2-mediated acute inflammation with increased bacterial persistence.

Like Mtb infection, bacterial burden in C56BL/6 mice intravenously infected with 106 CFU of M. avium was unaffected by Nod2 genotype at 30 and 100 days postinfection, yet immunological defects were detected along the Th1 axis [48]. Liver granuloma size plus lymphocyte and granulocyte numbers were reduced in M. avium-infected Nod2-KO mice compared to WT at day 100 [48]. Although the organ and phenotype are different, together with the aforementioned pulmonary necrotizing foci, the data point to Nod2 in the long-term regulation of mycobacterial lesions such as the granuloma. As discussed in our Mtb study, pulmonary necrotizing foci are a pathological characteristic of Mtb infection in the “Kramnik mouse” [190,191] which has been linked to excessive type I IFN [171]. An NTM model of infection could permit a more rapid and/or convenient investigation of NOD2 in this phenotype.

High-dose (1.5 × 107 CFU) intranasal infection with M. abscessus proceeds differently than Mtb and M. avium infections: M. abscessus is cleared from the lungs of immunocompetent mice. However, with Nod2-KO, clearance lagged behind WT C57BL/6 mice by about 1 log up to 20 days postinfection [62]. Concomitantly, pulmonary histopathology was worse in Nod2-KO. IFN-γ remained reduced in lung homogenates of Nod2-KO, while TNF and IL-6 were dysregulated [62]. The results suggested insufficient immunity early resulting in pathological inflammation later, as bacterial clearance was not efficient. In a subsequent M. abscessus study of the same model, Nod2-KO in C57BL/6 mice resulted in defective NO production and bacterial control in the lung, both of which could be rescued with recombinant IFN-β, suggesting a NOD2-IFN-β-NO pathway of M. abscessus control early in infection [68].

NOD2 and mycobacteria beyond infection

The potent adjuvant activity of mycobacteria has been renowned since Jules Freund invented his eponymous adjuvant in the mid-20th century. Various mycobacteria were discovered to contain N-glycolylated muropeptides from 1969 to 1970 [135,192195], and by the 1970s, the adjuvant activity of mycobacteria was attributed to MDP [196]. The adjuvant effect of mycobacteria has been recapitulated with purified TDM plus PGN in IFA [43], and more specifically with synthetic TDM (GlcC14C18) plus N-glycolyl MDP in IFA to generate experimental autoimmune encephalomyelitis (EAE) [64], demonstrating that N-glycolylated PGN/MDP is sufficient to recreate the adjuvant effect in combination with one other mycobacterial MAMP. Synergistic activation of the immune response in vitro and in vivo was also reported between MDP and purified Mtb Hsp70, a putative TLR/MyD88 MAMP [52,197]. While Nod2-KO and Ripk2-KO mice are resistant to EAE induced with CFA [33], one study reported that NOD2 played an immunoregulatory role in CFA-induced experimental autoimmune uveitis through T cell–intrinsic, RIPK2-independent NOD2 activity [67]; this curious result still affirms a role for NOD2 in the mycobacterial adjuvant effect.

Antimycobacterial MDP synergy was explored with the unnatural combination of purified LPS and N-glycolyl MDP (gram-negative and actinobacterial MAMPs) to successfully boost various BMDC effector functions during Mtb infection [5557,69]. Similar results were reported with murine mesenchymal stem cells [70]. Prophylactically, a NOD2-RIG-I dual activating small molecule called Inarigivir enhanced BCG vaccine protection during murine Mtb infection [65]. Even alone, N-glycolyl MDP administered during intranasal infection with M. abscessus or influenza A virus improved pulmonary clearance and reduced pathology [62,198].

Trained innate immunity elicited by BCG may be key to the utility of the bacillus [38,199]. Cancer is being targeted nonspecifically with BCG [200], Mip [201,202], and the Mtb extract Z-100 containing PGN [50,203]. With monocytes homozygous for the NOD2 frameshift mutation, trained innate immunity was defective after in vitro BCG [38] or gamma-irradiated BCG training [47]. Corroborative results were obtained using an RIPK2 inhibitor, and MDP alone induced a similar effect as BCG [38]. Thus, mycobacterial stimulation of NOD2 may contribute to trained innate immunity. However, double knockout of Nod1 and Nod2 did not significantly reduce the ability of BCG to control murine Mtb infection, indicating that BCG provides Nod2-independent protection too [63]. Targeting the NOD2 pathway in the classical vaccine approach, treating infection and trained-innate immunity, continues to be promising and can be informed by studies on mycobacteria but requires further research before clinical applications can be tested.

Conclusions

Nearly 2 decades of research have asserted a significant relationship between mycobacteria and NOD2. Polymorphisms in NOD2 have been associated with mycobacterial diseases. The mycobacterial NOD2 ligand N-glycolyl MDP is unique, shows enhanced NOD2 activity, and contributes significantly to the immune response. NOD2 sensing of mycobacteria manifests as heightened production of various immune effectors and likely contributes to bacteriologic control. NOD2 is protective against death during Mtb infection of animals by an unknown mechanism. Future studies will need to focus on NOD2-dependent mechanisms of immunopathogenesis (potentially type I IFN and/or cell death) to uncover how absence of functional NOD2 leads to altered disease phenotypes. Conversely, promoting the NOD2 pathway has demonstrated immunotherapeutic potential and merits further development.

Acknowledgments

J.-Y.D. would like to thank Hanna Ostapska for advice on graphically representing glycans and editing the figures.

References

  1. 1. Koch R. Die Aetiologie der Tuberculose. Berliner Klinische Wochenschrift. 1882;15:287–296.
  2. 2. Sizemore CF, Hafner R, Fauci AS. NIH Statement on World Tuberculosis Day. 2018:2018.
  3. 3. World Health. Organization. 2022. Global Tuberculosis Report 2022.
  4. 4. World Health Organization. 2021. Towards zero leprosy. Global Leprosy (Hansen’s disease) Strategy 2021–2030.
  5. 5. Duffy SC, Srinivasan S, Schilling MA, Stuber T, Danchuk SN, Michael JS, et al. Reconsidering Mycobacterium bovis as a proxy for zoonotic tuberculosis: a molecular epidemiological surveillance study. Lancet Microbe. 2020;1:e66–e73.
  6. 6. APHIS. 2008. Johne’s Disease on U.S. Dairies, 1991–2007. United States Department of Agriculture. Available from: https://www.aphis.usda.gov/animal_health/nahms/dairy/downloads/dairy07/Dairy07_is_Johnes_1.pdf.
  7. 7. Gill CO, Saucier L, Meadus WJ. Mycobacterium avium subsp. paratuberculosis in dairy products, meat, and drinking water. J Food Prot. 2011;74:480–499.
  8. 8. Dubé J-Y, Fava VM, Schurr E, Behr MA. Underwhelming or Misunderstood? Genetic Variability of Pattern Recognition Receptors in Immune Responses and Resistance to Mycobacterium tuberculosis. Front Immunol. 2021;12.
  9. 9. Hampe J, Cuthbert A, Croucher PJ, Mirza MM, Mascheretti S, Fisher S, et al. Association between insertion mutation in NOD2 gene and Crohn’s disease in German and British populations. Lancet. 2001;357:1925–1928. pmid:11425413
  10. 10. Hugot JP, Chamaillard M, Zouali H, Lesage S, Cézard JP, Belaiche J, et al. Association of NOD2 leucine-rich repeat variants with susceptibility to Crohn’s disease. Nature. 2001;411:599–603. pmid:11385576
  11. 11. Ogura Y, Bonen DK, Inohara N, Nicolae DL, Chen FF, Ramos R, et al. A frameshift mutation in NOD2 associated with susceptibility to Crohn’s disease. Nature. 2001;411:603–606.
  12. 12. Miceli-Richard C, Lesage S, Rybojad M, Prieur AM, Manouvrier-Hanu S, Hafner R, et al. CARD15 mutations in Blau syndrome. Nat Genet. 2001;29:19–20. pmid:11528384
  13. 13. Girardin SE, Boneca IG, Viala J, Chamaillard M, Labigne A, Thomas G, et al. Nod2 is a general sensor of peptidoglycan through muramyl dipeptide (MDP) detection. J Biol Chem. 2003;278:8869–8872. pmid:12527755
  14. 14. Inohara N, Ogura Y, Fontalba A, Gutierrez O, Pons F, Crespo J, et al. Host recognition of bacterial muramyl dipeptide mediated through NOD2. Implications for Crohn’s disease. J Biol Chem. 2003;278:5509–5512. pmid:12514169
  15. 15. Ferwerda G, Girardin SE, Kullberg BJ, Le Bourhis L, de Jong DJ, Langenberg DM, et al. NOD2 and toll-like receptors are nonredundant recognition systems of Mycobacterium tuberculosis. PLoS Pathog. 2005;1:279–285.
  16. 16. Loeuillet C, Martinon F, Perez C, Munoz M, Thome M, Meylan PR. Mycobacterium tuberculosis subverts innate immunity to evade specific effectors. J Immunol. 2006;177:6245–6255.
  17. 17. Ferwerda G, Kullberg BJ, de Jong DJ, Girardin SE, Langenberg DM, van Crevel R, et al. Mycobacterium paratuberculosis is recognized by Toll-like receptors and NOD2. J Leukoc Biol. 2007;82:1011–1018.
  18. 18. Gandotra S, Jang S, Murray PJ, Salgame P, Ehrt S. Nucleotide-binding oligomerization domain protein 2-deficient mice control infection with Mycobacterium tuberculosis. Infect Immun. 2007;75:5127–5134.
  19. 19. Lala S, Dheda K, Chang JS, Huggett JF, Kim LU, Johnson MA, et al. The pathogen recognition sensor, NOD2, is variably expressed in patients with pulmonary tuberculosis. BMC Infect Dis. 2007;7:96. pmid:17705850
  20. 20. Srivastava M, Meinders A, Steinwede K, Maus R, Lucke N, Bühling F, et al. Mediator responses of alveolar macrophages and kinetics of mononuclear phagocyte subset recruitment during acute primary and secondary mycobacterial infections in the lungs of mice. Cell Microbiol. 2007;9:738–752. pmid:17054437
  21. 21. Stanley SA, Johndrow JE, Manzanillo P, Cox JS. The Type I IFN Response to Infection with Mycobacterium tuberculosis Requires ESX-1-Mediated Secretion and Contributes to Pathogenesis. J Immunol. 2007;178:3143–3152.
  22. 22. Divangahi M, Mostowy S, Coulombe F, Kozak R, Guillot L, Veyrier F, et al. NOD2-Deficient Mice Have Impaired Resistance to Mycobacterium tuberculosis Infection through Defective Innate and Adaptive Immunity. J Immunol. 2008;181:7157–7165.
  23. 23. Leber JH, Crimmins GT, Raghavan S, Meyer-Morse NP, Cox JS, Portnoy DA. Distinct TLR- and NLR-mediated transcriptional responses to an intracellular pathogen. PLoS Pathog. 2008;4:e6. pmid:18193943
  24. 24. Magalhaes JG, Fritz JH, Le Bourhis L, Sellge G, Travassos LH, Selvanantham T, et al. Nod2-dependent Th2 polarization of antigen-specific immunity. J Immunol. 2008;181:7925–7935. pmid:19017983
  25. 25. Coulombe F, Divangahi M, Veyrier F, de Leseleuc L, Gleason JL, Yang Y, et al. Increased NOD2-mediated recognition of N-glycolyl muramyl dipeptide. J Exp Med. 2009;206:1709–1716.
  26. 26. Kleinnijenhuis J, Joosten LA, van de Veerdonk FL, Savage N, van Crevel R, Kullberg BJ, et al. Transcriptional and inflammasome-mediated pathways for the induction of IL-1beta production by Mycobacterium tuberculosis. Eur J Immunol. 2009;39:1914–1922.
  27. 27. Pandey AK, Yang Y, Jiang Z, Fortune SM, Coulombe F, Behr MA, et al. NOD2, RIP2 and IRF5 play a critical role in the type I interferon response to Mycobacterium tuberculosis. PLoS Pathog. 2009;5:e1000500.
  28. 28. Sinsimer D, Fallows D, Peixoto B, Krahenbuhl J, Kaplan G, Manca C. Mycobacterium leprae actively modulates the cytokine response in naive human monocytes. Infect Immun. 2010;78:293–300.
  29. 29. van de Veerdonk FL, Teirlinck AC, Kleinnijenhuis J, Kullberg BJ, van Crevel R, van der Meer JW, et al. Mycobacterium tuberculosis induces IL-17A responses through TLR4 and dectin-1 and is critically dependent on endogenous IL-1. J Leukoc Biol. 2010;88:227–232.
  30. 30. Brooks MN, Rajaram MV, Azad AK, Amer AO, Valdivia-Arenas MA, Park JH, et al. NOD2 controls the nature of the inflammatory response and subsequent fate of Mycobacterium tuberculosis and M. bovis BCG in human macrophages. Cell Microbiol. 2011;13:402–418.
  31. 31. Glubb DM, Gearry RB, Barclay ML, Roberts RL, Pearson J, Keenan JI, et al. NOD2 and ATG16L1 polymorphisms affect monocyte responses in Crohn’s disease. World J Gastroenterol. 2011;17:2829–2837.
  32. 32. Kang TJ, Chae GT. The Role of Intracellular Receptor NODs for Cytokine Production by Macrophages Infected with Mycobacterium leprae. Immune Netw. 2011;11:424–427.
  33. 33. Shaw PJ, Barr MJ, Lukens JR, McGargill MA, Chi H, Mak TW, et al. Signaling via the RIP2 adaptor protein in central nervous system-infiltrating dendritic cells promotes inflammation and autoimmunity. Immunity. 2011;34:75–84. pmid:21236705
  34. 34. Sun Y, Liu H, Yu G, Chen X, Liu H, Tian H, et al. Gene expression analysis of leprosy by using a multiplex branched DNA assay. Exp Dermatol. 2011;20:520–522. pmid:21585556
  35. 35. Abdalla H, Srinivasan L, Shah S, Mayer-Barber KD, Sher A, Sutterwala FS, et al. Mycobacterium tuberculosis infection of dendritic cells leads to partially caspase-1/11-independent IL-1β and IL-18 secretion but not to pyroptosis. PLoS ONE. 2012;7:e40722.
  36. 36. Dorhoi A, Nouailles G, Jörg S, Hagens K, Heinemann E, Pradl L, et al. Activation of the NLRP3 inflammasome by Mycobacterium tuberculosis is uncoupled from susceptibility to active tuberculosis. Eur J Immunol. 2012;42:374–384.
  37. 37. Juárez E, Carranza C, Hernández-Sánchez F, León-Contreras JC, Hernández-Pando R, Escobedo D, et al. NOD2 enhances the innate response of alveolar macrophages to Mycobacterium tuberculosis in humans. Eur J Immunol. 2012;42:880–889.
  38. 38. Kleinnijenhuis J, Quintin J, Preijers F, Joosten LA, Ifrim DC, Saeed S, et al. Bacille Calmette-Guerin induces NOD2-dependent nonspecific protection from reinfection via epigenetic reprogramming of monocytes. Proc Natl Acad Sci U S A. 2012;109:17537–17542. pmid:22988082
  39. 39. Manzanillo PS, Shiloh MU, Portnoy DA, Cox JS. Mycobacterium tuberculosis activates the DNA-dependent cytosolic surveillance pathway within macrophages. Cell Host Microbe. 2012;11:469–480.
  40. 40. Pandey RK, Sodhi A, Biswas SK, Dahiya Y, Dhillon MK. Mycobacterium indicus pranii mediates macrophage activation through TLR2 and NOD2 in a MyD88 dependent manner. Vaccine. 2012;30:5748–5754.
  41. 41. Schenk M, Krutzik SR, Sieling PA, Lee DJ, Teles RM, Ochoa MT, et al. NOD2 triggers an interleukin-32-dependent human dendritic cell program in leprosy. Nat Med. 2012;18:555–563. pmid:22447076
  42. 42. Nabatov AA, Hatzis P, Rouschop KM, van Diest P, Vooijs M. Hypoxia inducible NOD2 interacts with 3-O-sulfogalactoceramide and regulates vesicular homeostasis. Biochim Biophys Acta. 2013;1830:5277–5286. pmid:23880069
  43. 43. Shenderov K, Barber DL, Mayer-Barber KD, Gurcha SS, Jankovic D, Feng CG, et al. Cord factor and peptidoglycan recapitulate the Th17-promoting adjuvant activity of mycobacteria through Mincle/CARD9 signaling and the inflammasome. J Immunol. 2013;190:5722–5730. pmid:23630357
  44. 44. Verway M, Bouttier M, Wang TT, Carrier M, Calderon M, An BS, et al. Vitamin D induces interleukin-1β expression: paracrine macrophage epithelial signaling controls M. tuberculosis infection. PLoS Pathog. 2013;9:e1003407.
  45. 45. Xie J, Hodgkinson JW, Katzenback BA, Kovacevic N, Belosevic M. Characterization of three Nod-like receptors and their role in antimicrobial responses of goldfish (Carassius auratus L.) macrophages to Aeromonas salmonicida and Mycobacterium marinum. Dev Comp Immunol. 2013;39:180–187.
  46. 46. Hansen JM, Golchin SA, Veyrier FJ, Domenech P, Boneca IG, Azad AK, et al. N-glycolylated peptidoglycan contributes to the immunogenicity but not pathogenicity of Mycobacterium tuberculosis. J Infect Dis. 2014;209:1045–1054.
  47. 47. Arts RJ, Blok BA, Aaby P, Joosten LA, de Jong D, van der Meer JW, et al. Long-term in vitro and in vivo effects of γ-irradiated BCG on innate and adaptive immunity. J Leukoc Biol. 2015;98:995–1001.
  48. 48. Carvalho NB, Oliveira FS, Marinho FA, de Almeida LA, Fahel JS, Báfica A, et al. Nucleotide-binding oligomerization domain-2 (NOD2) regulates type-1 cytokine responses to Mycobacterium avium but is not required for host control of infection. Microbes Infect. 2015;17:337–344.
  49. 49. Coussens AK, Wilkinson RJ, Martineau AR. Phenylbutyrate Is Bacteriostatic against Mycobacterium tuberculosis and Regulates the Macrophage Response to Infection, Synergistically with 25-Hydroxy-Vitamin D3. PLoS Pathog. 2015;11:e1005007.
  50. 50. Katsunuma K, Yoshinaga K, Ohira Y, Eta R, Sato T, Horii T, et al. Z-100, extracted from Mycobacterium tuberculosis strain Aoyama B, promotes TNF-α production via nucleotide-binding oligomerization domain containing 2 (Nod2)-dependent NF-κB activation in RAW264.7 cells. Mol Immunol. 2015;64:218–227.
  51. 51. Kim TH, Chae GT, Kang TJ. Expression of Nucleotide-oligomerization Domain (NOD) and Related Genes in Mouse Tissues Infected with Mycobacterium leprae. Immune Netw. 2015;15:319–324.
  52. 52. Kim TH, Park JH, Park YM, Ryu SW, Shin SJ, Park JH, et al. Synergistic effect of muramyl dipeptide with heat shock protein 70 from Mycobacterium tuberculosis on immune activation. Immunobiology. 2015;220:26–31.
  53. 53. Landes MB, Rajaram MV, Nguyen H, Schlesinger LS. Role for NOD2 in Mycobacterium tuberculosis-induced iNOS expression and NO production in human macrophages. J Leukoc Biol. 2015;97:1111–1119.
  54. 54. Salem M, Seidelin JB, Eickhardt S, Alhede M, Rogler G, Nielsen OH. Species-specific engagement of human nucleotide oligomerization domain 2 (NOD)2 and Toll-like receptor (TLR) signalling upon intracellular bacterial infection: role of Crohn’s associated NOD2 gene variants. Clin Exp Immunol. 2015;179:426–434. pmid:25335775
  55. 55. Khan N, Aqdas M, Vidyarthi A, Negi S, Pahari S, Agnihotri T, et al. Triggering through NOD-2 Differentiates Bone Marrow Precursors to Dendritic Cells with Potent Bactericidal activity. Sci Rep. 2016;6:27263. pmid:27265209
  56. 56. Khan N, Pahari S, Vidyarthi A, Aqdas M, Agrewala JN. NOD-2 and TLR-4 Signaling Reinforces the Efficacy of Dendritic Cells and Reduces the Dose of TB Drugs against Mycobacterium tuberculosis. J Innate Immun. 2016;8:228–242.
  57. 57. Khan N, Vidyarthi A, Pahari S, Negi S, Aqdas M, Nadeem S, et al. Signaling through NOD-2 and TLR-4 Bolsters the T cell Priming Capability of Dendritic cells by Inducing Autophagy. Sci Rep. 2016;6:19084. pmid:26754352
  58. 58. Lachmandas E, Beigier-Bompadre M, Cheng SC, Kumar V, van Laarhoven A, Wang X, et al. Rewiring cellular metabolism via the AKT/mTOR pathway contributes to host defence against Mycobacterium tuberculosis in human and murine cells. Eur J Immunol. 2016;46:2574–2586.
  59. 59. Lee JY, Hwang EH, Kim DJ, Oh SM, Lee KB, Shin SJ, et al. The role of nucleotide-binding oligomerization domain 1 during cytokine production by macrophages in response to Mycobacterium tuberculosis infection. Immunobiology. 2016;221:70–75.
  60. 60. Schenk M, Mahapatra S, Le P, Kim HJ, Choi AW, Brennan PJ, et al. Human NOD2 Recognizes Structurally Unique Muramyl Dipeptides from Mycobacterium leprae. Infect Immun. 2016;84:2429–2438.
  61. 61. Andreu N, Phelan J, de Sessions PF, Cliff JM, Clark TG, Hibberd ML. Primary macrophages and J774 cells respond differently to infection with Mycobacterium tuberculosis. Sci Rep. 2017;7:42225.
  62. 62. Lee JY, Lee MS, Kim DJ, Yang SJ, Lee SJ, Noh EJ, et al. Nucleotide-Binding Oligomerization Domain 2 Contributes to Limiting Growth of Mycobacterium abscessus in the Lung of Mice by Regulating Cytokines and Nitric Oxide Production. Front Immunol. 2017;8:1477.
  63. 63. Bickett TE, McLean J, Creissen E, Izzo L, Hagan C, Izzo AJ, et al. Characterizing the BCG Induced Macrophage and Neutrophil Mechanisms for Defense Against Mycobacterium tuberculosis. Front Immunol. 2020;11:1202.
  64. 64. Dubé J-Y, McIntosh F, Zarruk JG, David S, Nigou J, Behr MA. Synthetic mycobacterial molecular patterns partially complete Freund’s adjuvant. Sci Rep. 2020;10:5874. pmid:32246076
  65. 65. Khan A, Singh VK, Mishra A, Soudani E, Bakhru P, Singh CR, et al. NOD2/RIG-I Activating Inarigivir Adjuvant Enhances the Efficacy of BCG Vaccine Against Tuberculosis in Mice. Front Immunol. 2020;11:592333. pmid:33365029
  66. 66. Lu Q, Zhang W, Fang J, Zheng J, Dong C, Xiong S. Mycobacterium tuberculosis Rv1096, facilitates mycobacterial survival by modulating the NF-κB/MAPK pathway as peptidoglycan N-deacetylase. Mol Immunol. 2020;127:47–55.
  67. 67. Napier RJ, Lee EJ, Davey MP, Vance EE, Furtado JM, Snow PE, et al. T cell-intrinsic role for Nod2 in protection against Th17-mediated uveitis. Nat Commun. 2020;11:5406. pmid:33106495
  68. 68. Ahn JH, Park JY, Kim DY, Lee TS, Jung DH, Kim YJ, et al. Type I Interferons Are Involved in the Intracellular Growth Control of Mycobacterium abscessus by Mediating NOD2-Induced Production of Nitric Oxide in Macrophages. Front Immunol. 2021;12:738070.
  69. 69. Aqdas M, Maurya SK, Pahari S, Singh S, Khan N, Sethi K, et al. Immunotherapeutic Role of NOD-2 and TLR-4 Signaling as an Adjunct to Antituberculosis Chemotherapy. ACS Infect Dis. 2021;7:2999–3008. pmid:34613696
  70. 70. Aqdas M, Singh S, Amir M, Maurya SK, Pahari S, Agrewala JN. Cumulative Signaling Through NOD-2 and TLR-4 Eliminates the Mycobacterium Tuberculosis Concealed Inside the Mesenchymal Stem Cells. Front Cell Infect Microbiol. 2021;11:669168.
  71. 71. Correa-Macedo W, Fava VM, Orlova M, Cassart P, Olivenstein R, Sanz J, et al. Alveolar macrophages from persons living with HIV show impaired epigenetic response to Mycobacterium tuberculosis. J Clin Invest. 2021:131.
  72. 72. Mandala JP, Thada S, Sivangala R, Ponnana M, Myakala R, Gaddam S. Influence of NOD-like receptor 2 gene polymorphisms on muramyl dipeptide induced pro-inflammatory response in patients with active pulmonary tuberculosis and household contacts. Immunobiology. 2021;226:152096. pmid:34058448
  73. 73. Dubé J-Y, McIntosh F, Behr MA. Mice Dually Disrupted for Nod2 and Mincle Manifest Early Bacteriological Control but Late Susceptibility During Mycobacterium tuberculosis Infection. Front Immunol. 2022;13.
  74. 74. Dallmann-Sauer M, Xu YZ, da Costa ALF, Tao S, Gomes TA, Prata RBS, et al. Allele-dependent interaction of LRRK2 and NOD2 in leprosy. PLoS Pathog. 2023;19:e1011260. pmid:36972292
  75. 75. Mo J, Boyle JP, Howard CB, Monie TP, Davis BK, Duncan JA. Pathogen sensing by nucleotide-binding oligomerization domain-containing protein 2 (NOD2) is mediated by direct binding to muramyl dipeptide and ATP. J Biol Chem. 2012;287:23057–23067. pmid:22549783
  76. 76. Lauro ML D’Ambrosio EA Bahnson BJ, Grimes CL. Molecular Recognition of Muramyl Dipeptide Occurs in the Leucine-rich Repeat Domain of Nod2. ACS Infect Dis. 2017;3:264–270. pmid:27748583
  77. 77. Tanabe T, Chamaillard M, Ogura Y, Zhu L, Qiu S, Masumoto J, et al. Regulatory regions and critical residues of NOD2 involved in muramyl dipeptide recognition. EMBO J. 2004;23:1587–1597. pmid:15044951
  78. 78. Jumper J, Evans R, Pritzel A, Green T, Figurnov M, Ronneberger O, et al. Highly accurate protein structure prediction with AlphaFold. Nature. 2021;596:583–589. pmid:34265844
  79. 79. Pettersen EF, Goddard TD, Huang CC, Meng EC, Couch GS, Croll TI, et al. UCSF ChimeraX: Structure visualization for researchers, educators, and developers. Protein Sci. 2021;30:70–82. pmid:32881101
  80. 80. Davis KM, Nakamura S, Weiser JN. Nod2 sensing of lysozyme-digested peptidoglycan promotes macrophage recruitment and clearance of S. pneumoniae colonization in mice. J Clin Invest. 2011;121:3666–3676.
  81. 81. Nakamura N, Lill JR, Phung Q, Jiang Z, Bakalarski C, de Maziere A, et al. Endosomes are specialized platforms for bacterial sensing and NOD2 signalling. Nature. 2014;509:240–244. pmid:24695226
  82. 82. Ismair MG, Vavricka SR, Kullak-Ublick GA, Fried M, Mengin-Lecreulx D, Girardin SE. hPepT1 selectively transports muramyl dipeptide but not Nod1-activating muramyl peptides. Can J Physiol Pharmacol. 2006;84:1313–1319. pmid:17487240
  83. 83. Charrière GM, Ip WE, Dejardin S, Boyer L, Sokolovska A, Cappillino MP, et al. Identification of Drosophila Yin and PEPT2 as evolutionarily conserved phagosome-associated muramyl dipeptide transporters. J Biol Chem. 2010;285:20147–20154.
  84. 84. Schaefer AK, Melnyk JE, Baksh MM, Lazor KM, Finn MG, Grimes CL. Membrane Association Dictates Ligand Specificity for the Innate Immune Receptor NOD2. ACS Chem Biol. 2017;12:2216–2224. pmid:28708377
  85. 85. Dixon CL, Fairn GD. S-palmitoylation of NOD2 controls its localization to the plasma membrane. J Lipid Res. 2021;62.
  86. 86. Lu Y, Zheng Y, Coyaud É, Zhang C, Selvabaskaran A, Yu Y, et al. Palmitoylation of NOD1 and NOD2 is required for bacterial sensing. Science. 2019;366:460–467. pmid:31649195
  87. 87. Stafford CA, Gassauer A-M, de Oliveira Mann CC, Tanzer MC, Fessler E, Wefers B, et al. Phosphorylation of muramyl peptides by NAGK is required for NOD2 activation. Nature. 2022. pmid:36002575
  88. 88. Grimes CL, Ariyananda Lde Z, Melnyk JE, O’Shea EK. The innate immune protein Nod2 binds directly to MDP, a bacterial cell wall fragment. J Am Chem Soc. 2012;134:13535–13537. pmid:22857257
  89. 89. Mukherjee T, Hovingh ES, Foerster EG, Abdel-Nour M, Philpott DJ, Girardin SE. NOD1 and NOD2 in inflammation, immunity and disease. Arch Biochem Biophys. 2019;670:69–81. pmid:30578751
  90. 90. Trindade BC, Chen GY. NOD1 and NOD2 in inflammatory and infectious diseases. Immunol Rev. 2020;297:139–161. pmid:32677123
  91. 91. Cuthbert AP, Fisher SA, Mirza MM, King K, Hampe J, Croucher PJ, et al. The contribution of NOD2 gene mutations to the risk and site of disease in inflammatory bowel disease. Gastroenterology. 2002;122:867–874.
  92. 92. Lesage S, Zouali H, Cézard JP, Colombel JF, Belaiche J, Almer S, et al. CARD15/NOD2 mutational analysis and genotype-phenotype correlation in 612 patients with inflammatory bowel disease. Am J Hum Genet. 2002;70:845–857.
  93. 93. Dalziel TK. Chronic Interstitial Enteritis. Br Med J. 1913;2:1068–1070.
  94. 94. Behr MA, Semret M, Poon A, Schurr E. Crohn’s disease, mycobacteria, and NOD2. Lancet Infect Dis. 2004;4:136–137. pmid:14998497
  95. 95. Sechi LA, Mura M, Tanda E, Lissia A, Fadda G, Zanetti S. Mycobacterium avium sub. paratuberculosis in tissue samples of Crohn’s disease patients. New Microbiol. 2004;27:75–77.
  96. 96. Sabbah A, Chang TH, Harnack R, Frohlich V, Tominaga K, Dube PH, et al. Activation of innate immune antiviral responses by NOD2. Nat Immunol. 2009;10:1073–1080. pmid:19701189
  97. 97. Morosky SA, Zhu J, Mukherjee A, Sarkar SN, Coyne CB. Retinoic acid-induced gene-I (RIG-I) associates with nucleotide-binding oligomerization domain-2 (NOD2) to negatively regulate inflammatory signaling. J Biol Chem. 2011;286:28574–28583. pmid:21690088
  98. 98. Keestra-Gounder AM, Byndloss MX, Seyffert N, Young BM, Chávez-Arroyo A, Tsai AY, et al. NOD1 and NOD2 signalling links ER stress with inflammation. Nature. 2016;532:394–397. pmid:27007849
  99. 99. Singh K, Han K, Tilve S, Wu K, Geller HM, Sack MN. Parkin targets NOD2 to regulate astrocyte endoplasmic reticulum stress and inflammation. Glia. 2018;66:2427–2437. pmid:30378174
  100. 100. Pham OH, Lee B, Labuda J, Keestra-Gounder AM, Byndloss MX, Tsolis RM, et al. NOD1/NOD2 and RIP2 Regulate Endoplasmic Reticulum Stress-Induced Inflammation during Chlamydia Infection. mBio. 2020;11(3):e00979–20. pmid:32487756
  101. 101. Fritz T, Niederreiter L, Adolph T, Blumberg RS, Kaser A. Crohn’s disease: NOD2, autophagy and ER stress converge. Gut. 2011;60:1580–1588. pmid:21252204
  102. 102. Lécine P, Esmiol S, Métais JY, Nicoletti C, Nourry C, McDonald C, et al. The NOD2-RICK complex signals from the plasma membrane. J Biol Chem. 2007;282:15197–15207. pmid:17355968
  103. 103. Ogura Y, Inohara N, Benito A, Chen FF, Yamaoka S, Nunez G. Nod2, a Nod1/Apaf-1 family member that is restricted to monocytes and activates NF-kappaB. J Biol Chem. 2001;276:4812–4818. pmid:11087742
  104. 104. Park JH, Kim YG, McDonald C, Kanneganti TD, Hasegawa M, Body-Malapel M, et al. RICK/RIP2 mediates innate immune responses induced through Nod1 and Nod2 but not TLRs. J Immunol. 2007;178:2380–2386. pmid:17277144
  105. 105. Zhang FR, Huang W, Chen SM, Sun LD, Liu H, Li Y, et al. Genomewide association study of leprosy. N Engl J Med. 2009;361:2609–2618. pmid:20018961
  106. 106. Berrington WR, Macdonald M, Khadge S, Sapkota BR, Janer M, Hagge DA, et al. Common polymorphisms in the NOD2 gene region are associated with leprosy and its reactive states. J Infect Dis. 2010;201:1422–1435.
  107. 107. Leturiondo AL, Noronha AB, Mendonça CYR, Ferreira CO, Alvarado-Arnez LE, Manta FSN, et al. Association of NOD2 and IFNG single nucleotide polymorphisms with leprosy in the Amazon ethnic admixed population. PLoS Negl Trop Dis. 2020;14:e0008247.
  108. 108. Sales-Marques C, Salomão H, Fava VM, Alvarado-Arnez LE, Amaral EP, Cardoso CC, et al. NOD2 and CCDC122-LACC1 genes are associated with leprosy susceptibility in Brazilians. Hum Genet. 2014;133:1525–1532. pmid:25367361
  109. 109. Sales-Marques C, Cardoso CC, Alvarado-Arnez LE, Illaramendi X, Sales AM, Hacker MA, et al. Genetic polymorphisms of the IL6 and NOD2 genes are risk factors for inflammatory reactions in leprosy. PLoS Negl Trop Dis. 2017;11:e0005754.
  110. 110. Zhang Q, Pan Y, Yan R, Zeng B, Wang H, Zhang X, et al. Commensal bacteria direct selective cargo sorting to promote symbiosis. Nat Immunol. 2015;16:918–926. pmid:26237551
  111. 111. Yan R, Liu Z. LRRK2 enhances Nod1/2-mediated inflammatory cytokine production by promoting Rip2 phosphorylation. Protein Cell. 2017;8:55–66. pmid:27830463
  112. 112. Wang H, Zhang X, Zuo Z, Zhang Q, Pan Y, Zeng B, et al. Rip2 Is Required for Nod2-Mediated Lysozyme Sorting in Paneth Cells. J Immunol. 2017;198:3729–3736. pmid:28330897
  113. 113. Barrett JC, Hansoul S, Nicolae DL, Cho JH, Duerr RH, Rioux JD, et al. Genome-wide association defines more than 30 distinct susceptibility loci for Crohn’s disease. Nat Genet. 2008;40:955–962. pmid:18587394
  114. 114. Fava VM, Manry J, Cobat A, Orlova M, Van Thuc N, Ba NN, et al. A Missense LRRK2 Variant Is a Risk Factor for Excessive Inflammatory Responses in Leprosy. PLoS Negl Trop Dis. 2016;10:e0004412.
  115. 115. Liu TC, Naito T, Liu Z, VanDussen KL, Haritunians T, Li D, et al. LRRK2 but not ATG16L1 is associated with Paneth cell defect in Japanese Crohn’s disease patients. JCI Insight. 2017;2:e91917.
  116. 116. Hui KY, Fernandez-Hernandez H, Hu J, Schaffner A, Pankratz N, Hsu NY, et al. Functional variants in the LRRK2 gene confer shared effects on risk for Crohn’s disease and Parkinson’s disease. Sci Transl Med. 2018;10(423):eaai7795.
  117. 117. Rivas MA, Avila BE, Koskela J, Huang H, Stevens C, Pirinen M, et al. Insights into the genetic epidemiology of Crohn’s and rare diseases in the Ashkenazi Jewish population. PLoS Genet. 2018;14:e1007329. pmid:29795570
  118. 118. Hampe J, Franke A, Rosenstiel P, Till A, Teuber M, Huse K, et al. A genome-wide association scan of nonsynonymous SNPs identifies a susceptibility variant for Crohn disease in ATG16L1. Nat Genet. 2007;39:207–211.
  119. 119. Rioux JD, Xavier RJ, Taylor KD, Silverberg MS, Goyette P, Huett A, et al. Genome-wide association study identifies new susceptibility loci for Crohn disease and implicates autophagy in disease pathogenesis. Nat Genet. 2007;39:596–604. pmid:17435756
  120. 120. Austin CM, Ma X, Graviss EA. Common nonsynonymous polymorphisms in the NOD2 gene are associated with resistance or susceptibility to tuberculosis disease in African Americans. J Infect Dis. 2008;197:1713–1716.
  121. 121. Hall NB, Igo RP Jr, Malone LL, Truitt B, Schnell A, Tao L, et al. Polymorphisms in TICAM2 and IL1B are associated with TB. Genes Immun. 2015;16:127–133.
  122. 122. Gutierrez J, Kroon EE, Möller M, Stein CM. Phenotype Definition for “Resisters” to Mycobacterium tuberculosis Infection in the Literature—A Review and Recommendations. Front Immunol. 2021;12.
  123. 123. Schurz H, Glanzmann B, Bowker N, van Toorn R, Solomons R, Schoeman J, et al. Deciphering Genetic Susceptibility to Tuberculous Meningitis. Front Neurol. 2022;13:820168. pmid:35401413
  124. 124. Capela C, Dossou AD, Silva-Gomes R, Sopoh GE, Makoutode M, Menino JF, et al. Genetic Variation in Autophagy-Related Genes Influences the Risk and Phenotype of Buruli Ulcer. PLoS Negl Trop Dis. 2016;10:e0004671. pmid:27128681
  125. 125. Pinedo PJ, Buergelt CD, Donovan GA, Melendez P, Morel L, Wu R, et al. Association between CARD15/NOD2 gene polymorphisms and paratuberculosis infection in cattle. Vet Microbiol. 2009;134:346–352.
  126. 126. Gonzalez-Mancera MS, Forghani I, Mirsaeidi M. Missense (p.Glu778Lys) and (p.Gly908Arg) variants of NOD2 gene are associated with recurrent pulmonary non-tuberculous mycobacterial infections. Scand J Immunol. 2020;92:e12935. pmid:32654169
  127. 127. Vincent AT, Nyongesa S, Morneau I, Reed MB, Tocheva EI, Veyrier FJ. The Mycobacterial Cell Envelope: A Relict From the Past or the Result of Recent Evolution? Front Microbiol. 2018;9. pmid:30369911
  128. 128. Dulberger CL, Rubin EJ, Boutte CC. The mycobacterial cell envelope—a moving target. Nat Rev Microbiol. 2020;18:47–59. pmid:31728063
  129. 129. Kalscheuer R, Palacios A, Anso I, Cifuente J, Anguita J, Jacobs WR Jr, et al. The Mycobacterium tuberculosis capsule: a cell structure with key implications in pathogenesis. Biochem J. 2019;476:1995–2016.
  130. 130. Lemassu A, Ortalo-Magné A, Bardou F, Silve G, Lanéelle MA, Daffé M. Extracellular and surface-exposed polysaccharides of non-tuberculous mycobacteria. Microbiology (Reading). 1996;142 (Pt 6):1513–1520. pmid:8704991
  131. 131. Abrahams KA, Besra GS. Mycobacterial cell wall biosynthesis: a multifaceted antibiotic target. Parasitology. 2018;145:116–133. pmid:27976597
  132. 132. Essers L, Schoop HJ. Evidence for the incorporation of molecular oxygen, a pathway in biosynthesis of N-glycolylmuramic acid in Mycobacterium phlei. Biochim Biophys Acta. 1978;544:180–184.
  133. 133. Raymond JB, Mahapatra S, Crick DC, Pavelka MS Jr. Identification of the namH gene, encoding the hydroxylase responsible for the N-glycolylation of the mycobacterial peptidoglycan. J Biol Chem. 2005;280:326–333. pmid:15522883
  134. 134. Mahapatra S, Crick DC, McNeil MR, Brennan PJ. Unique structural features of the peptidoglycan of Mycobacterium leprae. J Bacteriol. 2008;190:655–661.
  135. 135. Azuma I, Thomas DW, Adam A, Ghuysen JM, Bonaly R, Petit JF, et al. Occurrence of N-glycolylmuramic acid in bacterial cell walls. A preliminary survey. Biochim Biophys Acta. 1970;208:444–451.
  136. 136. Menardo F, Duchêne S, Brites D, Gagneux S. The molecular clock of Mycobacterium tuberculosis. PLoS Pathog. 2019;15:e1008067.
  137. 137. Sarathy JP, Dartois V. Caseum: a Niche for Mycobacterium tuberculosis Drug-Tolerant Persisters. Clin Microbiol Rev. 2020;33.
  138. 138. Du P, Sohaskey CD, Shi L. Transcriptional and Physiological Changes during Mycobacterium tuberculosis Reactivation from Non-replicating Persistence. Front Microbiol. 2016;7:1346.
  139. 139. Wolf AJ, Reyes CN, Liang W, Becker C, Shimada K, Wheeler ML, et al. Hexokinase Is an Innate Immune Receptor for the Detection of Bacterial Peptidoglycan. Cell. 2016;166:624–636. pmid:27374331
  140. 140. Beckwith KS, Beckwith MS, Ullmann S, Sætra RS, Kim H, Marstad A, et al. Plasma membrane damage causes NLRP3 activation and pyroptosis during Mycobacterium tuberculosis infection. Nat Commun. 2020;11:2270.
  141. 141. Qu Z, Zhou J, Zhou Y, Xie Y, Jiang Y, Wu J, et al. Mycobacterial EST12 activates a RACK1-NLRP3-gasdermin D pyroptosis-IL-1β immune pathway. Sci Adv. 2020;6.
  142. 142. McElvania Tekippe E, Allen IC, Hulseberg PD, Sullivan JT, McCann JR, Sandor M, et al. Granuloma formation and host defense in chronic Mycobacterium tuberculosis infection requires PYCARD/ASC but not NLRP3 or caspase-1. PLoS ONE. 2010;5:e12320.
  143. 143. Faustin B, Lartigue L, Bruey J-M, Luciano F, Sergienko E, Bailly-Maitre B, et al. Reconstituted NALP1 Inflammasome Reveals Two-Step Mechanism of Caspase-1 Activation. Mol Cell. 2007;25:713–724. pmid:17349957
  144. 144. Reubold TF, Hahne G, Wohlgemuth S, Eschenburg S. Crystal structure of the leucine-rich repeat domain of the NOD-like receptor NLRP1: implications for binding of muramyl dipeptide. FEBS Lett. 2014;588:3327–3332. pmid:25064844
  145. 145. Guan R, Roychowdhury A, Ember B, Kumar S, Boons GJ, Mariuzza RA. Structural basis for peptidoglycan binding by peptidoglycan recognition proteins. Proc Natl Acad Sci U S A. 2004;101:17168–17173. pmid:15572450
  146. 146. Guan R, Brown PH, Swaminathan CP, Roychowdhury A, Boons GJ, Mariuzza RA. Crystal structure of human peptidoglycan recognition protein I alpha bound to a muramyl pentapeptide from Gram-positive bacteria. Protein Sci. 2006;15:1199–1206. pmid:16641493
  147. 147. Sharma P, Dube D, Sinha M, Mishra B, Dey S, Mal G, et al. Multiligand specificity of pathogen-associated molecular pattern-binding site in peptidoglycan recognition protein. J Biol Chem. 2011;286:31723–31730. pmid:21784863
  148. 148. Duerr CU, Salzman NH, Dupont A, Szabo A, Normark BH, Normark S, et al. Control of intestinal Nod2-mediated peptidoglycan recognition by epithelium-associated lymphocytes. Mucosal Immunol. 2011;4:325–334. pmid:20980996
  149. 149. Melnyk JE, Mohanan V, Schaefer AK, Hou CW, Grimes CL. Peptidoglycan Modifications Tune the Stability and Function of the Innate Immune Receptor Nod2. J Am Chem Soc. 2015;137:6987–6990. pmid:26035228
  150. 150. Cole ST, Eiglmeier K, Parkhill J, James KD, Thomson NR, Wheeler PR, et al. Massive gene decay in the leprosy bacillus. Nature. 2001;409:1007–1011. pmid:11234002
  151. 151. Drobniewski FA, Pozniak AL, Uttley AH. Tuberculosis and AIDS. J Med Microbiol. 1995;43:85–91. pmid:7629858
  152. 152. Ehlers S, Schaible UE. The granuloma in tuberculosis: dynamics of a host-pathogen collusion. Front Immunol. 2013;3:411–411. pmid:23308075
  153. 153. Wang Q, Matsuo Y, Pradipta AR, Inohara N, Fujimoto Y, Fukase K. Synthesis of characteristic Mycobacterium peptidoglycan (PGN) fragments utilizing with chemoenzymatic preparation of meso-diaminopimelic acid (DAP), and their modulation of innate immune responses. Org Biomol Chem. 2016;14:1013–1023.
  154. 154. Chen KT, Huang DY, Chiu CH, Lin WW, Liang PH, Cheng WC. Synthesis of Diverse N-Substituted Muramyl Dipeptide Derivatives and Their Use in a Study of Human NOD2 Stimulation Activity. Chemistry (Easton). 2015;21:11984–11988.
  155. 155. Bersch KL, DeMeester KE, Zagani R, Chen S, Wodzanowski KA, Liu S, et al. Bacterial Peptidoglycan Fragments Differentially Regulate Innate Immune Signaling. ACS Cent Sci. 2021;7:688–696. pmid:34056099
  156. 156. Rubino SJ, Magalhaes JG, Philpott D, Bahr GM, Blanot D, Girardin SE. Identification of a synthetic muramyl peptide derivative with enhanced Nod2 stimulatory capacity. Innate Immun. 2013;19:493–503. pmid:23339926
  157. 157. Lachmandas E, Thiem K, van den Heuvel C, Hijmans A, de Galan BE, Tack CJ, et al. Patients with type 1 diabetes mellitus have impaired IL-1β production in response to Mycobacterium tuberculosis. Eur J Clin Microbiol Infect Dis. 2018;37:371–380.
  158. 158. Yang Y, Yin C, Pandey A, Abbott D, Sassetti C, Kelliher MA. NOD2 pathway activation by MDP or Mycobacterium tuberculosis infection involves the stable polyubiquitination of Rip2. J Biol Chem. 2007;282:36223–36229.
  159. 159. Tigno-Aranjuez JT, Asara JM, Abbott DW. Inhibition of RIP2’s tyrosine kinase activity limits NOD2-driven cytokine responses. Genes Dev. 2010;24:2666–2677. pmid:21123652
  160. 160. Nembrini C, Kisielow J, Shamshiev AT, Tortola L, Coyle AJ, Kopf M, et al. The kinase activity of Rip2 determines its stability and consequently Nod1- and Nod2-mediated immune responses. J Biol Chem. 2009;284:19183–19188. pmid:19473975
  161. 161. Flynn JL, Goldstein MM, Chan J, Triebold KJ, Pfeffer K, Lowenstein CJ, et al. Tumor necrosis factor-alpha is required in the protective immune response against Mycobacterium tuberculosis in mice. Immunity. 1995;2:561–572.
  162. 162. Drennan MB, Nicolle D, Quesniaux VJ, Jacobs M, Allie N, Mpagi J, et al. Toll-like receptor 2-deficient mice succumb to Mycobacterium tuberculosis infection. Am J Pathol. 2004;164:49–57.
  163. 163. Fremond CM, Yeremeev V, Nicolle DM, Jacobs M, Quesniaux VF, Ryffel B. Fatal Mycobacterium tuberculosis infection despite adaptive immune response in the absence of MyD88. J Clin Invest. 2004;114:1790–1799.
  164. 164. Bourigault ML, Segueni N, Rose S, Court N, Vacher R, Vasseur V, et al. Relative contribution of IL-1α, IL-1β and TNF to the host response to Mycobacterium tuberculosis and attenuated M. bovis BCG. Immun Inflamm Dis. 2013;1:47–62.
  165. 165. Keane J, Gershon S, Wise RP, Mirabile-Levens E, Kasznica J, Schwieterman WD, et al. Tuberculosis associated with infliximab, a tumor necrosis factor alpha-neutralizing agent. N Engl J Med. 2001;345:1098–1104. pmid:11596589
  166. 166. Cooper AM, Dalton DK, Stewart TA, Griffin JP, Russell DG, Orme IM. Disseminated tuberculosis in interferon gamma gene-disrupted mice. J Exp Med. 1993;178:2243–2247. pmid:8245795
  167. 167. Flynn JL, Chan J, Triebold KJ, Dalton DK, Stewart TA, Bloom BR. An essential role for interferon gamma in resistance to Mycobacterium tuberculosis infection. J Exp Med. 1993;178:2249–2254.
  168. 168. Sakai S, Kauffman KD, Sallin MA, Sharpe AH, Young HA, Ganusov VV, et al. CD4 T Cell-Derived IFN-gamma Plays a Minimal Role in Control of Pulmonary Mycobacterium tuberculosis Infection and Must Be Actively Repressed by PD-1 to Prevent Lethal Disease. PLoS Pathog. 2016;12:e1005667.
  169. 169. Cooper AM, Magram J, Ferrante J, Orme IM. Interleukin 12 (IL-12) is crucial to the development of protective immunity in mice intravenously infected with Mycobacterium tuberculosis. J Exp Med. 1997;186:39–45.
  170. 170. Döffinger R, Dupuis S, Picard C, Fieschi C, Feinberg J, Barcenas-Morales G, et al. Inherited disorders of IL-12- and IFNgamma-mediated immunity: a molecular genetics update. Mol Immunol. 2002;38:903–909. pmid:12009568
  171. 171. Ji DX, Yamashiro LH, Chen KJ, Mukaida N, Kramnik I, Darwin KH, et al. Type I interferon-driven susceptibility to Mycobacterium tuberculosis is mediated by IL-1Ra. Nat Microbiol. 2019;4:2128–2135.
  172. 172. Berry MP, Graham CM, McNab FW, Xu Z, Bloch SA, Oni T, et al. An interferon-inducible neutrophil-driven blood transcriptional signature in human tuberculosis. Nature. 2010;466:973–977. pmid:20725040
  173. 173. Moreira-Teixeira L, Mayer-Barber K, Sher A, O’Garra A. Type I interferons in tuberculosis: Foe and occasionally friend. J Exp Med. 2018;215:1273–1285. pmid:29666166
  174. 174. Juffermans NP, Florquin S, Camoglio L, Verbon A, Kolk AH, Speelman P, et al. Interleukin-1 signaling is essential for host defense during murine pulmonary tuberculosis. J Infect Dis. 2000;182:902–908. pmid:10950787
  175. 175. Sugawara I, Yamada H, Hua S, Mizuno S. Role of interleukin (IL)-1 type 1 receptor in mycobacterial infection. Microbiol Immunol. 2001;45:743–750. pmid:11791667
  176. 176. Fremond CM, Togbe D, Doz E, Rose S, Vasseur V, Maillet I, et al. IL-1 receptor-mediated signal is an essential component of MyD88-dependent innate response to Mycobacterium tuberculosis infection. J Immunol. 2007;179:1178–1189.
  177. 177. Yamada H, Mizumo S, Horai R, Iwakura Y, Sugawara I. Protective Role of Interleukin-1 in Mycobacterial Infection in IL-1 α/β Double-Knockout Mice. Lab Investig. 2000;80:759–767.
  178. 178. Mayer-Barber KD, Barber DL, Shenderov K, White SD, Wilson MS, Cheever A, et al. Cutting Edge: Caspase-1 Independent IL-1β Production Is Critical for Host Resistance to Mycobacterium tuberculosis and Does Not Require TLR Signaling In Vivo. J Immunol. 2010;184:3326–3330.
  179. 179. Redford PS, Murray PJ, O’Garra A. The role of IL-10 in immune regulation during M. tuberculosis infection. Mucosal Immunol. 2011;4:261–270.
  180. 180. Netea MG, Azam T, Lewis EC, Joosten LA, Wang M, Langenberg D, et al. Mycobacterium tuberculosis induces interleukin-32 production through a caspase- 1/IL-18/interferon-gamma-dependent mechanism. PLoS Med. 2006;3:e277.
  181. 181. Bai X, Kim SH, Azam T, McGibney MT, Huang H, Dinarello CA, et al. IL-32 is a host protective cytokine against Mycobacterium tuberculosis in differentiated THP-1 human macrophages. J Immunol. 2010;184:3830–3840.
  182. 182. Bai X, Shang S, Henao-Tamayo M, Basaraba RJ, Ovrutsky AR, Matsuda JL, et al. Human IL-32 expression protects mice against a hypervirulent strain of Mycobacterium tuberculosis. Proc Natl Acad Sci. 2015;112:5111–5116.
  183. 183. MacMicking JD, North RJ, LaCourse R, Mudgett JS, Shah SK, Nathan CF. Identification of nitric oxide synthase as a protective locus against tuberculosis. Proc Natl Acad Sci. 1997;94:5243–5248. pmid:9144222
  184. 184. Jung YJ, LaCourse R, Ryan L, North RJ. Virulent but not avirulent Mycobacterium tuberculosis can evade the growth inhibitory action of a T helper 1-dependent, nitric oxide Synthase 2-independent defense in mice. J Exp Med. 2002;196:991–998.
  185. 185. Scanga CA, Mohan VP, Tanaka K, Alland D, Flynn JL, Chan J. The inducible nitric oxide synthase locus confers protection against aerogenic challenge of both clinical and laboratory strains of Mycobacterium tuberculosis in mice. Infect Immun. 2001;69:7711–7717.
  186. 186. Behar SM, Martin CJ, Booty MG, Nishimura T, Zhao X, Gan HX, et al. Apoptosis is an innate defense function of macrophages against Mycobacterium tuberculosis. Mucosal Immunol. 2011;4:279–287.
  187. 187. Watson RO, Manzanillo PS, Cox JS. Extracellular M. tuberculosis DNA targets bacteria for autophagy by activating the host DNA-sensing pathway. Cell. 2012;150:803–815.
  188. 188. Travassos LH, Carneiro LA, Ramjeet M, Hussey S, Kim YG, Magalhães JG, et al. Nod1 and Nod2 direct autophagy by recruiting ATG16L1 to the plasma membrane at the site of bacterial entry. Nat Immunol. 2010;11:55–62. pmid:19898471
  189. 189. Cooney R, Baker J, Brain O, Danis B, Pichulik T, Allan P, et al. NOD2 stimulation induces autophagy in dendritic cells influencing bacterial handling and antigen presentation. Nat Med. 2010;16:90–97. pmid:19966812
  190. 190. Kramnik I, Dietrich WF, Demant P, Bloom BR. Genetic control of resistance to experimental infection with virulent Mycobacterium tuberculosis. Proc Natl Acad Sci U S A. 2000;97:8560–8565.
  191. 191. Driver ER, Ryan GJ, Hoff DR, Irwin SM, Basaraba RJ, Kramnik I, et al. Evaluation of a mouse model of necrotic granuloma formation using C3HeB/FeJ mice for testing of drugs against Mycobacterium tuberculosis. Antimicrob Agents Chemother. 2012;56:3181–3195.
  192. 192. Adam A, Petit JF, Wietzerbin-Falszpan J, Sinay P, Thomas DW, Lederer E. L’acide N-glycolyl-muramique, constituant des parois de Mycobacterium smegmatis: Identification par spectrometrie de masse. FEBS Lett. 1969;4:87–92.
  193. 193. Petit JF, Adam A, Wietzerbin-Falszpan J, Lederer E, Ghuysen JM. Chemical structure of the cell wall of Mycobacterium smegmatis. I. Isolation and partial characterization of the peptidoglycan. Biochem Biophys Res Commun. 1969;35:478–485.
  194. 194. Petit J, Adam A, Wietzerbin-Falszpan J. Isolation of UDP-N-glycolylmuramyl-(Ala, Glu, DAP) from Mycobacterium phlei. FEBS Lett. 1970;6:55–57.
  195. 195. Takayama K, David HL, Wang L, Goldman DS. Isolation and characterization of uridine diphosphate-N-glycolylmuramyl-L-alanyl-gamma-D-glutamyl-meso-alpha,alpha’-diaminopimelic acid from Mycobacterium tuberculosis. Biochem Biophys Res Commun. 1970;39:7–12.
  196. 196. Ellouz F, Adam A, Ciorbaru R, Lederer E. Minimal structural requirements for adjuvant activity of bacterial peptidoglycan derivatives. Biochem Biophys Res Commun. 1974;59:1317–1325. pmid:4606813
  197. 197. Qazi KR, Oehlmann W, Singh M, López MC, Fernández C. Microbial heat shock protein 70 stimulatory properties have different TLR requirements. Vaccine. 2007;25:1096–1103. pmid:17049413
  198. 198. Coulombe F, Fiola S, Akira S, Cormier Y, Gosselin J. Muramyl Dipeptide Induces NOD2-Dependent Ly6Chigh Monocyte Recruitment to the Lungs and Protects Against Influenza Virus Infection. PLoS ONE. 2012;7:e36734. pmid:22590599
  199. 199. Kaufmann E, Sanz J, Dunn JL, Khan N, Mendonça LE, Pacis A, et al. BCG Educates Hematopoietic Stem Cells to Generate Protective Innate Immunity against Tuberculosis. Cell. 2018;172:176–190.e19. pmid:29328912
  200. 200. van Puffelen JH, Keating ST, Oosterwijk E, van der Heijden AG, Netea MG, Joosten LAB, et al. Trained immunity as a molecular mechanism for BCG immunotherapy in bladder cancer. Nat Rev Urol. 2020;17:513–525. pmid:32678343
  201. 201. Sur PK, Dastidar AG. Role of mycobacterium w as adjuvant treatment of lung cancer (non-small cell lung cancer). J Indian Med Assoc. 2003;101(118):120.
  202. 202. Subramaniam M, Arshad NM, Mun KS, Malagobadan S, Awang K, Nagoor NH. Anti-Cancer Effects of Synergistic Drug–Bacterium Combinations on Induced Breast Cancer in BALB/c Mice. Biomol Ther. 2019;9:626. pmid:31635311
  203. 203. Sugiyama T, Fujiwara K, Ohashi Y, Yokota H, Hatae M, Ohno T, et al. Phase III placebo-controlled double-blind randomized trial of radiotherapy for stage IIB-IVA cervical cancer with or without immunomodulator Z-100: a JGOG study. Ann Oncol. 2014;25:1011–1017. pmid:24569914