Skip to main content
Advertisement
  • Loading metrics

A protective role for type I interferon signaling following infection with Mycobacterium tuberculosis carrying the rifampicin drug resistance-conferring RpoB mutation H445Y

  • Suhas Bobba,

    Roles Conceptualization, Formal analysis, Investigation, Methodology, Validation, Visualization, Writing – original draft, Writing – review & editing

    Affiliations Department of Molecular Microbiology, Washington University School of Medicine, St. Louis, Missouri, United States of America, Division of Infectious Diseases, Department of Medicine, Washington University School of Medicine, St. Louis, Missouri, United States of America

  • Kuldeep S. Chauhan,

    Roles Conceptualization, Formal analysis, Investigation, Validation, Writing – review & editing

    Affiliation Department of Microbiology, University of Chicago, Chicago, Illinois, United States of America

  • Sadia Akter,

    Roles Data curation, Formal analysis, Software, Visualization, Writing – review & editing

    Affiliation Department of Microbiology, University of Chicago, Chicago, Illinois, United States of America

  • Shibali Das,

    Roles Investigation, Methodology, Validation, Writing – review & editing

    Affiliation Department of Molecular Microbiology, Washington University School of Medicine, St. Louis, Missouri, United States of America

  • Ekansh Mittal,

    Roles Investigation, Methodology, Validation, Writing – review & editing

    Affiliation Division of Infectious Diseases, Department of Medicine, Washington University School of Medicine, St. Louis, Missouri, United States of America

  • Barun Mathema,

    Roles Conceptualization, Supervision, Writing – review & editing

    Affiliation Department of Epidemiology, Columbia University Mailman School of Public Health, New York, New York, United States of America

  • Jennifer A. Philips,

    Roles Funding acquisition, Resources, Supervision, Writing – review & editing

    Affiliations Department of Molecular Microbiology, Washington University School of Medicine, St. Louis, Missouri, United States of America, Division of Infectious Diseases, Department of Medicine, Washington University School of Medicine, St. Louis, Missouri, United States of America

  • Shabaana A. Khader

    Roles Conceptualization, Funding acquisition, Project administration, Resources, Supervision, Writing – original draft, Writing – review & editing

    khader@uchicago.edu

    Affiliation Department of Microbiology, University of Chicago, Chicago, Illinois, United States of America

Abstract

Interleukin-1 (IL-1) signaling is essential for controlling virulent Mycobacterium tuberculosis (Mtb) infection since antagonism of this pathway leads to exacerbated pathology and increased susceptibility. In contrast, the triggering of type I interferon (IFN) signaling is associated with the progression of tuberculosis (TB) disease and linked with negative regulation of IL-1 signaling. However, mice lacking IL-1 signaling can control Mtb infection if infected with an Mtb strain carrying the rifampin-resistance conferring mutation H445Y in its RNA polymerase β subunit (rpoB-H445Y Mtb). The mechanisms that govern protection in the absence of IL-1 signaling during rpoB-H445Y Mtb infection are unknown. In this study, we show that in the absence of IL-1 signaling, type I IFN signaling controls rpoB-H445Y Mtb replication, lung pathology, and excessive myeloid cell infiltration. Additionally, type I IFN is produced predominantly by monocytes and recruited macrophages and acts on LysM-expressing cells to drive protection through nitric oxide (NO) production to restrict intracellular rpoB-H445Y Mtb. These findings reveal an unexpected protective role for type I IFN signaling in compensating for deficiencies in IL-1 pathways during rpoB-H445Y Mtb infection.

Author summary

Mycobacterium tuberculosis (Mtb) is the causative agent of tuberculosis (TB) and is responsible for 10 million active cases worldwide annually. The acquisition and dissemination of drug resistance to front-line antibiotics have stifled efforts to control and combat TB. Understanding the immune responses elicited during Mtb infection is imperative to develop novel therapeutics that bolster protective responses and limit detrimental processes. Recently, we found that mice do not require the inflammatory interleukin-1 (IL-1) signaling pathway to control Mtb infection if the infecting bacterial strains carry a specific drug resistance-conferring mutation. Instead, here we show that a different signaling pathway, type I interferon (IFN), mediates protection against drug-resistant mutant Mtb infection in the absence of IL-1 signaling. We also found that the primary producers of type I IFN during drug-resistant Mtb infection were recruited innate cells, namely monocytes and macrophages. The downstream impacts of IFN signals were varied, with macrophages restricting bacterial replication while other immune cell functions were likely impaired by IFN signaling. Our study provides new insights into the different host requirements for protection against infection by various Mtb strains and highlights the unexpected roles for type I IFN signaling during Mtb infection.

Introduction

Mycobacterium tuberculosis (Mtb) is estimated to cause 10 million active cases of tuberculosis (TB) annually [1]. Global dissemination of drug resistance to front and second-line antibiotics limits efforts to contain and combat this pervasive disease [27]. Alternatively, to avoid the toxicity associated with anti-TB chemotherapies, researchers are developing host-directed therapeutics to better augment the host immune response to Mtb infection and effective vaccines that protect against TB disease. These strategies require a thorough understanding of the host-pathogen interactions and the immune events correlated with protective responses to Mtb infection.

Substantial efforts have been made to understand important immune cell populations and signaling pathways necessary for host protection and survival during Mtb infection [814]. Interleukin-1 (IL-1) signaling is essential for protection during Mtb infection, as hosts with defective IL-1 signaling experience immunopathology and succumb to infection [15,16]. Conversely, type I interferon (IFN) signaling is thought to be detrimental during Mtb infection through antagonism of IL-1 signaling and induction of maladaptive inflammatory processes [1719]. In humans, type I IFN signaling is associated with TB disease progression from controlled Mtb infection [2023]. We previously found that the presence of a drug resistance-conferring mutation, rpoB-H445Y, can alter host-pathogen interactions such that IL-1 signaling is not necessary to control drug-resistant (rpoB-H445Y) Mtb infection [7,24]. Additionally, cytokine production is altered during infection of macrophages with rpoB-H445Y Mtb in vitro and in vivo [25], with heightened type I IFN and limited IL-1 production. The rpoB-H445Y Mtb strain contained an increased abundance of the long-chained lipid phthiocerol dimycocerosate (PDIM) in its cell wall. Treatment of wildtype (wt) Mtb-infected BMDMs with purified PDIM from rpoB-H445Y Mtb was sufficient to drive altered cytokine production [24]. However, we have not determined the IL-1-independent protective pathways that led to the control of this rpoB-H445Y Mtb infection.

In the current study, we demonstrate that, in contrast to what is seen with wild type (wt) Mtb infection, type I IFN signaling has a protective role during rpoB-H445Y Mtb infection. While a global loss of either IL-1 or type I IFN signaling alone did not affect rpoB-H445Y Mtb replication in vivo, mice with deficient IL-1 signaling were susceptible to rpoB-H445Y Mtb infection when they also lacked type I IFN signaling. Increased rpoB-H445Y Mtb burden was associated with an increase in inflammatory cell influx and inflammation. Furthermore, recruited myeloid cells, such as monocytes and macrophages, were the source of type I IFN during rpoB-H445Y Mtb infection. Interestingly, the role of type I IFN signaling in host protection depended upon cell type, such that it was protective in LysM+ cells but detrimental in CD11c+ cells. In vitro experiments suggested that both type I IFN and IL-1 signaling led to nitric oxide (NO) production, which limited rpoB-H445Y Mtb growth. Our results highlight the complex role of type I IFN signaling and that host protective mechanisms can be markedly different during infection with two closely related bacterial strains. Insights from this work could be used to suggest potential targets for future avenues of study in the development of host-directed therapies against drug-resistant Mtb.

Results

Type I IFN signaling mediates protection in the absence of IL-1 signaling in rpoB-H445Y Mtb infection

IL-1 signaling is essential in controlling wt Mtb infection [15,18]. However, this signaling pathway is dispensable when mice are infected with rpoB-H445Y Mtb [24], a drug-resistant but non-isogenic Mtb strain derived from wt Mtb [25], suggesting that there are alternate signaling pathways mediating protection. To determine if type I IFN signaling was protective in the absence of IL-1 signaling against rpoB-H445Y Mtb in vivo, we generated mice lacking both IL-1 and type I IFN receptors and evaluated their susceptibility to wt or rpoB-H445Y Mtb infection. As expected, Il1r1-/- mice were unable to control wt Mtb infection, with elevated lung burden at 14- and 30-days post-infection (dpi), relative to C57BL/6 (B6) mice (Fig 1A), whereas loss of Ifnar1 alone has no impact on Mtb burden. As before, following wt Mtb infection [18,20], the increase in lung burden caused by loss of Il1r1 was reversed in mice lacking both Ifnar1 and Il1r1 [double knockout (DKO)] (Fig 1A). An increase in the susceptibility of Il1r1-/- mice was accompanied by an increase in inflammatory damage in the lungs of wt Mtb infected (S1A Fig), but this inflammatory damage did not significantly decrease in the DKO mice where Mtb control was evident. We also observed increased numbers of recruited macrophages (RMs) expressing high levels of Major Histocompatibility Complex Class II (MHCIIhi), a surrogate marker for activation, of neutrophils, and of MHCIIhi monocytes in the lungs of wt Mtb-infected Il1r1-/- mice (S1B Fig). In the infected DKO mice where susceptibility was reversed, only the numbers of neutrophils were significantly decreased. We found no differences in the numbers of cytokine-producing, activated (CD44hi) T lymphocyte populations between the infected groups (S1C and S1D Fig). A different pattern of susceptibility was apparent when we infected the mice with rpoB-H445Y Mtb. We observed higher bacterial burdens only in the DKO mice at 30 dpi (Fig 1B), while loss of either IL1 or type I IFN signaling alone did not impact lung rpoB-H445Y Mtb burden. The increased bacterial load in the DKO rpoB-H445Y Mtb-infected lungs also coincided with increased lung inflammation (Fig 1C) and increased numbers of lung MHCIIhi RMs and MHCIIhi myeloid dendritic cells (mDCs) in the DKO mice at 30 dpi (Fig 1D). However, there were no differences in the numbers of neutrophils, MHCIIhi monocytes, MHCIIhi alveolar macrophages (AMs), and cytokine-producing, CD44hi T lymphocyte populations between the DKO mice and B6 and Il1r1-/- mice (S2A–S2E Fig). As previously noted [25], loss of Ifnar1 alone was sufficient to drive an increase in the numbers of these cellular populations and the amount of inflammation during rpoB-H445Y Mtb infection. Thus, while type I IFN signaling is dispensable globally for control of wt and rpoB-H445Y Mtb infection, in the absence of IL-1 signaling, type I IFN signaling is necessary for immunity against rpoB-H445Y Mtb.

thumbnail
Fig 1. Type I interferon signaling is protective in the absence of interleukin-1 signaling in rpoB-H445Y Mycobacterium tuberculosis infection.

C57Bl/6 (B6), Ifnar1-/-, Il1r1-/-, and Ifnar1-/- Il1r1-/- [double knockout (DKO)] mice were aerosol infected with a low dose of wt or rpoB-H445Y Mtb. Lung Mtb burden was evaluated at 14- and 30-days following A) wt or B) rpoB-H445Y Mtb infection. C) Formalin-fixed, paraffin-embedded (FFPE) lung sections were stained with hematoxylin and eosin (H&E), and the total inflammatory areas were quantified after rpoB-H445Y Mtb infection at 30 days post-infection (dpi). D) Using flow cytometry, the total numbers of lung MHCIIhi RMs and mDCs were determined at 30 dpi after infection with rpoB-H445Y Mtb. The data shown represent the means ± SD of seven to ten mice per experiment. The data were evaluated for normality using the Shapiro-Wilk Test and passed (p-value > 0.05). Two-way ANOVA was used for A and B, and One-way ANOVA was used for C and D, with Tukey’s multiple comparisons test. Significant differences are indicated with asterisks (*, p-value ≤ 0.05; **, p-value ≤ 0.01; ***, p-value ≤ 0.001; ****, p-value ≤ 0.0001) by appropriate statistical tests. One of three independent experiments shown.

https://doi.org/10.1371/journal.ppat.1012137.g001

Similar type I IFN transcriptional pathways are induced during wt and rpoB-H445Y Mtb infections but result in divergent infection outcomes

Type I IFN signaling has been thought to be detrimental during wt Mtb infection based upon its association with the progression of TB patients towards active disease and its ability to antagonize the protective IL-1 signaling pathway [18,20,22]. To discern the roles of type I IFN signaling during both Mtb infections, we determined the lung transcriptional landscape in response to either infection at 14 days post-infection, an early timepoint before significant differences in lung rpoB-H445Y Mtb burden were apparent in the gene-deficient mice. We identified the most differentially expressed genes (DEGs) across groups and categorized them by pathways (Fig 2). In the lungs of mice infected with wt Mtb, genes related to IFN signaling were upregulated in the B6 and Il1r1-/- mice compared with Ifnar1-/- and DKO mice (Fig 2A). Specific analysis of the type I IFN transcriptional signature revealed downregulation of genes in this pathway in the Ifnar1-/- and DKO mice when compared to the B6 and Il1r1-/- mice during wt Mtb infection (S3A Fig). Additionally, genes involved in antigen presentation via MHC I, such as H2-q4 and H2-t22, were induced by type I IFN signaling during wt Mtb infection at this early timepoint. We also observed induction of genes involved in branched chain amino acid catabolism (Aldh6a1 and Ivd), phagocytosis [26] (Snap23 and Stx4a), endoplasmic reticulum (ER) stress [27,28] (Hspa5 and Calr), and cellular response to heat stress (Hsp90aa1 and St13) specifically in the B6 mice infected with wt Mtb relative to the single knockout and DKO mice (Fig 2A). We observed enrichment of DEGs in pathways involved in type I IFN signaling during wt Mtb infection in both B6- and Il1r1-/—infected mouse groups, although they differed in their ability to control infection. Together, our data suggest that decreased co-expression of type I IFN-induced genes with genes involved in metabolism and stress could drive detrimental does not determine the outcomes of wt Mtb infection.

thumbnail
Fig 2. Type I IFN signaling signatures are common across wt and rpoB-H445Y Mtb infections.

C57Bl/6 (B6), Ifnar1-/-, Il1r1-/-, and Ifnar-/- Il1r1-/- [double knockout (DKO)] mice were aerosol infected with a low dose of wt or rpoB-H445Y Mtb and sacrificed at 14 dpi. RNA was extracted from homogenized lung tissue and sequenced. Heatmaps displaying average expression levels of top differentially expressed genes (DEGs) across mouse genotypes after A) wt or B) rpoB-H445Y Mtb infection with annotated pathways (four to six mice per group). Significant DEGs related to the corresponding reactome pathways are visualized using heatmaps(|log2FC| >0, adjusted p-value < 0.05). Average expression (normalized) of selected genes on log2 scale are shown.

https://doi.org/10.1371/journal.ppat.1012137.g002

In the lungs of mice infected with rpoB-H445Y Mtb, we also observed that DEGs related to IFN signaling were upregulated in B6 and Il1r1-/- mice, whereas these pathways were downregulated in the Ifnar1-/- and DKO mice (Fig 2B). Indeed, the type I IFN transcriptional signature was downregulated in both the Ifnar1-/- and DKO mouse groups when compared with either wt or Il1r1-/- mice following rpoB-H445Y Mtb infection (S3B Fig). Interestingly, expression of some interferon-stimulated genes (ISGs), such as Abce1 and Ptpn11, was higher in the lungs of Il1r1-/- mice than in B6 mice after rpoB-H445Y Mtb infection. Another IFN-related gene, Ube2l6, decreased in expression in the rpoB-H445Y Mtb infected Il1r1-/- mice lungs relative to infected B6 counterparts (Fig 2B). These findings suggest that IL-1 signaling plays a complex role in the regulation of IFN-related genes. Coincident with the upregulation of IFN-related genes in B6 and Il1r1-/- rpoB-H445Y Mtb-infected mice, genes involved in metabolic pathways, such as fatty acid metabolism and cholesterol biosynthesis were upregulated. In contrast, mice lacking IL-1 signaling exhibited decreased expression of genes, such as Slc2a3 and Slc4a1, that are part of the solute carrier (SLC)-mediated transmembrane transport family. Therefore, our data indicate that the loss of rpoB-H445Y Mtb control observed in DKO mice was associated with the downregulation of various cellular processes that could be triggered by either type I IFN or IL-1 signaling. Many IFN-related genes, such as Isg15, Ifnar1, and Mx2, were induced following either wt or rpoB-H445Y Mtb infection. In contrast, genes such as Abce1 and Ptpn11 were only upregulated following rpoB-H445Y Mtb infection, whereas Uba7 was upregulated only during wt Mtb infection. Furthermore, Ube2l6, which exhibited differential regulation in Il1r1-/- and B6 mice during rpoB-H445Y Mtb infection, was upregulated in both groups of mice during wt Mtb infection. So, while we observed broad induction of many IFN signaling-related genes during both Mtb infections, some IFN-related genes were differentially induced across wt and rpoB-H445Y Mtb infections.

CD11b+ myeloid cells produce type I IFN, while LysM+ cells respond to type I IFN following rpoB-H445Y Mtb infection

Type I IFN is rapidly produced by many cell types in response to infection [21,29]. We used IFNβ-YFP reporter mouse strain, in which cells that produce IFN β also co-express YFP and are readily detectable [3032] (Fig 3A), and found that type I IFN production was induced in various myeloid lung populations following infection with rpoB-H445Y Mtb (Fig 3B). As early as 7 dpi, rpoB-H445Y Mtb infection lead to significantly higher accumulation of IFN-producing lung monocytes and RMs as compared with uninfected controls. No discernable production of type I IFN in AMs, neutrophils, or any non-immune (CD45-) population was detected at this timepoint. Plasmacytoid dendritic cells (pDCs), a major producer of type I IFN during bacterial or viral infection [14,32], were not readily detected in our analysis. We then tracked type I IFN production in rpoB-H445Y Mtb infection at later timepoints (Fig 3B). Lung monocytes were also the primary cell type expressing type I IFN production at 14 dpi, and RMs were the major cellular population that was producing type I IFNs at 28 dpi. Taken together, our data indicate that the main sources of type I IFN production in rpoB-H445Y Mtb infection are CD11b+ myeloid cells, namely monocytes and RMs.

thumbnail
Fig 3. Type I IFN producers are CD11b+ myeloid cells, while LysM+ cells respond to type I IFN following rpoB-H445Y Mtb infection.

IFNβ-YFP mice were aerosol infected with rpoB-H445Y Mtb and sacrificed at the indicated timepoints along with uninfected controls. A) Representative histograms for various subsets [Monocytes (left), Recruited Macrophages (RMs) (middle left), Neutrophils (middle right), and Alveolar Macrophages (AMs) (right)] depicting YFP expression by intracellular antibody staining (green) compared to controls [uninfected IFNβ-YFP mice stained with intracellular antibody (orange), infected IFNβ-YFP mice stained with intracellular isotype antibody (blue), and infected B6 non-reporter mice stained with intracellular antibody (red)]. B) The frequencies of YFP+ cells were determined by flow cytometry following sacrifice of rpoB-H445Y Mtb-infected IFNβ-YFP mice at 7, 14, and 28 dpi, along with uninfected controls. C) Experimental scheme for the LysM-Cre Ifnar1fl/fl mice with the purple arrow indicating aerosol infection with rpoB-H445Y Mtb and the pink arrow indicating administration of a-IL1R or isotype antibodies i.p. D) Lung Mtb burden was determined at 28 dpi. E) LysM-Cre+ Ifnar1fl/fl (semi-closed green circles) and Cd11c-Cre+ Ifnar1fl/fl mice (semi-closed orange circles), along with littermate controls (open blue circles), were infected with rpoB-H445Y Mtb, and bacterial burden was determined in the lungs at 28 dpi. The data shown represent the means ± SD of three to eight mice per experiment. The data were evaluated for normality using the Shapiro-Wilk Test and passed (p-value > 0.05). Two-way ANOVA and One-way ANOVA, with Tukey’s multiple comparisons tests, were used for B and D-E, respectively. Significant differences are indicated with asterisks (*, p-value ≤ 0.05; **, p-value ≤ 0.01; ***, p-value ≤ 0.001; ****, p-value ≤ 0.0001) by appropriate statistical tests. One-way ANOVA. Two independent experiments shown.

https://doi.org/10.1371/journal.ppat.1012137.g003

Loss of control of rpoB-H445Y Mtb infection in mice lacking type I IFN and IL-1 signaling was accompanied by an influx of inflammatory myeloid cells, and the primary producers of type I IFN throughout rpoB-H445Y Mtb infection were myeloid cells. To test whether myeloid cells were also responding to type I IFN in rpoB-H445Y Mtb infection in the absence of IL-1 signaling, we generated LysM-Cre Ifnar1fl/fl mice (deletion of the type I IFN receptor on LysM-expressing myeloid cells [33]), infected these mice with rpoB-H445Y Mtb, and also treated them with IL-1 blocking antibodies or isotype controls (Fig 3C). As we observed in the rpoB-H445Y Mtb-infected Il1r1-/- mice, blocking of IL-1 signaling with antibody did not impact lung bacterial burden in the Ifnar1fl/fl mice (Fig 3D). But, similar to the rpoB-H445Y Mtb-infected DKO mice, loss of the type I IFN receptor in LysM-Cre+ Ifnar1fl/fl mice treated with IL-1 signaling blocking antibody led to an increase in lung Mtb burden. Unexpectedly, loss of the type I IFN receptor in the LysM-expressing cells alone led to increased rpoB-H445Y Mtb lung burden (Fig 3D). The increased susceptibility in LysM-Cre+ mice infected with rpoB-H445Y Mtb correlated with increased pulmonary inflammation (S4A Fig) and increased accumulation of lung MHCIIhi RMs and mDCs (S4B Fig). Together, our results suggest that type I IFN signaling is necessary in LysM+ cells to mediate control of rpoB-H445Y Mtb infection but dispensable when deleted globally as in the Ifnar1-/- mice. To better understand this apparent discrepancy, we assessed the impact of the loss of type I IFN signaling in CD11c-expressing cells, which include AMs and DCs (Fig 3E). Cd11c-Cre+ Ifnar1fl/fl mice better controlled rpoB-H445Y Mtb infection relative to littermate controls. Thus, during rpoB-H445Y Mtb infection, type I IFN signaling is protective in LysM-expressing cells, especially in the absence of IL-1 signaling, but detrimental in CD11c-expressing cells.

NO production restricts intracellular rpoB-H445Y Mtb growth in murine macrophages

To address the importance of macrophages in producing and responding to type I IFN signals, we utilized bone marrow-derived macrophages (BMDMs) to elucidate the mechanistic role of type I IFN signaling in the absence of IL-1 signaling during rpoB-H445Y Mtb infection. As before [24], BMDMs lacking IL-1 signaling were more susceptible to wt Mtb infection at 6 dpi (Fig 4A). Additional loss of type I IFN signaling rescued this defect in the DKO BMDMs. In contrast, during rpoB-H445Y Mtb infection, loss of either type I IFN or IL-1 signaling led to an increase in intracellular Mtb growth, with loss of both (DKO) causing a further increase in Mtb growth (Fig 4B). We also assessed cell death dynamics during the BMDM infections (S5 Fig). Across the four genotypes, we find that BMDMs lacking only Il1r1 exhibited increased cell death as seen with increased frequencies of cells at apoptotic (AnnexinVhi 7AADlo) and necrotic (AnnexinVhi 7AADhi) stages. While wt Mtb drove high levels of necrotic cell death early in infection, the rpoB Mtb mutant induced high levels of necrosis later during infection. Similarly, high levels of apoptosis were observed in Il1r1-/- BMDMs only later during rpoB mutant Mtb infection. As both infections induced high levels of cell death in the Il1r1-/- BMDMs, but not in the DKO BMDMs, these responses could be a general response to Mtb infection that is mediated by type I IFNs [34]. While these increased frequency of cell death in Il1r1-/- BMDMs coincided with their inability to control wt Mtb infection, the same does not appear to be the case during rpoB Mtb mutant infection, suggesting a role for other host cellular processes to mediate protection against rpoB-H445Y Mtb.

thumbnail
Fig 4. NO production restricts intracellular rpoB-H445Y Mtb growth in macrophages.

Bone marrow-derived macrophages (BMDMs) from B6, Ifnar1-/-, Il1r1-/-, DKO mice were infected with either A) wt Mtb or B) rpoB-H445Y Mtb, and Mtb CFU was determined at 6 dpi. C) The concentration of nitrite in the supernatants of wt or rpoB-H445Y Mtb infected BMDMs was determined, along with uninfected controls. D) B6, Il1r1-/-, Ifnar-/- and DKO BMDMs were infected with rpoB-H445Y Mtb and then treated with the indicated concentrations of an NO scavenger, Carboxy-PTIO (C-PTIO). Bacterial burden was determined on 6 dpi. Fold change in Mtb growth was determined relative to untreated and infected BMDMs. E) B6 and Nos2-/- BMDMs were infected with rpoB-H445Y Mtb, and Mtb CFU was determined at 6 dpi. The data shown represent the means ± SD of three to four biological replicates per experiment. The data were evaluated for normality using the Shapiro-Wilk Test and passed (p-value > 0.05). One-way ANOVA and Two-way ANOVA, with Tukey’s multiple comparisons tests, were used for A-B and C-D, respectively. Unpaired two-tailed student’s t-test was used for E. Significant differences are indicated with asterisks (*, p-value ≤ 0.05; **, p-value ≤ 0.01; ***, p-value ≤ 0.001; ****, p-value ≤ 0.0001) by appropriate statistical tests. One of three independent experiments shown.

https://doi.org/10.1371/journal.ppat.1012137.g004

Type I IFN signaling can drive nitric oxide (NO) production as part of its anti-bacterial functions [29]. Thus, we assessed the abundance of nitrite, the byproduct of NO generation, in the supernatants of wt or rpoB-H445Y Mtb infected BMDMs (Fig 4C). BMDMs from mice with functional type I IFN and IL-1 signaling produced increased NO in response to rpoB-H445Y Mtb infection when compared with other infected BMDMs. While there was a reduction in NO production in rpoB-H445Y Mtb infected BMDMs lacking either IL-1 or type I IFN signaling relative to infected B6 BMDMs, BMDMs from these two groups still produced more NO than their infected DKO counterparts. The stepwise decrease in NO production and the stepwise increase in CFU between B6 BMDMs to Ifnar1-/- and Il1r1-/- BMDMs to DKO BMDMs support the idea that NO production is associated with protection against rpoB-H445Y Mtb infection. To assess the role of NO in controlling rpoB-H445Y Mtb infection, we used carboxy-PTIO (C-PTIO), an NO scavenger, at varying concentrations (Fig 4D). Relative to untreated controls, the addition of 100 uM C-PTIO resulted in a loss of control in B6 BMDMs during rpoB-H445Y Mtb infection. In contrast, the addition of 1 uM and 10 uM of C-PTIO was sufficient to increase the susceptibility of Ifnar1-/- and Il1r1-/- BMDMs, respectively, while adding any amount of C-PTIO did not impact the course of infection in DKO BMDMs. This “minimum susceptibility” analysis suggests a stoichiometric role for NO in limiting rpoB-H445Y Mtb during infection. To more specifically assess the role of NO production in controlling rpoB-H445Y Mtb infection, we generated BMDMs from Nos2-/- bone marrow and compared their ability to control bacterial replication relative to B6 BMDMs (Fig 4E). Similar to B6 BMDMs where NO production was inhibited by C-PTIO addition, BMDMs that were unable to produce NO due to lack of the inducible nitric oxide synthase were unable to control mycobacterial growth. Collectively, our data show that the protective mechanisms underlying type I IFN signaling are partially through NO production to restrict intracellular rpoB-H445Y Mtb replication.

Discussion

IL-1 signaling is essential in the early immune response to Mtb infection [15,16]. However, the rifampin-resistant rpoB-H445Y Mtb bypass the need for IL-1 signaling, and the protective immune pathways in play have not been understood [24]. In this study, we show a protective role for type I IFN signaling in the absence of IL-1 signaling during infection with rpoB-H445Y Mtb. Additionally, the producers of type I IFN during infection are lung monocytes and recruited macrophages. During rpoB-H445Y Mtb infection, while LysM-expressing cells respond to the type I IFN signals to confer protection, type I IFN signaling in CD11c-expressing cells has a detrimental role. Finally, we show that type I IFN and IL-1 signaling cooperate to drive NO production to limit rpoB-H445Y Mtb replication. Together, our results demonstrate a protective role for type I IFN signaling during rpoB-H445Y Mtb infection, in contrast to its proposed detrimental role during wt Mtb infection. These results highlight fundamentally divergent host-pathogen interactions during closely related Mtb infections.

The impacts of type I IFN and IL-1 signaling in wt Mtb infection and on TB pathogenesis have been well studied [20,35,36]. In susceptible mouse models of TB such as C3Heb/FeJ and 129S2, type I IFNs facilitate the recruitment of Mtb-permissive cell types, drive neutrophil extracellular traps that impair Mtb control, and promote exacerbated immunopathology and disease [19,23,34,37]. Indeed, in human TB, type I IFNs and neutrophils are associated with and can predict the progression toward clinical TB diagnosis [22,38,39]. In models utilizing resistant mice such as B6, IL-1 signaling is necessary to control Mtb infection, especially through the production of eicosanoids, such as PGE2[15,16,18]. In these models, type I IFN signaling plays a detrimental role through antagonizing IL-1 production and signaling, limiting effective T cell immunity, and sensitizing infected macrophages towards death [17,18,34,37,40,41]. Our data also shows that lack of IL-1 signaling during wt Mtb infection results in increased frequencies of macrophage death in vitro as well as increased accumulation of neutrophils in the lungs in vivo and how these processes could be driven by type I IFN signaling. Heightening type I IFN induction through poly-IC treatment in B6 mice also led to increased Mtb lung burdens through recruitment of Mtb-permissive immune cells [42]. We have recently shown that type I IFN production is heightened during rpoB-H445Y Mtb infection, with type I IFN signaling leading to a suppressed immune response in the lungs of rpoB-H445Y Mtb-infected mice [25]. Correspondingly, lack of type I IFN signaling during rpoB-H445Y Mtb infection led to increased immune cell infiltration and activation and more inflammation, which is shown here as well. But, this more robust immune response in Infar-/- mice led to better control of rpoB-H445Y Mtb and limited inflammation during the chronic stages of infection [25]. However, some studies have found a protective role for type I IFN signaling during wt Mtb infection [35,43]. For example, autocrine type I IFN signaling can promote macrophage control of wt Mtb infection, but Mtb has been shown to modulate host macrophage responses to limit potential protective autocrine type I IFN signaling, while promoting detrimental paracrine signaling [43]. Also, in a context without protective type II IFN signaling, additional loss of type I IFN signaling led to the double IFN knockout mice succumbing to Mtb infection earlier than either single IFN knockout mice [35]. Together, these studies suggest a complex role for type I IFN signaling where wt Mtb could be hijacking this host signal during infection to subvert protective immune responses such as IL-1 signaling.

In our work with the DR Mtb strains containing the rpoB-H445Y mutation, we have found that the impact of type I IFN and IL-1 signaling on Mtb infection are drastically changed. Previously, we found antagonism of IL-1 production by type I IFN signaling and heightened type I IFN production during infection with rpoB-H445Y Mtb. Yet, rpoB-H445Y Mtb-infected B6 mice were able to control infection, even without requiring IL-1 signaling [24]. In the current study, we find that type I IFN signaling pathway genes are induced in the lungs following wt or rpoB-H445Y Mtb infection. However, while decreased expression of genes induced by type I IFN was seen in infected DKO mice relative to their Il1r1-/- counterparts, these transcriptional changes were associated with protective or detrimental host responses during wt or rpoB-H445Y Mtb infection, respectively. So, the difference in host control of these two infections could be driven by differences in which specific ISGs were induced as well as how those ISGs function in tandem with other upregulated cellular processes, such as different metabolic pathways and stress-related responses [24,44]. Thus, contextual cues could lead to diverse outcomes of type I IFN signaling during Mtb infections. Further work is needed to study how the wt and rpoB-H445Y Mtb strains drive differential expression of host genes and processes and how those pathways contribute to protective or detrimental effects of type I IFN signaling during infection.

Type I IFNs can be produced by many cell types, especially in response to microbial infection [36]. Following rpoB-H445Y Mtb infection, monocytes and RMs were the major producers of type I IFNs. Kotov et al. also recently investigated sources of type I IFNs during wt Mtb infection and found interstitial macrophages and pDCs to be dominant producers [45]. While we also found recruited monocyte and macrophage populations to be contributing to type I IFN production, we did not observe a significant population of pDCs in the lungs during infection, a cell type known to be a potent producer of type I IFNs [14,46], nor did we find IFN-producing non-immune epithelial cells [47]. These findings could be attributed to pDCs and epithelial cells acting earlier than 7 dpi or our approach not being sensitive enough to capture the pDC and epithelial cell populations. For example, Kotov et al. used a double fluorescent reporter sytem to identify all cells that have expressed Ifnb1 at any point during infection, which would capture early or transient expression of type I IFNs. Thus, we cannot rule out a role for pDCs in our rpoB-H445Y Mtb infection model, which could be studied in follow-up work. Also, Kotov et al. carried out their studies using the Mtb Erdman strain, while we used the Mtb HN878 strain as our reference wt strain, and differences between these strains could impact the dynamics of type I IFN production by different cell types. Future experiments could also test whether myeloid cells directly infected with Mtb are type I IFN producers and how the kinetics of infection influences the kinetics of IFN production found in this study [48,49]. Additional insights can also be gained by utilizing newly emerging technologies such as spatial transcriptomics, to study localization of the IFN-producing cell types with respect to infected and other responding cell types [5052].

Type I IFNs signal through the IFNAR1/IFNAR2 heterodimeric receptor, which is ubiquitously expressed [36]. Thus, depending on the cell type and other signaling processes, a wide variety of cellular responses can be induced, some that are protective and others that are detrimental. In the rpoB-H445Y Mtb infection model, even with functional IL-1 signaling, LysM-expressing cells required type I IFN signaling to drive a protective immune response, while type I IFN signaling was detrimental in CD11c-expressing cells. It should be noted that LysM is broadly expressed in a variety of immune cell types such as monocytes, macrophages, and granulocytes as well as some non-immune type II alveolar epithelial cells [33,5355]. Our results show cell-specific effects of type I IFN signaling during rpoB-H445Y Mtb infection, but we can only speculate on the specific contribution of each cell type during infection. Neutrophils are a major cell type targeted by the LysM-Cre model and not affected in the CD11c-Cre model and, as noted above, have been associated with type I IFN signaling in other Mtb infections and shown here in wt HN878 Mtb infection. However, we saw no increases in numbers of neutrophils in either of the susceptible DKO mice or the LysM-Cre+ Ifnar1fl/fl mice. While there could be differences in functionality between the neutrophils from protected and susceptible mice, such possibilities require follow-up investigation in the future. The consistent findings of increased abundance of activated RMs across susceptible mouse models, RMs as a source of type I IFN production, and a susceptible BMDM infection model suggest that these infiltrating macrophages are likely an immune subset contributing to a deficient immune response to rpoB-H445Y Mtb in the absence of type I IFN signaling.

On the other hand, CD11c expression in the lungs is restricted to AMs and DCs. As AMs are a common subset targeted by both Cre mouse models, it is likely that loss of type I IFN signaling in this subset does not play a role in rpoB-H445Y Mtb infection. However, our findings suggest a likely detrimental role for type I IFN signaling on DCs, which is supported by our previous study where we observed reduced BMDC migration in vitro and limited trafficking of Mtb to the draining lymph nodes in vivo during rpoB-H445Y Mtb infection [25]. We also previously found that global loss of Ifnar does not appear to impact rpoB-H445Y Mtb bacterial burden by 30 dpi, but Ifnar-/- mice carry reduced bacterial loads during chronic stages of infection (≥ 50 dpi)[25]. Together, our data suggest that, in infected Ifnar-/- mice, the deficiency in the ability of RMs to control rpoB-H445Y Mtb growth might be offset by a more robust adaptive immune response triggered by DCs that are not limited by type I IFN signaling, which also improves long-term control of rpoB-H445Y Mtb infection [56,57]. The susceptibility of the DKO mice to rpoB-H445Y Mtb infection and limited adaptive responses, relative to infected Ifnar-/- mice, suggest that IL-1 signaling might play a role in the DC priming of the adaptive responses and is a mechanism contributing to control. The cell-cytokine specific contributions to protection against rpoB-H445Y Mtb should be explored in future studies.

In this study, we observed a role for NO in limiting rpoB-H445Y Mtb growth, with high levels of NO associated with control of rpoB-H445Y Mtb infection and limiting NO production resulted in increased rpoB-H445Y Mtb growth. This role for NO might be Mtb strain specific and not play a significant role during wt Mtb infection. Indeed, infection with rpoB-H445Y Mtb induced higher levels of NO than with the wt Mtb strain. Although we have not determined the mechanism behind the NO production, type I IFN and IL-1 signaling could function in tandem with another signal, such as TNF, to promote Nos2 expression and NO production [58]. Similarly, on the bacterial side, PDIM or another virulence factor might contribute to suppression of NO production during wt Mtb infection, and this suppression might be lost during rpoB-H445Y Mtb infection due to differences in virulence factors between wt Mtb and and rpoB-H445Y Mtb strains [24]. Future work can aim to dissect the roles of host and pathogen factors in contributing to NO production. Studies in murine models suggest that NO contributes to limiting Mtb replication directly or indirectly, especially in macrophage subsets [59,60]. In primary human AM studies, NO induction has been mechanistically implicated in the killing of Mycobacterium species in vitro [61,62]. Thus, while NO production is elevated in murine cells it may not be reflective of Mtb control in human macrophages [63]. Our results show that both type I IFN and IL-1 signaling cooperate to induce production of NO in an additive manner. While type I IFN signaling has been more closely linked with NO production through the inducible nitric oxide synthase (Nos2) [43,64], it is possible that IL-1 could also drive NO production as in the case of Brucella infection [65]. Collectively, the data suggest that the dispensability of either signaling pathway during in vivo rpoB-H445Y Mtb infection is tied to compensatory NO production that is sufficient for protection.

An important question arising from this work is why wt and rpoB-H445Y Mtb are driving different infection outcomes. As we observe induction of similar pathways, albeit at different magnitudes, we hypothesize that differences in the expression of virulence factors between wt and rpoB-H445Y Mtb could influence the magnitude and timing of host cell responses. Concurrently, differential expression of virulence factors could impact the physiology of rpoB-H445Y Mtb and alter the response of the mutant Mtb to the pathways induced by type I IFN signaling, such as susceptibility to NO. We speculate that the location of the mutation in rpoB, which is a critical component of the bacterial transcription machinery, may result in broad transcriptional changes and expression of virulence factors [24,66]. We hypothesize that the increased production of long-chained PDIM in the rpoB-H445Y Mtb strain could be affecting the integrity of host phagosomes and lead to increased bacterial escape into the cytosol [6770]. But, we have not thoroughly assessed all of the changes these DR mutations drive and how those changes are linked with the divergent host-pathogen interactions. A caveat of this study is that we cannot establish a causal link between the rpoB-H445Y mutation and the immune phenotypes observed in the study. The rpoB-H445Y Mtb strain used in this study was derived from the wt Mtb strain through antibiotic selection, and multiple independent DR Mtb isolates with the rpoB-H445Y mutation drove the same responses in vitro and in vivo [24]. However, a few additional mutations exist that separate the two strains outside of the rpoB-H445Y mutation, and we cannot rule out the possibility that additional mutations contribute to the differential host response [25]. The question of causality remains open and must be established in future work.

In summary, our studies have mechanistically shown how type I IFN signaling can drive protective or detrimental outcomes during Mtb infections and impact TB disease pathogenesis. Additionally, we uncover novel immune circuits through the use of a rpoB-H445Y Mtb strain with two mutually antagonistic signaling pathways acting in a compensatory but not redundant manner. Thus, our work highlights the necessity for greater investigation of host-pathogen interactions using a broader repertoire of Mtb strains and including drug-resistant isolates, which can reveal novel targets for future development of host-directed therapeutics.

Materials and methods

Ethics statement

Protocols involving the use of animals were approved by the Washington University in St. Louis or The University of Chicago Institutional Animal Care and Use Committee (IACUC) guidelines. All of the experiments were performed in accordance with the protocols.

Mice

C57BL/6J (B6), Ifnar1−/−, Il1r1-/-, Nos2-/-, and Ifnbmob mice on the B6 background were purchased from The Jackson Laboratory (Bar Harbor, ME). Ifnar1-/- Il1r1-/- mice were generated by crossing Ifnar1−/− and Il1r1-/- mice. LysM-Cre Ifnar1fl/fl and Cd11c-Cre Ifnar1fl/fl mice were generously provided by the Baldridge Lab (Washington University in St. Louis). In some experiments, LysM-Cre+ and LysM-Cre- Ifnar1fl/fl mice were administered sterile PBS or a-IL1R blocking antibody (BioXCell) at 100 ug/mouse in 200uL every other day intraperitoneally (IP) starting at -1 dpi. All mice were maintained in the animal facility at Washington University in St. Louis and bred in house. Experimental mice were age and sex matched and infected between the ages of 6 and 8 weeks. All mice were maintained and used in accordance with the approved Washington University in St. Louis Institutional Animal Care and Use Committee guidelines.

Generation of BMDMs

Bone marrow derived macrophages (BMDMs) were generated as previously described [11,24]. Briefly, bone marrow cells were collected from the tibia and femur of B6 mice and cultured in complete Dulbecco’s modified eagle’s medium (cDMEM) with 20 ng/mL of recombinant granulocyte-macrophage colony-stimulating factor (GM-CSF). Cells were cultured at 37°C in 5% CO2 and supplemented with media on day 3. On day 7, floating cells were taken up and discarded, and adherent cells were collected as BMDMs.

Mtb infections

Mtb strain HN878 (wt) was obtained from BEI resources (Manassas, VA) under National Institutes of Health contract AI-75320. Independent rifampicin resistant Mtb colonies (biological replicates) were selected from rifampicin (2 μg/ml) containing 7H11 agar plates [3]. The sequences of rpoB were confirmed by Sanger sequencing (Genewiz), and Mtb stocks were created for further experimentation. The whole genomes of all Mtb strains were previously sequenced, with all SNPs between both, wt and rpoB-H445Y, Mtb strains are published [25]. All Mtb strains were cultured in Proskauer Beck medium supplemented with 0.05% Tween 80 and frozen at −80°C while in mid-log phase prior to use in mouse and macrophage infections. Colony forming units (CFU) of the bacterial stocks were determined through serial dilutions on 7H11 agar plates. Mice were aerosol infected with low doses (~100 CFU) of indicated Mtb strains in sterile PBS using a Glas-col nebulizer [71]. Mtb bacterial burden/organ was quantitated by plating serial dilutions of homogenized lung tissue on 7H11 agar plates (BD Biosciences, Franklin Lakes, NJ). Plates were incubated for 2–3 weeks at 37°C, and the number of colonies were counted. In vitro infections were carried out at a multiplicity of infection (MOI) of 1 in antibiotic free media for six days. In some experiments, macrophages were treated with indicated amounts of carboxy-PTIO (C-PTIO) four hours after the start of infection. Bacterial burden in each well was determined through serial plating of supernatant and lysed macrophages after six days post infection.

Flow cytometry

Lungs were perfused with heparin in PBS, minced, digested with DNAse/collagenase, lysed for red blood cells, and pressed through a 70 μm filter to generate a single cell suspension [71]. For quantification of cytokine responses, single cell suspensions were stimulated ex vivo with purified Mtb antigens ESAT6 and Ag85B (10 μg/mL) overnight followed by a five hour stimulation with brefeldin and monensin (BD Biosciences) as previously described [9,10]. Cells were treated with Fc Block (CD16/CD32,2.4G2, Tonbo Biosciences) and stained with appropriate fluorochrome-labeled specific antibodies or isotype control antibodies. Intracellular cytokine staining was performed using the BD Cytofix/Cytoperm kit (BD Biosciences). Mouse antibodies used include CD11b (M1/70; Tonbo Biosciences), CD11c (HL3; BD Biosciences), Gr-1 (RB6–8C5, eBioscience), SiglecF (E50-2440; BD Biosciences), CD64 (X54-5/7.1; Biolegend), GFP polyclonal antibody (Cat#A21311; Thermo Fisher) or rabbit IgG isotype control (Catalog # 11-4614-80), CD3 (500A2; BioLegend), CD4 (RM4–5; BD Biosciences), CD44 (IM7; Tonbo Biosciences), IFN-γ (XMG1.2; Tonbo Biosciences) or isotype control IgG1, k (A85-1; BD Biosciences), CD8 (53–6.7; BD Biosciences), and MHC Class II (M5/114.15.2; Tonbo Biosciences).

Cells were processed with the Becton Dickinson (BD) Fortessa X-20 flow cytometer using FACS Diva software, or the BD FACSJazz flow cytometer using FACS software (BD). Flow cytometry experiments were analyzed using FlowJo (Tree Star Inc.). As before [11,24,72], neutrophils were defined as CD11b+CD11c-Gr-1hi, monocytes were defined as CD11b+CD11c-Gr-1med cells, and recruited macrophages (RMs) were defined as CD11b+CD11c-Gr-1low cells. Myeloid dendritic cells (mDCs) were defined as CD11b+CD11c+ cells and alveolar macrophages (AMs) were defined as CD11b-CD11c+SiglecFhiCD64hi. T cells were defined as CD3+CD4+ or CD3+CD8+ cells. Total numbers of cells within each gate were back-calculated based on cell counts/individual sample.

For the determination of frequencies of apoptotic and necrotic cells, BMDMs were collected at the indicated timepoints and stained with Annexin V (PE) and 7AAD as per the instructions of the PE Annexin V Apoptosis Detection Kit (559763, BD Pharmigen), prior to processing on the X-20 flow cytometer.

Histology

Lung lobes were perfused with 10% neutral buffered formalin and embedded in paraffin. Formalin-fixed paraffin-embedded (FFPE) lung sections were stained with hematoxylin and eosin (H&E) and inflammatory features were evaluated by light microscopy. Inflammatory lesions were outlined with the automated tool of the Zeiss Axioplan 2 microscope (Carl Zeiss) and the total inflammatory area in each lung lobe was measured.

Protein determination

BMDMs were either stimulated with Mtb antigens for 1 day or left unstimulated, lysed in RIPA Lysis Buffer (Santa Cruz Biotechnology), and heat denatured in reducing sample buffer (Thermo Fisher Scientific). Proteins were separated in 4% to 20% polyacrylamide gradient gels (Bio-Rad) and transferred onto PVDF membranes. Membranes were probed with primary antibodies against IFNAR1 (R&D, Cat#AF3039), and GAPDH (Santa Cruz Biotechnology, Cat# SC-25778) followed by incubation with HRP-conjugated anti–rabbit IgG secondary antibody (Santa Cruz Biotechnology) and developed by Clarity Western ECL Substrate (Bio-Rad).

RNA isolation, sequencing, and processing

In brief, lung tissue was homogenized in RLT buffer, and total RNA was extracted with RNeasy Mini Kit (Qiagen) as previously described [11]. Total RNA integrity was determined using Agilent Bioanalyzer or 4200 Tapestation. Library preparation was performed with 5 to 10ug of total RNA with a Bioanalyzer RIN score greater than 8.0. Ribosomal RNA was removed by poly-A selection using Oligo-dT beads (mRNA Direct kit, Life Technologies). mRNA was then fragmented in reverse transcriptase buffer and heating to 94 degrees for 8 minutes. mRNA was reverse transcribed to yield cDNA using SuperScript III RT enzyme (Life Technologies, per manufacturer’s instructions) and random hexamers. A second strand reaction was performed to yield ds-cDNA. cDNA was blunt ended, had an A base added to the 3’ ends, and then had Illumina sequencing adapters ligated to the ends. Ligated fragments were then amplified for 12–15 cycles using primers incorporating unique dual index tags. Fragments were sequenced on an Illumina NovaSeq-6000 using paired end reads extending 150 bases. RNA-seq reads were then aligned and quantitated to the Ensembl release 101 primary assembly with an Illumina DRAGEN Bio-IT on-premise server running version 3.9.3–8 software (GTAC MGI).

The read count data of each gene were generated using featureCounts from Rsubread package [73] with default parameter settings. Significantly differentially expressed genes across different groups were identified using DESeq2, version 1.30.1[74], with default settings and a minimum p-value significance threshold of 0.05 (after FDR correction for the number of tests)[75]. Pathway enrichment analysis among differentially expressed gene sets of interest was performed for Reactome pathways using ReactomePA [76] with p-value ≤ 0.05 after FDR correction. Gene set enrichment analysis was also performed among differentially expressed gene sets of interest with hallmark gene sets for Mus musculus using fgsea [77] with eps set to zero to estimate p-value more accurately and p-value ≤ 0.05 after FDR correction. Additionally, several R packages were used including stringr [78], pheatmap [79], ggplot2[80], and org.Hs.eg.db [81] for data processing and generate figures.

Statistical analysis

All datasets were evaluated for normality using the Shapiro-Wilk Test. Differences between the means of more than two groups were analyzed using One-way ANOVA or Two-way ANOVA, where applicable, with Tukey’s post-tests for normally distributed distributions or Kruskal Wallis Test with Dunn’s multiple comparisons tests. Differences between the means of two groups were analyzed by the unpaired two-tailed student’s t-test. All statistical analyses were done in GraphPad Prism 9. A p-value < 0.05 was considered significant. The data points across the figures represent the mean (±SD or +SEM) of values as noted. *, p≤0.05; **, p≤0.01; ***, p≤0.001; ****, p≤0.0001; ns-not significant (p>0.05). All experiments were replicated for reproducibility.

Supporting information

S1 Fig. Loss of control of wt Mtb infection is associated with increased inflammatory damage and increased abundance of some myeloid subsets.

C57Bl/6 (B6), Ifnar1-/-, Il1r1-/-, and Ifnar1-/- Il1r1-/- [double knockout (DKO)] mice were aerosol infected with a low dose of wt Mtb and sacrificed after 30 dpi. A) Formalin-fixed, paraffin embedded (FFPE) lung sections were stained with hematoxylin and eosin (H&E), and the total inflammatory areas were quantified after wt Mtb infection at 30 days post infection (dpi). B) Total numbers of Major Histocompatibility Complex Class IIhi (MHCIIhi) recruited macrophages (RMs), myeloid dendritic cells (mDCs) MHCIIhi, neutrophils, MHCIIhi alveolar macrophages (AMs), and MHCIIhi monocytes were determined via flow cytometry. The numbers of C) CD4+ CD44hi IFNg+ T cells and D) CD8+ CD44hi IFNg+ T cells were also determined by flow cytometry following ex vivo stimulation with purified Mtb antigens ESAT6 and Ag85B. The data shown represent the means ± SD of four to seven mice per experiment. The data were evaluated for normality using the Shapiro-Wilk Test and passed (p-value > 0.05). One-way ANOVA, with Tukey’s multiple comparisons tests, were used for A-D. Significant differences are indicated with asterisks (*, p-value ≤ 0.05; ns, p-value > 0.05) by appropriate statistical tests. One of three independent experiments shown.

https://doi.org/10.1371/journal.ppat.1012137.s001

(TIF)

S2 Fig. Loss of control of rpoB-H445Y Mtb infection is not associated with changes in other myeloid or T cell subsets.

C57Bl/6 (B6), Ifnar1-/-, Il1r1-/-, and Ifnar1-/- Il1r1-/- [double knockout (DKO)] mice were aerosol infected with a low dose of rpoB-H445Y Mtb and sacrificed after 30 dpi. Total numbers of A) neutrophils, B) Major Histocompatibility Complex Class IIhi (MHCIIhi) monocytes, and C) MHCIIhi alveolar macrophages (AMs) were determined via flow cytometry. The numbers of D) CD4+ CD44hi IFNg+ T cells and E) CD8+ CD44hi IFNg+ T cells were also determined by flow cytometry following ex vivo stimulation with purified Mtb antigens ESAT6 and Ag85B. The data shown represent the means ± SD of five to nine mice per experiment. The data were evaluated for normality using the Shapiro-Wilk Test and passed (p-value > 0.05). One-way ANOVA, with Tukey’s multiple comparisons tests, were used for A-E. Significant differences are indicated with asterisks (*, p-value ≤ 0.05; **, p-value ≤ 0.01; ***, p-value ≤ 0.001; ****, p-value ≤ 0.0001) by appropriate statistical tests. One of three independent experiments shown.

https://doi.org/10.1371/journal.ppat.1012137.s002

(TIF)

S3 Fig. Type I IFN signatures are upregulated in B6 and IL-1 knockout mice infected with either wt or rpoB-H445Y Mtb.

C57Bl/6 (B6), Ifnar1-/-, Il1r1-/-, and Ifnar1-/- Il1r1-/- (DKO) mice were aerosol infected with a low dose of wt or rpoB-H445Y Mtb and sacrificed at 14 dpi. RNA was extracted from homogenized lung tissue and sequenced. The expression of genes in the type I IFN pathway in bulk lung cells from DKO mice over bulk lung cells from the wt mice is shown (top), DKO mice over Il1r1-/- mice (middle), and Ifnar1-/- over wt mice (bottom) after A) wt or B) rpoB-H445Y Mtb infection with annotated pathways (n = 4–6). Gene set enrichment analysis was done using an fgsea R package as noted in Methods. The pathways are enriched in the second group of each comparison.

https://doi.org/10.1371/journal.ppat.1012137.s003

(TIF)

S4 Fig. Increased susceptibility of LysM-Cre+ Ifnar1fl/fl mice to rpoB-H445Y Mtb infection is associated with exacerbated immunopathology.

As described in Fig 3C, LysM-Cre+ Ifnar1fl/fl mice and littermate controls were aerosol infected with a low dose of rpoB-H445Y Mtb and administered a-IL1R or isotype antibodies i.p. every other day, starting one day before infection, and sacrificed at 28 dpi. A) FFPE lung sections that were H&E stained were prepared and the inflammatory areas were quantified at 28 dpi. Representative histological images are shown, with quantification of total inflammatory area divided by total lung area for each lobe depicted below. B) Through flow cytometry, total numbers of MHCIIhi RMs and mDCs were calculated in the lungs at 28 dpi with the rpoB-H445Y Mtb strain. The data shown represent the means ± SD of five to nine mice per experiment. The data were evaluated for normality using the Shapiro-Wilk Test and passed (p-value > 0.05). One-way ANOVA, with Tukey’s multiple comparisons tests, were used for A and B. Significant differences are indicated with asterisks (*, p-value ≤ 0.05; **, p-value ≤ 0.01; ***, p-value ≤ 0.001; ****, p-value ≤ 0.0001) by appropriate statistical tests. One of two independent experiments shown.

https://doi.org/10.1371/journal.ppat.1012137.s004

(TIF)

S5 Fig. Cell death dynamics of rpoB-H445Y Mtb-infected BMDMs are influenced by genotype but are not directly correlated with susceptibility.

BMDMs from B6, Ifnar1-/-, Il1r1-/-, DKO mice were infected with either wt Mtb (circles) or rpoB-H445Y Mtb (squares). At indicated timepoints post infection, BMDMs were collected, stained with Annexin V and 7AAD, per manufacturer’s instruction, and processed using flow cytometry. Uninfected controls were also included. The frequencies of A) apoptotic (7AADlo AnnexinVhi) and B) necrotic (7AADhi AnnexinVhi) BMDMs were quantified. The data shown represent the means ± SD of three to four biological replicates per experiment. The data were evaluated for normality using the Shapiro-Wilk Test and passed (p-value > 0.05). Two-way ANOVA, with Tukey’s multiple comparisons tests, was used for A and B. Significant differences are indicated with asterisks (*, p-value ≤ 0.05; **, p-value ≤ 0.01; ***, p-value ≤ 0.001; ****, p-value ≤ 0.0001) by appropriate statistical tests. One of two independent experiments shown.

https://doi.org/10.1371/journal.ppat.1012137.s005

(TIF)

S1 Data. This file contains the source data used to generate every figure in this manuscript.

Each tab of the Excel file corresponds to a specific figure panel (main and supplemental) and is labeled as such.

https://doi.org/10.1371/journal.ppat.1012137.s006

(XLSX)

Acknowledgments

We thank Dr. Megan Baldridge and her lab for providing the LysM-Cre Ifnar1fl/fl and CD11c-Cre Ifnar1fl/fl mice. We thank Lan Lu for her help in generating the Ifnar1-/- Il1r1-/- mice and maintaining all mice colonies. We thank Dr. Julia Tao for her help in the preparation and analysis of the lung sections by histology. We thank the Genome Technology Access Center at the McDonnell Genome Institute at Washington University School of Medicine for help with genomic analysis. The Center is partially supported by NCI Cancer Center Support Grant #P30 CA91842 to the Siteman Cancer Center from the National Center for Research Resources (NCRR), a component of the National Institutes of Health (NIH), and NIH Roadmap for Medical Research. This publication is solely the responsibility of the authors and does not necessarily represent the official view of NCRR or NIH.

References

  1. 1. Global tuberculosis report 2022. Geneva: World Health Organization; Available: https://www.who.int/publications-detail-redirect/9789240061729
  2. 2. Dookie N, Rambaran S, Padayatchi N, Mahomed S, Naidoo K. Evolution of drug resistance in Mycobacterium tuberculosis: a review on the molecular determinants of resistance and implications for personalized care. J Antimicrob Chemother. 2018;73: 1138–1151. pmid:29360989
  3. 3. Ford CB, Shah RR, Maeda MK, Gagneux S, Murray MB, Cohen T, et al. Mycobacterium tuberculosis mutation rate estimates from different lineages predict substantial differences in the emergence of drug resistant tuberculosis. Nat Genet. 2013;45: 784–790. pmid:23749189
  4. 4. Thirumurugan R, Kathirvel M, Vallayyachari K, Surendar K, Samrot AV, Muthaiah M. Molecular analysis of rpoB gene mutations in rifampicin resistant Mycobacterium tuberculosis isolates by multiple allele specific polymerase chain reaction in Puducherry, South India. J Infect Public Health. 2015;8: 619–625. pmid:26117709
  5. 5. Chikaonda T, Ketseoglou I, Nguluwe N, Krysiak R, Thengolose I, Nyakwawa F, et al. Molecular characterisation of rifampicin-resistant Mycobacterium tuberculosis strains from Malawi. Afr J Lab Med. 2017;6: 463. pmid:28879159
  6. 6. Kigozi E, Kasule GW, Musisi K, Lukoye D, Kyobe S, Katabazi FA, et al. Prevalence and patterns of rifampicin and isoniazid resistance conferring mutations in Mycobacterium tuberculosis isolates from Uganda. PLOS ONE. 2018;13: e0198091. pmid:29847567
  7. 7. Bobba S, Khader SA. Rifampicin drug resistance and host immunity in tuberculosis: more than meets the eye. Trends Immunol. 2023 [cited 3 Aug 2023]. pmid:37543504
  8. 8. Khader SA, Partida-Sanchez S, Bell G, Jelley-Gibbs DM, Swain S, Pearl JE, et al. Interleukin 12p40 is required for dendritic cell migration and T cell priming after Mycobacterium tuberculosis infection. J Exp Med. 2006;203: 1805–1815. pmid:16818672
  9. 9. Das S, Marin ND, Esaulova E, Ahmed M, Swain A, Rosa BA, et al. Lung Epithelial Signaling Mediates Early Vaccine-Induced CD4+ T Cell Activation and Mycobacterium tuberculosis Control. mBio. 2021;0: e01468–21. pmid:34253059
  10. 10. Griffiths KL, Ahmed M, Das S, Gopal R, Horne W, Connell TD, et al. Targeting dendritic cells to accelerate T-cell activation overcomes a bottleneck in tuberculosis vaccine efficacy. Nat Commun. 2016;7: 13894. pmid:28004802
  11. 11. Dunlap MD, Howard N, Das S, Scott N, Ahmed M, Prince O, et al. A novel role for C-C motif chemokine receptor 2 during infection with hypervirulent Mycobacterium tuberculosis. Mucosal Immunol. 2018;11: 1727–1742. pmid:30115997
  12. 12. Slight SR, Rangel-Moreno J, Gopal R, Lin Y, Fallert Junecko BA, Mehra S, et al. CXCR5+ T helper cells mediate protective immunity against tuberculosis. J Clin Invest. 2013;123: 712–726. pmid:23281399
  13. 13. Monin L, Griffiths KL, Slight S, Lin Y, Rangel-Moreno J, Khader SA. Immune requirements for protective Th17 recall responses to Mycobacterium tuberculosis challenge. Mucosal Immunol. 2015;8: 1099–1109. pmid:25627812
  14. 14. Esaulova E, Das S, Singh DK, Choreño-Parra JA, Swain A, Arthur L, et al. The immune landscape in tuberculosis reveals populations linked to disease and latency. Cell Host Microbe. 2021;29: 165–178.e8. pmid:33340449
  15. 15. Bohrer AC, Tocheny C, Assmann M, Ganusov VV, Mayer–Barber KD. Cutting Edge: IL-1R1 Mediates Host Resistance to Mycobacterium tuberculosis by Trans-Protection of Infected Cells. J Immunol. 2018;201: 1645–1650. pmid:30068597
  16. 16. 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 Baltim Md 1950. 2010;184: 3326–3330. pmid:20200276
  17. 17. Novikov A, Cardone M, Thompson R, Shenderov K, Kirschman KD, Mayer-Barber KD, et al. Mycobacterium tuberculosis triggers host type I interferon signaling to regulate IL-1β production in human macrophages. J Immunol Baltim Md 1950. 2011;187: 2540–2547. pmid:21784976
  18. 18. Mayer-Barber KD, Andrade BB, Oland SD, Amaral EP, Barber DL, Gonzales J, et al. Host-directed therapy of tuberculosis based on interleukin-1 and type I interferon crosstalk. Nature. 2014;511: 99–103. pmid:24990750
  19. 19. Dorhoi A, Yeremeev V, Nouailles G, Weiner J, Jörg S, Heinemann E, et al. Type I IFN signaling triggers immunopathology in tuberculosis-susceptible mice by modulating lung phagocyte dynamics. Eur J Immunol. 2014;44: 2380–2393. pmid:24782112
  20. 20. Mayer-Barber KD, Yan B. Clash of the Cytokine Titans: counter-regulation of interleukin-1 and type I interferon-mediated inflammatory responses. Cell Mol Immunol. 2017;14: 22–35. pmid:27264686
  21. 21. 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
  22. 22. Berry MPR, Graham CM, McNab FW, Xu Z, Bloch SAA, Oni T, et al. An Interferon-Inducible Neutrophil-Driven Blood Transcriptional Signature in Human Tuberculosis. Nature. 2010;466: 973–977. pmid:20725040
  23. 23. Moreira-Teixeira L, Stimpson PJ, Stavropoulos E, Hadebe S, Chakravarty P, Ioannou M, et al. Type I IFN exacerbates disease in tuberculosis-susceptible mice by inducing neutrophil-mediated lung inflammation and NETosis. Nat Commun. 2020;11: 5566. pmid:33149141
  24. 24. Howard NC, Marin ND, Ahmed M, Rosa BA, Martin J, Bambouskova M, et al. Mycobacterium tuberculosis carrying a rifampicin drug resistance mutation reprograms macrophage metabolism through cell wall lipid changes. Nat Microbiol. 2018;3: 1099–1108. pmid:30224802
  25. 25. Bobba S, Howard NC, Das S, Ahmed M, Tang L, Thirunavukkarasu S, et al. Mycobacterium tuberculosis carrying the rifampicin drug resistance conferring rpoB mutation H445Y is associated with suppressed immunity through type I interferons. mBio. 2023. pmid:37682004
  26. 26. Hatsuzawa K, Sakurai C. Regulatory Mechanism of SNAP23 in Phagosome Formation and Maturation. Yonago Acta Med. 2020;63: 135–145. pmid:32884432
  27. 27. Jo SH, Choi J-A, Lim Y-J, Lee J, Cho S-N, Oh S-M, et al. Calreticulin modulates the intracellular survival of mycobacteria by regulating ER-stress-mediated apoptosis. Oncotarget. 2017;8: 58686–58698. pmid:28938588
  28. 28. Wu Y, Guo Z, Liu F, Yao K, Gao M, Luo Y, et al. Sp110 enhances macrophage resistance to Mycobacterium tuberculosis via inducing endoplasmic reticulum stress and inhibiting anti-apoptotic factors. Oncotarget. 2017;8: 64050–64065. pmid:28969051
  29. 29. Trinchieri G. Type I interferon: friend or foe? J Exp Med. 2010;207: 2053–2063. pmid:20837696
  30. 30. Zemek RM, Chin WL, Fear VS, Wylie B, Casey TH, Forbes C, et al. Temporally restricted activation of IFNβ signaling underlies response to immune checkpoint therapy in mice. Nat Commun. 2022;13: 4895. pmid:35986006
  31. 31. Spaulding E, Fooksman D, Moore JM, Saidi A, Feintuch CM, Reizis B, et al. STING-Licensed Macrophages Prime Type I IFN Production by Plasmacytoid Dendritic Cells in the Bone Marrow during Severe Plasmodium yoelii Malaria. PLOS Pathog. 2016;12: e1005975. pmid:27792766
  32. 32. Scheu S, Dresing P, Locksley RM. Visualization of IFNβ production by plasmacytoid versus conventional dendritic cells under specific stimulation conditions in vivo. Proc Natl Acad Sci U S A. 2008;105: 20416–20421. pmid:19088190
  33. 33. Shi J, Hua L, Harmer D, Li P, Ren G. Cre Driver Mice Targeting Macrophages. Methods Mol Biol Clifton NJ. 2018;1784: 263–275. pmid:29761406
  34. 34. Zhang L, Jiang X, Pfau D, Ling Y, Nathan CF. Type I interferon signaling mediates Mycobacterium tuberculosis–induced macrophage death. J Exp Med. 2020;218: e20200887. pmid:33125053
  35. 35. Desvignes L, Wolf AJ, Ernst JD. Dynamic Roles of Type I and Type II IFNs in Early Infection with Mycobacterium tuberculosis. J Immunol. 2012;188: 6205–6215. pmid:22566567
  36. 36. McNab F, Mayer-Barber K, Sher A, Wack A, O’Garra A. Type I interferons in infectious disease. Nat Rev Immunol. 2015;15: 87–103. pmid:25614319
  37. 37. 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 interleukin-1 receptor antagonist IL-1Ra. Nat Microbiol. 2019;4: 2128–2135. pmid:31611644
  38. 38. Zak DE, Penn-Nicholson A, Scriba TJ, Thompson E, Suliman S, Amon LM, et al. A prospective blood RNA signature for tuberculosis disease risk. Lancet Lond Engl. 2016;387: 2312–2322. pmid:27017310
  39. 39. Singhania A, Verma R, Graham CM, Lee J, Tran T, Richardson M, et al. A modular transcriptional signature identifies phenotypic heterogeneity of human tuberculosis infection. Nat Commun. 2018;9: 2308. pmid:29921861
  40. 40. Mayer-Barber K, Andrade B, Barber D, Hieny S, Feng C, Caspar P, et al. Innate and adaptive interferons suppress IL-1α and IL-1β production by distinct pulmonary myeloid subsets during Mycobacterium tuberculosis infection. Immunity. 2011;35: 1023–1034. pmid:22195750
  41. 41. Manca C, Tsenova L, Bergtold A, Freeman S, Tovey M, Musser JM, et al. Virulence of a Mycobacterium tuberculosis clinical isolate in mice is determined by failure to induce Th1 type immunity and is associated with induction of IFN-α/β. Proc Natl Acad Sci U S A. 2001;98: 5752–5757. pmid:11320211
  42. 42. Antonelli LRV, Gigliotti Rothfuchs A, Gonçalves R, Roffê E, Cheever AW, Bafica A, et al. Intranasal Poly-IC treatment exacerbates tuberculosis in mice through the pulmonary recruitment of a pathogen-permissive monocyte/macrophage population. J Clin Invest. 2010;120: 1674–1682. pmid:20389020
  43. 43. Banks DA, Ahlbrand SE, Hughitt VK, Shah S, Mayer-Barber KD, Vogel SN, et al. Mycobacterium tuberculosis Inhibits Autocrine Type I IFN Signaling to Increase Intracellular Survival. J Immunol. 2019;202: 2348–2359. pmid:30833347
  44. 44. Olson GS, Murray TA, Jahn AN, Mai D, Diercks AH, Gold ES, et al. Type I interferon decreases macrophage energy metabolism during mycobacterial infection. Cell Rep. 2021;35: 109195. pmid:34077724
  45. 45. Kotov DI, Lee OV, Fattinger SA, Langner CA, Guillen JV, Peters JM, et al. Early cellular mechanisms of type I interferon-driven susceptibility to tuberculosis. Cell. 2023 [cited 4 Dec 2023]. pmid:38029747
  46. 46. Reizis B. Plasmacytoid Dendritic Cells: Development, Regulation, and Function. Immunity. 2019;50: 37–50. pmid:30650380
  47. 47. Swiecki M, Colonna M. Type I interferons: diversity of sources, production pathways and effects on immune responses. Curr Opin Virol. 2011;1: 463–475. pmid:22440910
  48. 48. Huang L, Nazarova EV, Tan S, Liu Y, Russell DG. Growth of Mycobacterium tuberculosis in vivo segregates with host macrophage metabolism and ontogeny. J Exp Med. 2018;215: 1135–1152. pmid:29500179
  49. 49. Tan S, Yates RM, Russell DG. Mycobacterium tuberculosis: Readouts of bacterial fitness and the environment within the phagosome. Methods Mol Biol Clifton NJ. 2017;1519: 333–347. pmid:27815891
  50. 50. Stoltzfus CR, Filipek J, Gern BH, Olin BE, Leal JM, Wu Y, et al. CytoMAP: A Spatial Analysis Toolbox Reveals Features of Myeloid Cell Organization in Lymphoid Tissues. Cell Rep. 2020;31: 107523. pmid:32320656
  51. 51. Carow B, Hauling T, Qian X, Kramnik I, Nilsson M, Rottenberg ME. Spatial and temporal localization of immune transcripts defines hallmarks and diversity in the tuberculosis granuloma. Nat Commun. 2019;10: 1823. pmid:31015452
  52. 52. Sawyer AJ, Patrick E, Edwards J, Wilmott JS, Fielder T, Yang Q, et al. Spatial mapping reveals granuloma diversity and histopathological superstructure in human tuberculosis. J Exp Med. 2023;220: e20221392. pmid:36920308
  53. 53. Miyake Y, Kaise H, Isono K, Koseki H, Kohno K, Tanaka M. Protective Role of Macrophages in Noninflammatory Lung Injury Caused by Selective Ablation of Alveolar Epithelial Type II Cells1. J Immunol. 2007;178: 5001–5009. pmid:17404282
  54. 54. McCubbrey AL, Allison KC, Lee-Sherick AB, Jakubzick CV, Janssen WJ. Promoter Specificity and Efficacy in Conditional and Inducible Transgenic Targeting of Lung Macrophages. Front Immunol. 2017;8: 1618. pmid:29225599
  55. 55. Abram CL, Roberge GL, Hu Y, Lowell CA. Comparative analysis of the efficiency and specificity of myeloid-Cre deleting strains using ROSA-EYFP reporter mice. J Immunol Methods. 2014;408: 89–100. pmid:24857755
  56. 56. Green AM, DiFazio R, Flynn JL. IFN-γ from CD4 T cells is essential for host survival and enhances CD8 T cell function during Mycobacterium tuberculosis infection. J Immunol Baltim Md 1950. 2013;190: 270–277. pmid:23233724
  57. 57. Sakai S, Kauffman KD, Schenkel JM, McBerry CC, Mayer-Barber KD, Masopust D, et al. Control of Mycobacterium tuberculosis infection by a subset of lung parenchyma homing CD4 T cells. J Immunol Baltim Md 1950. 2014;192: 2965–2969. pmid:24591367
  58. 58. Schäffer M, Bongartz M, Hoffmann W, Viebahn R. Regulation of Nitric Oxide Synthesis in Wounds by IFN-γ Depends on TNF-α. J Invest Surg. 2006;19: 371–379. pmid:17101606
  59. 59. Pisu D, Huang L, Grenier JK, Russell DG. Dual RNA-Seq of Mtb-Infected Macrophages In Vivo Reveals Ontologically Distinct Host-Pathogen Interactions. Cell Rep. 2020;30: 335–350.e4. pmid:31940480
  60. 60. Yang K, Wu Y, Xie H, Li M, Ming S, Li L, et al. Macrophage-mediated inflammatory response decreases mycobacterial survival in mouse MSCs by augmenting NO production. Sci Rep. 2016;6: 27326. pmid:27251437
  61. 61. Rich EA, Torres M, Sada E, Finegan CK, Hamilton BD, Toossi Z. Mycobacterium tuberculosis (MTB)- stimulated production of nitric oxide by human alveolar macrophages and relationship of nitric oxide production to growth inhibition of MTB. Tuber Lung Dis. 1997;78: 247–255. pmid:10209679
  62. 62. Nozaki Y, Hasegawa Y, Ichiyama S, Nakashima I, Shimokata K. Mechanism of nitric oxide-dependent killing of Mycobacterium bovis BCG in human alveolar macrophages. Infect Immun. 1997;65: 3644–3647. pmid:9284131
  63. 63. Verberk SGS, de Goede KE, Gorki FS, van Dierendonck XAMH, Argüello RJ, Van den Bossche J. An integrated toolbox to profile macrophage immunometabolism. Cell Rep Methods. 2022;2: 100192. pmid:35497494
  64. 64. 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 U S A. 1997;94: 5243–5248. pmid:9144222
  65. 65. Hop HT, Reyes AWB, Arayan LT, Huy TXN, Vu SH, Min W, et al. Interleukin 1 alpha (IL-1α) restricts Brucella abortus 544 survival through promoting lysosomal-mediated killing and NO production in macrophages. Vet Microbiol. 2019;232: 128–136. pmid:31030836
  66. 66. Lahiri N, Shah RR, Layre E, Young D, Ford C, Murray MB, et al. Rifampin Resistance Mutations Are Associated with Broad Chemical Remodeling of Mycobacterium tuberculosis. J Biol Chem. 2016;291: 14248–14256. pmid:27226566
  67. 67. Quigley J, Hughitt VK, Velikovsky CA, Mariuzza RA, El-Sayed NM, Briken V. The Cell Wall Lipid PDIM Contributes to Phagosomal Escape and Host Cell Exit of Mycobacterium tuberculosis. mBio. 2017;8. pmid:28270579
  68. 68. Barczak AK, Avraham R, Singh S, Luo SS, Zhang WR, Bray M-A, et al. Systematic, multiparametric analysis of Mycobacterium tuberculosis intracellular infection offers insight into coordinated virulence. PLOS Pathog. 2017;13: e1006363. pmid:28505176
  69. 69. Augenstreich J, Arbues A, Simeone R, Haanappel E, Wegener A, Sayes F, et al. ESX-1 and phthiocerol dimycocerosates of Mycobacterium tuberculosis act in concert to cause phagosomal rupture and host cell apoptosis. Cell Microbiol. 2017;19. pmid:28095608
  70. 70. Lerner TR, Queval CJ, Lai RP, Russell MRG, Fearns A, Greenwood DJ, et al. Mycobacterium tuberculosis cords within lymphatic endothelial cells to evade host immunity. JCI Insight. 2020;5: e136937. pmid:32369443
  71. 71. Khader SA, Bell GK, Pearl JE, Fountain JJ, Rangel-Moreno J, Cilley GE, et al. IL-23 and IL-17 in the establishment of protective pulmonary CD4+ T cell responses after vaccination and during Mycobacterium tuberculosis challenge. Nat Immunol. 2007;8: 369–377. pmid:17351619
  72. 72. Scott NR, Swanson RV, Al-Hammadi N, Domingo-Gonzalez R, Rangel-Moreno J, Kriel BA, et al. S100A8/A9 regulates CD11b expression and neutrophil recruitment during chronic tuberculosis. J Clin Invest. 130: 3098–3112. pmid:32134742
  73. 73. Liao Y, Smyth GK, Shi W. The R package Rsubread is easier, faster, cheaper and better for alignment and quantification of RNA sequencing reads. Nucleic Acids Res. 2019;47: e47. pmid:30783653
  74. 74. Love MI, Huber W, Anders S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 2014;15: 550. pmid:25516281
  75. 75. Benjamini Y, Hochberg Y. Controlling the False Discovery Rate: A Practical and Powerful Approach to Multiple Testing. J R Stat Soc Ser B Methodol. 1995;57: 289–300.
  76. 76. Yu G, He Q-Y. ReactomePA: an R/Bioconductor package for reactome pathway analysis and visualization. Mol Biosyst. 2016;12: 477–479. pmid:26661513
  77. 77. Korotkevich G, Sukhov V, Budin N, Shpak B, Artyomov MN, Sergushichev A. Fast gene set enrichment analysis. bioRxiv; 2021. p. 060012.
  78. 78. Wickham H. Simple, Consistent Wrappers for Common String Operations. 2022. Available: https://stringr.tidyverse.org/
  79. 79. Kolde R. pheatmap: Pretty Heatmaps. 2019. Available: https://cran.r-project.org/web/packages/pheatmap/index.html
  80. 80. Wickham H. ggplot2. Cham: Springer International Publishing; 2016. https://doi.org/10.1007/978-3-319-24277-4
  81. 81. Carlson M. org.Hs.eg.db. 2019. Available: http://bioconductor.org/packages/org.Hs.eg.db/