Skip to main content
Frontiers in Cellular and Infection Microbiology logoLink to Frontiers in Cellular and Infection Microbiology
. 2014 Mar 5;4:31. doi: 10.3389/fcimb.2014.00031

Cell death paradigms in the pathogenesis of Mycobacterium tuberculosis infection

Dinesh Kumar Parandhaman 1,2, Sujatha Narayanan 1,*
PMCID: PMC3943388  PMID: 24634891

Abstract

Cell death or senescence is a fundamental event that helps maintain cellular homeostasis, shapes the growth of organism, and provides protective immunity against invading pathogens. Decreased or increased cell death is detrimental both in infectious and non-infectious diseases. Cell death is executed both by regulated enzymic reactions and non-enzymic sudden collapse. In this brief review we have tried to summarize various cell death modalities and their impact on the pathogenesis of Mycobacterium tuberculosis.

Keywords: apoptosis, pyroptosis, intrinsic pathway, extrinsic pathway, autophagy, Mycobacterium tuberculosis

Introduction

Cell death is a primordial event in embryogenesis, metamorphosis, and in innate immune response against the invading pathogens. Cell death as a defense mechanism is also documented in the plant kingdom (Kabbage et al., 2013). Cell death is executed in a series of ordered biochemical cascades and is referred as programmed cell death or PCD.

Till early 2000, cell death was discussed as dichotomy in terms of either apoptosis or necrosis. However, with the growth of science many distinct modes of cell death with well-organized signaling cascades were unraveled. Currently, there exists nine different forms of cell death namely apoptosis (Fink and Cookson, 2005), autophagy (Fink and Cookson, 2005), mitoptosis (Chaabane et al., 2012), necrosis (Fink and Cookson, 2005), necroptosis (Galluzzi and Kroemer, 2008), netosis (Remijsen et al., 2011), oncosis (Fink and Cookson, 2005), pyroptosis (Fink and Cookson, 2005), and pyronecrosis (Willingham et al., 2007). It is still a puzzle whether these pathways are different features of the same response or physiologically distinct responses. Apoptosis as an defense mechanism initiates both innate and adaptive immunity (Behar et al., 2010). However, pathogenic organisms have developed mechanisms to modulate apoptosis for their survival. Apoptosis of the infected cells have been reported to be a favorable outcome for the dissemination of infections like Yersinia, Francisella, etc. (Ruckdeschel et al., 1997; Wickstrum et al., 2009). On the contrary, impairment of apoptosis provides a survival niche to many intracellular pathogens including Mycobacterium tuberculosis (Behar et al., 2010), leads to auto immunity, cancer and degenerative disorders (Elmore, 2007). Studies in M. tuberculosis have identified a causal relationship between virulence of the strain and induction of apoptosis. Inhibition of apoptosis favors M. tuberculosis survival in many ways like preventing bactericidal effects, T-cell priming, etc. (Velmurugan et al., 2007). In contrast, a recent report states that apoptosis inducing strains could disseminate M. tuberculosis infection (Aguilo et al., 2013). Necrotic cell death of burdened M. tuberculosis infected cells was shown to pave way for re-infection (Butler et al., 2012). In here, we summarize various apoptotic modalities and their role in the pathogenesis of M. tuberculosis. Furthermore, we share our experience in analyzing these responses in M. tuberculosis infection.

Models of cell death

Apoptosis

First represented in the article by Kerr, Wyllie, and Currie in 1972 (Elmore, 2007). Apoptosis is an energy dependent regulatory process that disintegrates the dying cell by enclosing the cytoplasmic contents inside membrane bound vesicles called apoptotic bodies. These apoptotic bodies are engulfed by the phagocytic cells by a process called efferocytosis thereby efficiently clearing the dying cell without any inflammatory responses (Lee et al., 2009). Three pathways namely extrinsic/ligand-mediated pathway, intrinsic/mitochondrial pathway, and the granzyme B-mediated pathway regulate the process of apoptosis upon activation by physiological or pathological conditions (Elmore, 2007). The major players in apoptosis are caspases, adaptor proteins, tumor necrosis factor (TNF) receptor (TNF-R) super family, and Bcl-2 family of proteins (Strasser et al., 2000). There are three categories of caspases; initiators (caspase-2,-8,-9,-10), effectors or executioners (caspase-3,-6,-7), and inflammatory caspases (caspase-1,-4,-5) (Elmore, 2007). Caspase-activated DNases activate endonuclease that produce the typical internucleosomal DNA cleavage during apoptosis (Strasser et al., 2000). Adapter proteins play a major role in apoptosis as a link between caspases and the TNF-R by mediating homotypic interactions between the domains death domain, the death effector domain, and the caspase recruitment domains (Strasser et al., 2000).

Bcl-2 family of proteins are classified into three types that fall into pro-survival and pro-apoptotic categories based on the amino acid sequence homology to Bcl-2 homology regions BH1–BH4. Pro -survival Bcl-xL, Bcl-w, A1/Bfl-1, Mcl-1, and Boo/Diva have three or four bcl-2 homology regions while the pro-apoptotic members called Bax-like death factors Bax, Bcl-xS, Bak, and Bok/Mtd contain two or three homology regions (Pecina-Slaus, 2010). The third group of proteins Bad, Bik/Nbk, Bid, Hrk/DP5, Bim/Bod, and Blk, etc. that possess only a BH3 region are potent inducers of apoptosis (Strasser et al., 2000).

Apoptotic pathways

  • Extrinsic pathway is initiated by binding of the ligands like TNF-α, FasL, CD95L, TRAIL, etc. to their respective receptors TNFR, Fas/CD95, and DR3 on the cell surface. This activates the initiator caspases such as caspases 8 and 10 that results in the formation and activation of death inducing signaling complex (DISC) that activates caspase 3 (Pecina-Slaus, 2010; Kalimuthu and Se-Kwon, 2013). Caspase 3 activation leads to cleavage of various death substrates that results in the characteristic hallmarks of apoptosis like DNA fragmentation, membrane blebbing, etc. (Kalimuthu and Se-Kwon, 2013).

  • Intrinsic pathway of apoptosis is trigged due to the intracellular death signals. Mitochondrial enzyme endonuclease G, Bcl-2 family of proteins like Bax, Bid, and other mitochondrial proteins AIF, DIABLO [SMAC (second mitochondria-derived activator of caspases)], and cytochrome C plays a major role in this response (Kalimuthu and Se-Kwon, 2013). Upon the stimulus, the BH3-only protein Bid activates Bax and Bak that results in conformational change and oligomerization, forming an oligomeric pore in the outer mitochondrial membrane called permeability transition pores (Ferri and Kroemer, 2001; Kalimuthu and Se-Kwon, 2013). This results in the release of cytochrome C and other pro-apoptotic factors from the mitochondria into the cytosol. Cytochrome C interacts with Apaf and activates caspase-9 forming a multi-protein subunit complex called casposome (apoptosome) comprising cytochrome C, Apaf-1, procaspase-9, and ATP. In the absence of death stimulus, inhibitor of apoptosis family proteins (IAP) inactivates the caspase activity by direct binding. However, upon apoptotic stimuli IAPs are negatively regulated by SMAC and that leads to the activation of caspase-3 (Pecina-Slaus, 2010; Kalimuthu and Se-Kwon, 2013). Furthermore, extrinsic pathway was found to influence the intrinsic pathway of apoptosis by truncation of Bid (Cillessen et al., 2007).

  • Granzyme B-mediated pathway utilizing the extrinsic mode of apoptosis is used by cytotoxic T lymphocytes as a mechanism to kill its target. Besides this, the secretion of pore forming granules containing serine proteases granzyme A and granzyme B also execute apoptosis that is both dependent and independent of caspase activation (Elmore, 2007).

Autophagy

It is a regulated homeostatic response conserved in all living cells degrading their own cytoplasm. Autophagy is a predominant cell survival response that is involved either in nutrient turnover or energy production during stress or removal of long lived cells or to protect against invading intracellular pathogens (Chaabane et al., 2012). Three forms of autophagy namely macroautophagy, microautophagy, and chaperone-mediated autophagy exist. During the autophagy, damaged organelle is lined with an isolation membrane called the phagophore that enlarges forming the double membrane structure called autophagosome. The autophagosome fuses either with late endosomes or lysosomes causing cell death (Levine and Deretic, 2007; Remijsen et al., 2011). Autophagy is regulated by autophagy-related proteins, serine/threonine kinase, mammalian target of rapamycin (mTOR), class I and class III phosphoinositide 3-kinases (PI3Ks) (Levine and Deretic, 2007; Su et al., 2013).

Mitoptosis

Apoptotic changes inside the mitochondria are called mitoptosis. Mitoptosis is still in infancy and no specific factors have been identified. The identification is based on morphological changes like disintegrating cristae, swollen mitochondria, etc. (Chaabane et al., 2012).

Necrosis

Accidental cell death induced due to pathological or physiological conditions are called necrosis. During necrosis, swelling of organelles like endoplasmic reticulum, mitochondria occurs thereby rupturing the plasma membrane. This leaks the intracellular contents of the necrotic cell into the intercellular space causing inflammatory responses (Fink and Cookson, 2005; Chaabane et al., 2012).

Necroptosis

In the year 2008, Hitomi et al. reported that necrosis could be a regulated process of cell death. The activation of serine/threonine kinase RIP1, BH3 only protein Bmf, and mitochondrial dysfunction executes necroptosis (Galluzzi and Kroemer, 2008).

NETosis

In 2004, the findings of Brinkman group unveiled another cell death program named by Steinberg in 2007 called NETosis (Mesa and Vasquez, 2013). One among the defense mechanisms used by neutrophils is the extrusion of intracellular material in the form of extracellular traps (ETs) to the surrounding extracellular medium. This concentrates the microbicidal substances to trap and kill pathogens (Mesa and Vasquez, 2013). Release of ETs by neutrophils is called NETs and mast cells as MCETs. NETs are composed of DNA and histones, and they are resistant to degradation by proteases, insensitive to caspase inhibition and necrostatins (cytoprotective agents) (Mesa and Vasquez, 2013). During NETosis both the nuclear and granular membranes disintegrate leaving the plasma membrane intact (Remijsen et al., 2011). NETosis is activated by pathogens, platelets activated with LPS and in eosinophils (Remijsen et al., 2011). Formation of NET is both nuclear and mitochondrial in origin.

Oncosis

It is the swelling of cells that involves rapid plasma membrane breakdown, and swollen nuclei without internucleosomal DNA fragmentation. Oncosis depletes cellular energy and leads to failure of the ionic pumps in the plasma membrane. It is elicited by agents that disrupt the ATP production of the cell (Fink and Cookson, 2005).

Pyroptosis

Apoptosis in general does not induce an inflammatory response. However, apoptosis in Shigella, Salmonella, Francisella, and Legionella infections produce inflammatory responses that are called as pyroptosis (Carneiro et al., 2009; Lee et al., 2011). Pyroptosis is executed by the formation of inflammasomes by bacterial products involving NLRC 4 (Nod-like receptor—NLR), that activates caspase-1 and the processing of IL-1β and IL-18 cytokines promoting cell death (Fink and Cookson, 2005; Carneiro et al., 2009).

Pyronecrosis

Cathepsin B-dependent apoptosis that is independent of caspase-1 activation and inflammasome formation is called pyronecrosis. This mode of apoptosis is observed in shigellosis (Willingham et al., 2007; Carneiro et al., 2009).

Other apoptotic models

  • Tumor suppressor protein 53 (TP53) induced apoptosis involves the transcriptional induction of redox proteins, generation of reactive oxygen species, and oxidative degradation of mitochondrial components that results in cell death. TP53 was shown to transcriptionally regulate proapoptotic proteins like Bax and NOXA (Yamada et al., 2002).

  • NF-kB expression is implicated in the survival of living cells. NF-kB family contains five proteins namely c-Rel, RelA, RelB, p50/p105, and p52/p100. NF-kB as a homo or hetero dimers bind to the kB sites on their target DNA and regulate their expression (Barkett and Gilmore, 1999). NF-kB is activated by various stimuli like pathogens, mitogens, proinflammatory cytokines, etc. It plays a major role in immune responses and affects the expression of genes c-IAP-1 and c-IAP-2, Fas ligand, c-myc, p53, etc. involved in apoptosis (Zhang and Ghosh, 2001). Two TNF receptors TNFRSF8 and TNFRSF9 were shown to promote apoptosis, former activating, and latter inactivating NF-kB expression (Wang et al., 2008).

Apoptosis and mycobacterium tuberculosis

M. tuberculosis infections with virulent strains have been reported to inhibit macrophage apoptosis (Behar et al., 2010). Varied mechanisms of apoptotic suppression have been reported in M. tuberculosis infections (Table 1) unraveling the tactics of this pathogen to generate a protective niche inside the host. Among the various cell death modalities described above, only three apoptotic responses were documented in M. tuberculosis infection namely apoptosis (nuoG, SecA2, pknE, lpqH, esxA (ESAT-6), PE_PGRS33, pstS-1, Rv3654c, and Rv3655c), pyroptosis (zmp1, Rv3364c), and autophagy (eis) (Hinchey et al., 2007; Velmurugan et al., 2007; Jayakumar et al., 2008; Master et al., 2008; Sanchez et al., 2009, 2012; Danelishvili et al., 2010, 2012; Shin et al., 2010).

Table 1.

Apoptotic mechanisms in the pathogenesis of M. tuberculosis.

S.no Mechanisms of apoptosis Year References
1 Treatment of macrophages post-infection with exogenous ATP reduces viability 1994 Molloy et al., 1994
2 Extrinsic apoptosis 1997 Keane et al., 1997
3 Virulent strains induce IL-10-dependent sTNFR2 forming inactive TNF-α-TNFR2 complex 1998 Fratazzi et al., 1999
4 Granulysin and perforin reduce the viability of M. tuberculosis 1998 Stenger et al., 1998
5 Treatment of Fas ligand post-infection reduces the viability 1998 Oddo et al., 1998
6 Degree of apoptosis is strain-dependent 2000 Keane et al., 2000
7 ManLam prevents apoptosis by altering Ca2+ levels 2000 Rojas et al., 2000
8 M. tuberculosis apoptosis down regulates CD14 2000 Santucci et al., 2000
9 Apoptosis of avirulent strains dependent on group IV cytosolic phospholipase A2 and TNF-α 2001 Duan et al., 2001
10 Reduced viability using exogenous ATP is executed using P2X7 receptor 2001 Fairbairn et al., 2001
11 Anti-apoptotic Mcl-1expression by virulent strains decreases apoptosis 2003 Sly et al., 2003
12 Detour pathway of antigen presentation 2003 Schaible et al., 2003
13 19 kDa lipoprotein induces apoptosis by TLR2 signaling 2003 Lopez et al., 2003
14 Virulent strains induce necrosis 2006 Park et al., 2006
15 Methyl glyoxal plays role in apoptosis 2006 Rachman et al., 2006
16 TLR-2-mediated activation of NF-kB and c-FLIP protects infected cells from FasL-induced apoptosis 2006 Loeuillet et al., 2006
17 PE_PGRS33 induces TNF-α secretion using TLR-2 signaling and genetic alterations in PE_PGRS33 decreases TNF-α secretion 2006 Basu et al., 2007
18 High MOI induces TNF-α independent apoptosis leading to mycobacterial spread 2007 Lee et al., 2006
19 Higher MOI leads to caspase independent apoptosis involving both mitochondria and lysosymes 2007 O'Sullivan et al., 2007
20 ESAT-6 induces apoptosis 2007 Derrick and Morris, 2007
21 Bystander apoptosis elicited by avirulent strains are independent of TNF-α,Fas,TRAIL, TGF-β, TLR2, and MyD88 2008 Kelly et al., 2008
22 Virulent strains prevents apoptotic envelope formation leading to necrosis 2008 Gan et al., 2008
23 Virulent strains produce more lipoxinA4 promoting necrosis and avirulent strain induces PGE2 that prevents necrosis 2008 Chen et al., 2008
24 Formation of NETs unable to kill M. tuberculosis 2008 Ramos-Kichik et al., 2009
25 Prevents pyroptosis using zmp1 by inhibiting inflammasome formation required for IL-1β secretion 2008 Master et al., 2008
26 pstS1 induces TNF-α, FasL,Fas TNFR1, TNFR2, and TLR-2 mediated apoptosis 2008 Sanchez et al., 2009
27 TNF-α-mediated caspase-8 apoptosis by p38MAPK, ASK-1, and FLIPS degradation 2009 Kundu et al., 2009
28 Virulent strains inhibit plasma membrane repair promoting necrosis 2009 Divangahi et al., 2009
29 Neutrophil activation leads to ectososme release 2010 Gonzalez-Cano et al., 2010
30 nuoG neutralize NOX2 derived ROS inhibiting extrinsic apoptosis 2010 Miller et al., 2010
31 Rv3654c and Rv3655c genes prevent extrinsic apoptosis 2010 Danelishvili et al., 2010
32 eis is involved in suppressing autophagy in a redox dependent JNK activation 2010 Shin et al., 2010
33 Higher MOI induces host cell lipolysis and PHOPR kinase plays a role in this response 2011 Divangahi et al., 2009
34 PE_PGRS33 interacts with host mitochondria and probably involved in primary necrosis 2011 Cadieux et al., 2011
35 Dendritic cells undergo caspase independent apoptosis 2011 Ryan et al., 2011
36 ROS mediated necrosis as a survival strategy in neutrophils 2012 Corleis et al., 2012
37 ESAT-6 induced apoptosis is regulated by BAT3 2012 Grover and Izzo, 2012
38 Rv3364c prevents pyroptosis by inhibiting cathepsinG 2012 Danelishvili et al., 2012
39 pknE inhibits various modes of apoptosis in response to nitric oxide stress of the macrophages 2012 Kumar and Narayanan, 2012
40 nuoG mutant reveals decreased neutrophil apoptosis reduces CD4 T cell activation 2012 Blomgran et al., 2012
41 Virulence determines cytotoxicity whereas strain characteristics determine the mode of cell death 2012 Butler et al., 2012
42 ESAT-6 is involved in inhibiting autophagy 2012 Romagnoli et al., 2012
43 sigH or its regulated genes suppresses apoptosis, modulates innate immune responses, and reduces chemotaxis 2012 Dutta et al., 2012
44 Infection with avirulent mycobacteria induces mitochondrial exhaustion while virulent promotes mitochondrial function thereby increasing ATP synthesis 2012 Jamwal et al., 2013
45 LpqH induces both extrinsic and intrinsic apoptosis 2012 Sanchez et al., 2012
46 Virulent Mycobacterial strains induce apoptosis by ESX-1 system and colonize new cells 2013 Aguilo et al., 2013
47 Validation of burst size hypothesis in in vivo model 2013 Repasy et al., 2013
48 pknE involved in the copathogenesis of HIV/TB coinfection 2014 Parandhaman et al., 2014

This table illustrates varied apoptotic mechanisms identified in the pathogenesis of M. tuberculosis. The abbreviations MOI denote multiplicity of infection, ManLam, mannosylated lipoarabinomannan; PGE2, prostaglandinE2; ROS, reactive oxygen species; ATP, adenosine tri phosphate.

Serine/threonine protein kinases (STPK)

Two component signaling systems were considered as the standalone mechanism of signaling in prokaryotes in response to environmental cues. However with the availability of various molecular techniques serine, threonine, and tyrosine mediated phosphorylation events unique to eukaryotes were documented in pathogenic prokaryotes like M. tuberculosis, Streptococcus species, Staphylococcus spp, Pseudomonas spp, etc. (Chao et al., 2009; Chakraborti et al., 2011). Among the 11 STPKs that M. tuberculosis encodes, only five of them pknE, pknG, pknH, pknI, and pknK were reported to support intracellular survival (Walburger et al., 2004; Papavinasasundaram et al., 2005; Jayakumar et al., 2008; Gopalaswamy et al., 2009; Malhotra et al., 2010). Our data for the first time proved that PknE was the only STPK to inhibit apoptosis (Jayakumar et al., 2008).

pknE in innate immunity

The function of pknE was established from our studies using the deletion mutant ΔpknE generated using specialized transduction. Deletion of pknE had reduced intracellular survival, increased apoptosis, and reduced proinflammatory responses (Jayakumar et al., 2008). Subsequent molecular pathogenesis studies revealed that the deletion of pknE promotes macrophage cell death dependent on intrinsic pathway of apoptosis, TP53, and Arg2. This apoptosis was independent of TNF-α, iNOS, Akt, Arg1, and pro-inflammatory cytokines (Kumar and Narayanan, 2012). M. tuberculosis encounters reactive nitrogen and oxygen intermediates inside the macrophages as one among the host defenses. Characterization of the promoter of the pknE gene showed its elevated expression during nitric oxide (NO) stress (Jayakumar et al., 2008). Macrophage experiments performed using NO donor sodium nitroprusside to mimic the host microbicidal activity confirmed that, pknE in response to NO stress suppresses innate immune responses (Kumar and Narayanan, 2012). In vitro studies carried with the deletion mutant showed defective growth in pH 7.0 and lysozyme (a cell wall-damaging agent) with better survival in pH 5.5, SDS (surfactant stress), and kanamycin (a second-line anti-tuberculosis drug). ΔpknE was reduced in cell size during growth in liquid media and exhibited hypervirulence in a guinea pig model of infection (Kumar et al., 2012). The data from the in vitro studies highlighted the role of pknE in adaptive responses of M. tuberculosis. Recently we reported that, deletion of pknE results in defective phosphorylation kinetics of MAPKs (p38MAPK, Erk½, and SAPK/JNK) and their transcription factors ATF-2 and c-JUN. Deletion of pknE also revealed crosstalks in the host macrophages where Erk½ signaling was found to be influenced by SAPK/JNK and p38 pathways independently. Modulations in intra cellular signaling altered the expression of coreceptors CCR5 and CXCR4 in macrophages infected with the deletion mutant of pknE that were authenticated using HIV tropic strains (Parandhaman et al., 2014). For the first time, our data showed that difference in apoptosis and intracellular signaling events, and the virulence capacity of the M. tuberculosis strain could influence the copathogenesis of HIV infection (Parandhaman et al., 2014). Collectively the reports show that pknE has a role suppression of innate immunity and help M. tuberculosis to adapt to the different environmental condition that it encounters.

Conclusion

Molecular techniques have revolutionized our understanding of pathogenic organisms and their interactions with the immune system. Pathogenic organisms have evolved host mimicking properties and utilize the host responses for their own survival and propagation. This review has addressed the various mechanisms of cell death that is vital for initiating an innate and adaptive immunity against the invading pathogen. As novel cell death paradigms evolve, it adds to the complexity of how temporally and spatially the immune system coordinates these responses. Most of the cell death models described here disrupt the energy source of the cell, mitochondria indicating whether these paradigms are interconnected response of a single biochemical event and this still remains a puzzle. Adding complexity to this conundrum is that, pathogenic organisms like M. tuberculosis is able to inhibit the various apoptotic models that were discovered so far. This arise the question whether M. tuberculosis by educating itself avoids cell death or has antigens that are poor inducers of cell death and that await further studies.

Conflict of interest statement

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

References

  1. Aguilo J. I., Alonso H., Uranga S., Marinova D., Arbues A., De Martino A., et al. (2013). ESX-1-induced apoptosis is involved in cell-to-cell spread of Mycobacterium tuberculosis. Cell. Microbiol. 15, 1994–2005 10.1111/cmi.12169 [DOI] [PubMed] [Google Scholar]
  2. Barkett M., Gilmore T. D. (1999). Control of apoptosis by Rel/NF-kappaB transcription factors. Oncogene 18, 6910–6924 10.1038/sj.onc.1203238 [DOI] [PubMed] [Google Scholar]
  3. Basu S., Pathak S. K., Banerjee A., Pathak S., Bhattacharyya A., Yang Z., et al. (2007). Execution of macrophage apoptosis by PE_PGRS33 of Mycobacterium tuberculosis is mediated by Toll-like receptor 2-dependent release of tumor necrosis factor-alpha. J. Biol. Chem. 282, 1039–1050 10.1074/jbc.M604379200 [DOI] [PubMed] [Google Scholar]
  4. Behar S. M., Divangahi M., Remold H. G. (2010). Evasion of innate immunity by Mycobacterium tuberculosis: is death an exit strategy? Nat. Rev. Microbiol. 8, 668–674 10.1038/nrmicro2387 [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Blomgran R., Desvignes L., Briken V., Ernst J. D. (2012). Mycobacterium tuberculosis inhibits neutrophil apoptosis, leading to delayed activation of naive CD4 T cells. Cell Host Microbe 11, 81–90 10.1016/j.chom.2011.11.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Butler R. E., Brodin P., Jang J., Jang M. S., Robertson B. D., Gicquel B., et al. (2012). The balance of apoptotic and necrotic cell death in Mycobacterium tuberculosis infected macrophages is not dependent on bacterial virulence. PLoS ONE 7:e47573 10.1371/journal.pone.0047573 [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Cadieux N., Parra M., Cohen H., Maric D., Morris S. L., Brennan M. J. (2011). Induction of cell death after localization to the host cell mitochondria by the Mycobacterium tuberculosis PE_PGRS33 protein. Microbiology 157, 793–804 10.1099/mic.0.041996-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Carneiro L. A., Travassos L. H., Soares F., Tattoli I., Magalhaes J. G., Bozza M. T., et al. (2009). Shigella induces mitochondrial dysfunction and cell death in nonmyleoid cells. Cell Host Microbe 5, 123–136 10.1016/j.chom.2008.12.011 [DOI] [PubMed] [Google Scholar]
  9. Chaabane W., User S. D., El-Gazzah M., Jaksik R., Sajjadi E., Rzeszowska-Wolny J., et al. (2012). Autophagy, apoptosis, mitoptosis and necrosis: interdependence between those pathways and effects on cancer. Arch. Immunol. Ther. Exp. (Warsz.) 61, 43–58 10.1007/s00005-012-0205-y [DOI] [PubMed] [Google Scholar]
  10. Chakraborti P. K., Matange N., Nandicoori V. K., Singh Y., Tyagi J. S., Visweswariah S. S. (2011). Signalling mechanisms in Mycobacteria. Tuberculosis (Edinb.) 91, 432–440 10.1016/j.tube.2011.04.005 [DOI] [PubMed] [Google Scholar]
  11. Chao J., Wong D., Zheng X., Poirier V., Bach H., Hmama Z., et al. (2009). Protein kinase and phosphatase signaling in Mycobacterium tuberculosis physiology and pathogenesis. Biochim. Biophys. Acta 1804, 620–627 10.1016/j.bbapap.2009.09.008 [DOI] [PubMed] [Google Scholar]
  12. Chen M., Divangahi M., Gan H., Shin D. S., Hong S., Lee D. M., et al. (2008). Lipid mediators in innate immunity against tuberculosis: opposing roles of PGE2 and LXA4 in the induction of macrophage death. J. Exp. Med. 205, 2791–2801 10.1084/jem.20080767 [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Cillessen S. A., Hess C. J., Hooijberg E., Castricum K. C., Kortman P., Denkers F., et al. (2007). Inhibition of the intrinsic apoptosis pathway downstream of caspase-9 activation causes chemotherapy resistance in diffuse large B-cell lymphoma. Clin. Cancer Res. 13, 7012–7021 10.1158/1078-0432.CCR-06-2891 [DOI] [PubMed] [Google Scholar]
  14. Corleis B., Korbel D., Wilson R., Bylund J., Chee R., Schaible U. E. (2012). Escape of Mycobacterium tuberculosis from oxidative killing by neutrophils. Cell. Microbiol. 14, 1109–1121 10.1111/j.1462-5822.2012.01783.x [DOI] [PubMed] [Google Scholar]
  15. Danelishvili L., Everman J. L., McNamara M. J., Bermudez L. E. (2012). Inhibition of the plasma-membrane-associated serine protease cathepsin G by Mycobacterium tuberculosis Rv3364c suppresses caspase-1 and pyroptosis in macrophages. Front. Microbiol. 2:281 10.3389/fmicb.2011.00281 [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Danelishvili L., Yamazaki Y., Selker J., Bermudez L. E. (2010). Secreted Mycobacterium tuberculosis Rv3654c and Rv3655c proteins participate in the suppression of macrophage apoptosis. PLoS ONE 5:e10474 10.1371/journal.pone.0010474 [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Derrick S. C., Morris S. L. (2007). The ESAT6 protein of Mycobacterium tuberculosis induces apoptosis of macrophages by activating caspase expression. Cell. Microbiol. 9, 1547–1555 10.1111/j.1462-5822.2007.00892.x [DOI] [PubMed] [Google Scholar]
  18. Divangahi M., Chen M., Gan H., Desjardins D., Hickman T. T., Lee D. M., et al. (2009). Mycobacterium tuberculosis evades macrophage defenses by inhibiting plasma membrane repair. Nat. Immunol. 10, 899–906 10.1038/ni.1758 [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Duan L., Gan H., Arm J., Remold H. G. (2001). Cytosolic phospholipase A2 participates with TNF-alpha in the induction of apoptosis of human macrophages infected with Mycobacterium tuberculosis H37Ra. J. Immunol. 166, 7469–7476 [DOI] [PubMed] [Google Scholar]
  20. Dutta N. K., Mehra S., Martinez A. N., Alvarez X., Renner N. A., Morici L. A., et al. (2012). The stress-response factor SigH modulates the interaction between Mycobacterium tuberculosis and host phagocytes. PLoS ONE 7:e28958 10.1371/journal.pone.0028958 [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Elmore S. (2007). Apoptosis: a review of programmed cell death. Toxicol. Pathol. 35, 495–516 10.1080/01926230701320337 [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Fairbairn I. P., Stober C. B., Kumararatne D. S., Lammas D. A. (2001). ATP-mediated killing of intracellular mycobacteria by macrophages is a P2X(7)-dependent process inducing bacterial death by phagosome-lysosome fusion. J. Immunol. 167, 3300–3307 [DOI] [PubMed] [Google Scholar]
  23. Ferri K. F., Kroemer G. (2001). Organelle-specific initiation of cell death pathways. Nat. Cell Biol. 3, E255–E263 10.1038/ncb1101-e255 [DOI] [PubMed] [Google Scholar]
  24. Fink S. L., Cookson B. T. (2005). Apoptosis, pyroptosis, and necrosis: mechanistic description of dead and dying eukaryotic cells. Infect. Immun. 73, 1907–1916 10.1128/IAI.73.4.1907-1916.2005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Fratazzi C., Arbeit R. D., Carini C., Balcewicz-Sablinska M. K., Keane J., Kornfeld H., et al. (1999). Macrophage apoptosis in mycobacterial infections. J. Leukoc. Biol. 66, 763–764 [DOI] [PubMed] [Google Scholar]
  26. Galluzzi L., Kroemer G. (2008). Necroptosis: a specialized pathway of programmed necrosis. Cell 135, 1161–1163 10.1016/j.cell.2008.12.004 [DOI] [PubMed] [Google Scholar]
  27. Gan H., Lee J., Ren F., Chen M., Kornfeld H., Remold H. G. (2008). Mycobacterium tuberculosis blocks crosslinking of annexin-1 and apoptotic envelope formation on infected macrophages to maintain virulence. Nat. Immunol. 9, 1189–1197 10.1038/ni.1654 [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Gonzalez-Cano P., Mondragon-Flores R., Sanchez-Torres L. E., Gonzalez-Pozos S., Silva-Miranda M., Monroy-Ostria A., et al. (2010). Mycobacterium tuberculosis H37Rv induces ectosome release in human polymorphonuclear neutrophils. Tuberculosis (Edinb.) 90, 125–134 10.1016/j.tube.2010.01.002 [DOI] [PubMed] [Google Scholar]
  29. Gopalaswamy R., Narayanan S., Chen B., Jacobs W. R., Av-Gay Y. (2009). The serine/threonine protein kinase PknI controls the growth of Mycobacterium tuberculosis upon infection. FEMS Microbiol. Lett. 295, 23–29 10.1111/j.1574-6968.2009.01570.x [DOI] [PubMed] [Google Scholar]
  30. Grover A., Izzo A. A. (2012). BAT3 regulates Mycobacterium tuberculosis protein ESAT-6-mediated apoptosis of macrophages. PLoS ONE 7:e40836 10.1371/journal.pone.0040836 [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Hinchey J., Lee S., Jeon B. Y., Basaraba R. J., Venkataswamy M. M., Chen B., et al. (2007). Enhanced priming of adaptive immunity by a proapoptotic mutant of Mycobacterium tuberculosis. J. Clin. Invest. 117, 2279–2288 10.1172/JCI31947 [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Jamwal S., Midha M. K., Verma H. N., Basu A., Rao K. V., Manivel V. (2013). Characterizing virulence-specific perturbations in the mitochondrial function of macrophages infected with Mycobacterium tuberculosis. Sci. Rep. 3:1328 10.1038/srep01328 [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Jayakumar D., Jacobs W. R., Jr., Narayanan S. (2008). Protein kinase E of Mycobacterium tuberculosis has a role in the nitric oxide stress response and apoptosis in a human macrophage model of infection. Cell. Microbiol. 10, 365–374 10.1111/j.1462-5822.2007.01049.x [DOI] [PubMed] [Google Scholar]
  34. Kabbage M., Williams B., Dickman M. B. (2013). Cell death control: the interplay of apoptosis and autophagy in the pathogenicity of Sclerotinia sclerotiorum. PLoS Pathog. 9:e1003287 10.1371/journal.ppat.1003287 [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Kalimuthu S., Se-Kwon K. (2013). Cell survival and apoptosis signaling as therapeutic target for cancer: marine bioactive compounds. Int. J. Mol. Sci. 14, 2334–2354 10.3390/ijms14022334 [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Keane J., Balcewicz-Sablinska M. K., Remold H. G., Chupp G. L., Meek B. B., Fenton M. J., et al. (1997). Infection by Mycobacterium tuberculosis promotes human alveolar macrophage apoptosis. Infect. Immun. 65, 298–304 [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Keane J., Remold H. G., Kornfeld H. (2000). Virulent Mycobacterium tuberculosis strains evade apoptosis of infected alveolar macrophages. J. Immunol. 164, 2016–2020 [DOI] [PubMed] [Google Scholar]
  38. Kelly D. M., Ten Bokum A. M., O'Leary S. M., O'Sullivan M. P., Keane J. (2008). Bystander macrophage apoptosis after Mycobacterium tuberculosis H37Ra infection. Infect. Immun. 76, 351–360 10.1128/IAI.00614-07 [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Kumar D., Narayanan S. (2012). pknE, a serine/threonine kinase of Mycobacterium tuberculosis modulates multiple apoptotic paradigms. Infect. Genet. Evol. 12, 737–747 10.1016/j.meegid.2011.09.008 [DOI] [PubMed] [Google Scholar]
  40. Kumar D., Palaniyandi K., Challu V. K., Kumar P., Narayanan S. (2012). PknE, a serine/threonine protein kinase from Mycobacterium tuberculosis has a role in adaptive responses. Arch. Microbiol. 195, 75–80 10.1007/s00203-012-0848-4 [DOI] [PubMed] [Google Scholar]
  41. Kundu M., Pathak S. K., Kumawat K., Basu S., Chatterjee G., Pathak S., et al. (2009). A TNF- and c-Cbl-dependent FLIP(S)-degradation pathway and its function in Mycobacterium tuberculosis-induced macrophage apoptosis. Nat. Immunol. 10, 918–926 10.1038/ni.1754 [DOI] [PubMed] [Google Scholar]
  42. Lee J., Hartman M., Kornfeld H. (2009). Macrophage apoptosis in tuberculosis. Yonsei Med. J. 50, 1–11 10.3349/ymj.2009.50.1.1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Lee J., Remold H. G., Ieong M. H., Kornfeld H. (2006). Macrophage apoptosis in response to high intracellular burden of Mycobacterium tuberculosis is mediated by a novel caspase-independent pathway. J. Immunol. 176, 4267–4274 [DOI] [PubMed] [Google Scholar]
  44. Lee J., Repasy T., Papavinasasundaram K., Sassetti C., Kornfeld H. (2011). Mycobacterium tuberculosis induces an atypical cell death mode to escape from infected macrophages. PLoS ONE 6:e18367 10.1371/journal.pone.0018367 [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Levine B., Deretic V. (2007). Unveiling the roles of autophagy in innate and adaptive immunity. Nat. Rev. Immunol. 7, 767–777 10.1038/nri2161 [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Loeuillet C., Martinon F., Perez C., Munoz M., Thome M., Meylan P. R. (2006). Mycobacterium tuberculosis subverts innate immunity to evade specific effectors. J. Immunol. 177, 6245–6255 [DOI] [PubMed] [Google Scholar]
  47. Lopez M., Sly L. M., Luu Y., Young D., Cooper H., Reiner N. E. (2003). The 19-kDa Mycobacterium tuberculosis protein induces macrophage apoptosis through Toll-like receptor-2. J. Immunol. 170, 2409–2416 [DOI] [PubMed] [Google Scholar]
  48. Malhotra V., Arteaga-Cortes L. T., Clay G., Clark-Curtiss J. E. (2010). Mycobacterium tuberculosis protein kinase K confers survival advantage during early infection in mice and regulates growth in culture and during persistent infection: implications for immune modulation. Microbiology 156, 2829–2841 10.1099/mic.0.040675-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Master S. S., Rampini S. K., Davis A. S., Keller C., Ehlers S., Springer B., et al. (2008). Mycobacterium tuberculosis prevents inflammasome activation. Cell Host Microbe 3, 224–232 10.1016/j.chom.2008.03.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Mesa M. A., Vasquez G. (2013). NETosis. Autoimmune Dis. 2013:651497 10.1155/2013/651497 [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Miller J. L., Velmurugan K., Cowan M. J., Briken V. (2010). The type I NADH dehydrogenase of Mycobacterium tuberculosis counters phagosomal NOX2 activity to inhibit TNF-alpha-mediated host cell apoptosis. PLoS Pathog. 6:e1000864 10.1371/journal.ppat.1000864 [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Molloy A., Laochumroonvorapong P., Kaplan G. (1994). Apoptosis, but not necrosis, of infected monocytes is coupled with killing of intracellular bacillus Calmette-Guerin. J. Exp. Med. 180, 1499–1509 10.1084/jem.180.4.1499 [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Oddo M., Renno T., Attinger A., Bakker T., Macdonald H. R., Meylan P. R. (1998). Fas ligand-induced apoptosis of infected human macrophages reduces the viability of intracellular Mycobacterium tuberculosis. J. Immunol. 160, 5448–5454 [PubMed] [Google Scholar]
  54. O'Sullivan M. P., O'Leary S., Kelly D. M., Keane J. (2007). A caspase-independent pathway mediates macrophage cell death in response to Mycobacterium tuberculosis infection. Infect. Immun. 75, 1984–1993 10.1128/IAI.01107-06 [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Papavinasasundaram K. G., Chan B., Chung J. H., Colston M. J., Davis E. O., Av-Gay Y. (2005). Deletion of the Mycobacterium tuberculosis pknH gene confers a higher bacillary load during the chronic phase of infection in BALB/c mice. J. Bacteriol. 187, 5751–5760 10.1128/JB.187.16.5751-5760.2005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Parandhaman D. K., Hanna L. E., Narayanan S. (2014). PknE, a serine/threonine protein kinase of Mycobacterium tuberculosis initiates survival crosstalk that also impacts HIV coinfection. PLoS ONE 9:e83541 10.1371/journal.pone.0083541 [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  57. Park J. S., Tamayo M. H., Gonzalez-Juarrero M., Orme I. M., Ordway D. J. (2006). Virulent clinical isolates of Mycobacterium tuberculosis grow rapidly and induce cellular necrosis but minimal apoptosis in murine macrophages. J. Leukoc. Biol. 79, 80–86 10.1189/jlb.0505250 [DOI] [PubMed] [Google Scholar]
  58. Pecina-Slaus N. (2010). Wnt signal transduction pathway and apoptosis: a review. Cancer Cell Int. 10, 22 10.1186/1475-2867-10-22 [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Rachman H., Kim N., Ulrichs T., Baumann S., Pradl L., Nasser Eddine A., et al. (2006). Critical role of methylglyoxal and AGE in mycobacteria-induced macrophage apoptosis and activation. PLoS ONE 1:e29 10.1371/journal.pone.0000029 [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Ramos-Kichik V., Mondragon-Flores R., Mondragon-Castelan M., Gonzalez-Pozos S., Muniz-Hernandez S., Rojas-Espinosa O., et al. (2009). Neutrophil extracellular traps are induced by Mycobacterium tuberculosis. Tuberculosis (Edinb.) 89, 29–37 10.1016/j.tube.2008.09.009 [DOI] [PubMed] [Google Scholar]
  61. Remijsen Q., Kuijpers T. W., Wirawan E., Lippens S., Vandenabeele P., Vanden Berghe T. (2011). Dying for a cause: NETosis, mechanisms behind an antimicrobial cell death modality. Cell Death Differ. 18, 581–588 10.1038/cdd.2011.1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Repasy T., Lee J., Marino S., Martinez N., Kirschner D. E., Hendricks G., et al. (2013). Intracellular bacillary burden reflects a burst size for Mycobacterium tuberculosis in vivo. PLoS Pathog. 9:e1003190 10.1371/journal.ppat.1003190 [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Rojas M., Garcia L. F., Nigou J., Puzo G., Olivier M. (2000). Mannosylated lipoarabinomannan antagonizes Mycobacterium tuberculosis-induced macrophage apoptosis by altering Ca+2-dependent cell signaling. J. Infect. Dis. 182, 240–251 10.1086/315676 [DOI] [PubMed] [Google Scholar]
  64. Romagnoli A., Etna M. P., Giacomini E., Pardini M., Remoli M. E., Corazzari M., et al. (2012). ESX-1 dependent impairment of autophagic flux by Mycobacterium tuberculosis in human dendritic cells. Autophagy 8, 1357–1370 10.4161/auto.20881 [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Ruckdeschel K., Roggenkamp A., Lafont V., Mangeat P., Heesemann J., Rouot B. (1997). Interaction of Yersinia enterocolitica with macrophages leads to macrophage cell death through apoptosis. Infect. Immun. 65, 4813–4821 [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Ryan R. C., O'Sullivan M. P., Keane J. (2011). Mycobacterium tuberculosis infection induces non-apoptotic cell death of human dendritic cells. BMC Microbiol. 11:237 10.1186/1471-2180-11-237 [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Sanchez A., Espinosa P., Esparza M. A., Colon M., Bernal G., Mancilla R. (2009). Mycobacterium tuberculosis 38-kDa lipoprotein is apoptogenic for human monocyte-derived macrophages. Scand. J. Immunol. 69, 20–28 10.1111/j.1365-3083.2008.02193.x [DOI] [PubMed] [Google Scholar]
  68. Sanchez A., Espinosa P., Garcia T., Mancilla R. (2012). The 19 kDa Mycobacterium tuberculosis lipoprotein (LpqH) induces macrophage apoptosis through extrinsic and intrinsic pathways: a role for the mitochondrial apoptosis-inducing factor. Clin. Dev. Immunol. 2012:950503 10.1155/2012/950503 [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Santucci M. B., Amicosante M., Cicconi R., Montesano C., Casarini M., Giosue S., et al. (2000). Mycobacterium tuberculosis-induced apoptosis in monocytes/macrophages: early membrane modifications and intracellular mycobacterial viability. J. Infect. Dis. 181, 1506–1509 10.1086/315371 [DOI] [PubMed] [Google Scholar]
  70. Schaible U. E., Winau F., Sieling P. A., Fischer K., Collins H. L., Hagens K., et al. (2003). Apoptosis facilitates antigen presentation to T lymphocytes through MHC-I and CD1 in tuberculosis. Nat. Med. 9, 1039–1046 10.1038/nm906 [DOI] [PubMed] [Google Scholar]
  71. Shin D. M., Jeon B. Y., Lee H. M., Jin H. S., Yuk J. M., Song C. H., et al. (2010). Mycobacterium tuberculosis eis regulates autophagy, inflammation, and cell death through redox-dependent signaling. PLoS Pathog. 6:e1001230 10.1371/journal.ppat.1001230 [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Sly L. M., Hingley-Wilson S. M., Reiner N. E., McMaster W. R. (2003). Survival of Mycobacterium tuberculosis in host macrophages involves resistance to apoptosis dependent upon induction of antiapoptotic Bcl-2 family member Mcl-1. J. Immunol. 170, 430–437 [DOI] [PubMed] [Google Scholar]
  73. Stenger S., Hanson D. A., Teitelbaum R., Dewan P., Niazi K. R., Froelich C. J., et al. (1998). An antimicrobial activity of cytolytic T cells mediated by granulysin. Science 282, 121–125 10.1126/science.282.5386.121 [DOI] [PubMed] [Google Scholar]
  74. Strasser A., O'Connor L., Dixit V. M. (2000). Apoptosis signaling. Annu. Rev. Biochem. 69, 217–245 10.1146/annurev.biochem.69.1.217 [DOI] [PubMed] [Google Scholar]
  75. Su M., Mei Y., Sinha S. (2013). Role of the crosstalk between autophagy and apoptosis in cancer. J. Oncol. 2013:102735 10.1155/2013/102735 [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Velmurugan K., Chen B., Miller J. L., Azogue S., Gurses S., Hsu T., et al. (2007). Mycobacterium tuberculosis nuoG is a virulence gene that inhibits apoptosis of infected host cells. PLoS Pathog. 3:e110 10.1371/journal.ppat.0030110 [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Walburger A., Koul A., Ferrari G., Nguyen L., Prescianotto-Baschong C., Huygen K., et al. (2004). Protein kinase G from pathogenic mycobacteria promotes survival within macrophages. Science 304, 1800–1804 10.1126/science.1099384 [DOI] [PubMed] [Google Scholar]
  78. Wang M., Windgassen D., Papoutsakis E. T. (2008). A global transcriptional view of apoptosis in human T-cell activation. BMC Med. Genomics 1:53 10.1186/1755-8794-1-53 [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Wickstrum J. R., Bokhari S. M., Fischer J. L., Pinson D. M., Yeh H. W., Horvat R. T., et al. (2009). Francisella tularensis induces extensive caspase-3 activation and apoptotic cell death in the tissues of infected mice. Infect. Immun. 77, 4827–4836 10.1128/IAI.00246-09 [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Willingham S. B., Bergstralh D. T., O'Connor W., Morrison A. C., Taxman D. J., Duncan J. A., et al. (2007). Microbial pathogen-induced necrotic cell death mediated by the inflammasome components CIAS1/cryopyrin/NLRP3 and ASC. Cell Host Microbe 2, 147–159 10.1016/j.chom.2007.07.009 [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Yamada T., Goto M., Punj V., Zaborina O., Kimbara K., Das Gupta T. K., et al. (2002). The bacterial redox protein azurin induces apoptosis in J774 macrophages through complex formation and stabilization of the tumor suppressor protein p53. Infect. Immun. 70, 7054–7062 10.1128/IAI.70.12.7054-7062.2002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Zhang G., Ghosh S. (2001). Toll-like receptor-mediated NF-kappaB activation: a phylogenetically conserved paradigm in innate immunity. J. Clin. Invest. 107, 13–19 10.1172/JCI11837 [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Frontiers in Cellular and Infection Microbiology are provided here courtesy of Frontiers Media SA

RESOURCES