Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Transl Res. Author manuscript; available in PMC 2019 Dec 1.
Published in final edited form as:
PMCID: PMC6218307
NIHMSID: NIHMS1503027
PMID: 30165038

Mitochondria in innate immune signaling

Abstract

Mitochondria are functionally versatile organelles. In addition to their conventional role of meeting the cell’s energy requirements, mitochondria also actively regulate innate immune responses against infectious and sterile insults. Components of mitochondria, when released or exposed in response to dysfunction or damage, can be directly recognized by receptors of the innate immune system and trigger an immune response. In addition, despite initiation that may be independent from mitochondria, numerous innate immune responses are still subject to mitochondrial regulation as discrete steps of their signaling cascades occur on mitochondria or require mitochondrial components. Finally, mitochondrial metabolites and the metabolic state of the mitochondria within an innate immune cell modulate the precise immune response and shape the direction and character of that cell’s response to stimuli. Together, these pathways result in a nuanced and very specific regulation of innate immune responses by mitochondria

Introduction:

Theories that mitochondria evolved from an independent prokaryotic organism to a symbiont residing within the cytosol of the eukaryotic cell suggest that this affiliation was to mutual benefit, with the mitochondrion generating energy for the cell and the cell providing reagents and security for the mitochondrion (1,2). This theory of a bacterial origin for mitochondria fits nicely with findings that the unique components of mitochondria, when exposed, reveal their prokaryotic history and are recognized as foreign by innate immune receptors triggering an inflammatory response. Intriguingly, more recent studies suggest that the relevance of mitochondria to the innate immune response extends beyond their identification as invading bacteria and instead profoundly impact many separate aspects of innate immune responses.

Mitochondria are dynamic organelles with inner and outer membranes and an internal negatively charged matrix. Their most described function is to provide energy for the cell as the site of oxidative phosphorylation, generating 32 molecules of ATP per molecule of glucose. The vast majority of mitochondrial proteins are encoded by the nuclear DNA, transcribed and translated by the eukaryotic machinery, and then transported to their functional sites in the mitochondrion (3). Mitochondria do have their own circular DNA that encodes thirteen proteins necessary for oxidative phosphorylation along with the ribosomal and transfer RNAs needed for their translation (4). While a basic view of mitochondria may be to regard them as simply a source of ATP, the paths by which ATP is made as well as the other functions and activities of mitochondria are more complex with significant impacts upon the cell and the organism. Mitochondria are neither static nor discrete structures. In response to the conditions within the cell and both the state of and the demands being placed on the mitochondrion, mitochondria undergo fusion to combine with other mitochondria or fission to separate and form new mitochondria (5, 6). The balance of fusion and fission events has relevance beyond just determining the number of mitochondria to a cell as these processes also impact calcium regulation, generation of reactive oxygen species, and impact oxidative phosphorylation (7).

Generation of ATP by mitochondria requires the negatively charged matrix that allows the passage of electrons over the electron transport chain, which consists of specialized complexes arranged on the inner mitochondrial membrane. Disruption of the negative charge of the matrix occurs in response to a number of stressors including antioxidants, oxidative phosphorylation substrates, and membrane uncoupling agents. This loss of negative potential results in a failure of ATP production and the generation and release of reactive oxygen species (ROS) that have the potential to cause widespread damage. Further, dysfunctional mitochondria can lose membrane integrity, allowing previously sequestered mitochondrial components to leak into the cytosol or out of damaged cells to the circulation (810). To limit the negative effects of ROS, damaged and dysfunctional mitochondria are removed through a process known as mitophagy (11).

The innate immune response has a critical role in the detection and correction of both infectious and sterile insults. The response begins with the recognition of the insult, commonly through germline encoded receptors termed pattern recognition receptors (PRRs). These receptors bind to conserved features of microbes that identify them as foreign, or to endogenous molecules with specific modifications or in locations that reveal tissue or cellular injury (12, 13). This recognition of the insult, with the PRR bound by its specific activating ligand, triggers the immune response. While the precise characteristic of that immune response reflects both the type of innate immune cell being activated and the specific receptor and ligand, in general these innate responses are initiated quickly and also quickly escalate and alert additional cells and tissues of the disorder. These early signals of the initial innate immune response have effects throughout the organism, driving recruitment of additional innate immune cells to the site, alerting and activating the more specific adaptive immune response, and triggering the production of molecules needed for inflammation and repair by numerous tissues (1418).

Both structural and functional aspects of the mitochondria can impact the innate immune response. There are two broad categories by which this occurs: first, by directly activating the immune response and second, by modulating a response. Direct activation commonly reflects mitochondrial damage or pathology while modulation can occur as a byproduct of normal mitochondrial functions and processes. In this review, we will discuss the current literature that defines the interactions between mitochondria and the innate immune response.

Mitochondria derived alarmins of innate immunity:

Pathogen associated molecular patterns, or PAMPs, are conserved features of invading organisms that serve to identify these organisms as foreign, while damage associated molecular patterns, or DAMPs, are endogenous molecules released or modified by sterile insults. Both DAMPs and PAMPs are specifically recognized as alarmins by discrete receptors of the innate immune system and trigger the appropriate immune response (12, 13). Their release by mitochondria is in response to cell stress and loss of homeostasis, similar to more typical DAMPs, but the recognition of these alarmins by PRR depends on their structural similarity to the PAMPs of invading microbes. In this review, we will refer to these alarmins as DAMPs to emphasize their pivotal function as markers of cell stress. The DAMPs that are released by mitochondria include the structural phospholipid cardiolipin, n- formyl peptides (n-fp), reactive oxygen species (ROS) and mitochondrial DNA (mtDNA). While the exact mechanism by which these mitochondrial alarmins are released has not been elucidated completely, studies have suggested loss of mitochondrial membrane integrity results in the escape of mitochondrial components to the cytosol. There are discrete pathways by which the inner and outer mitochondrial membranes are disrupted, through sustained opening of a mitochondrial permeability transition (MPT) pore and mitochondrial outer membrane permeabilization (MOMP), respectively (19, 20). Despite being recognized very early, the details of the structure of the MPT pore as well as the triggers that lead to its formation and opening are not yet clear (21). In general, the MPT pore opens and remains open after insults to the mitochondrion that are associated with disruption of calcium levels or oxidative stress, consistent with the insults believed to be associated with release of mitochondrial DAMPs (22). Similar gaps exist in our knowledge of the structure and regulation of MOMP. MOMP has been studied extensively as a trigger for apoptosis, when BCL-2 family members induce the formation of pores in the outer mitochondrial membrane in response to either death receptors or various cellular or mitochondrial stresses (23). However, MOMP has been reported to occur in the absence of cell death, providing a potential pathway by which MOMP could be associated with the release of mitochondrial contents leading to an innate immune response (24, 25). A recent breakthrough shows the large pore induced in the outer mitochondrial membrane during intrinsic apoptosis is associated with a herniation of the inner mitochondrial membrane. In the cases where this herniated membrane ruptures, mitochondrial DAMPs were released to the cytosol where they could be sensed by various PRRs (24). While the exact pathways and pores or channels that are associated with the release of mitochondrial alarmins are only beginning to be defined, it is clear that after escaping the mitochondrion these molecules are recognized by separate receptors and trigger discrete inflammatory pathways that culminate in the restoration of normal cellular function.

n-FP:

In a mann r to the initiation of protein translation in prokaryotes, mitochondrial initiation of protein translation requires N-formylated methionine (fMet), as mitochondrial translational initiation factor 2 can utilize only the formylated form of methionine, while unformylated methionine is used specifically for protein elongation (26). This unique characteristic of bacterial and mitochondrial proteins was proposed as a potential immune target well before the receptors and signaling pathways had been defined (27, 28). In the absence of tissue stress or injury these bacteria-like n-formyl peptides (n-FP) are sequestered within the mitochondria. However, during traumatic injury and cell death associated with infection, n-FP are released and bind specific receptors, formyl peptide receptors (FPRs) that in turn recruit immune cells and trigger an extensive inflammatory response as discussed below (2931).

Cardiolipin:

Cardiolipin is a unique phospholipid that was first identified in animal heart tissue and thus the family is known as cardiolipins. While the structure of cardiolipin varies, in general cardiolipin contains two phosphatidyl groups linked through a glycerol moiety. Cardiolipins are found in many prokaryotic membranes but in eukaryotic cells are limited to the mitochondrial membranes, primarily the inner mitochondrial membrane during normal mitochondrial function (32, 33). Cardiolipin contributes to the structural composition and integrity of the mitochondria membrane and constitutes about 20% of the phospholipid content of the mitochondrial inner membrane (3436). In addition to its structural role, cardiolipin has numerous non-redundant roles in the mitochondria that reflect its remarkable ability to interact via non-covalent bonds with a wide variety of unrelated molecules. This unique binding capability allows cardiolipin to serve as a discretely controlled regulator of numerous otherwise separate mitochondrial pathways, including mitochondrial dynamics (fission and fusion), the import of cargo from cytosol, metabolic functions, innate immune responses, ROS generation, and apoptotic signaling (37, 38). These functions of cardiolipin occur in its normal environment within mitochondria, while other functions of cardiolipin, in particular its ability to drive innate immune responses, occur after its release from or externalization on the surface of the mitochondrion during conditions of mitochondrial dysfunction, stress, or damage (39). It is in this context that cardiolipin may play a role in activation of the NLRP3 inflammasome (40, 41). In a separate pathway, following ischemia/reperfusion injury, cells release oxidized phospholipids, including cardiolipin, that serve as DAMPs and trigger multiple types of innate immune responses (42). In contrast to these pro-inflammatory roles for exposed or released cardiolipin, externalized cardiolipin has been reported to downregulate innate cytokine responses by upregulating mitophagy pathways (43, 44).

mtDNA:

Mitochondrial DNA is a circular double strand of approximately 17Kbp that contains 13 mRNAs that encode the unique proteins of oxidative phosphorylation as well as related ribosomal RNAs and tRNAs. Similar to the mitochondrial use of bacterial-like machinery for protein translation discussed above, mitochondria DNA has characteristics consistent with prokaryotic nucleic acid. Mitochondrial DNA is a small molecule with methylation patterns discrete from nuclear DNA and is present at hundreds of copies per cell. Under normal conditions, mtDNA is contained within the mitochondrial matrix where mitochondria have unique mechanisms to repair DNA damage as well as to preserve genetic integrity through the selective amplification of intact copies of mtDNA (45, 46). These repair pathways may also modify the mtDNA in such a way to mark it as non-self to innate immune sensors. During cell death or mitochondrial damage the mtDNA can be released into the cytosol or the circulation where it can be sensed by a number of innate immune receptors and trigger inflammatory responses as detailed below. This release of mtDNA associated with inflammation is supported by the finding of elevated levels of circulating mtDNA in patients suffering from trauma, rheumatoid arthritis, and femur fracture as well as in animal models of trauma and shock (4751).

mROS:

Mitochondrial reactive oxygen species (mROS) generation occurs via the electron transport chain (ETC) in response to altered substrate availability, hypoxia, or other abnormal mitochondrial or cellular conditions (52, 53). Electrons that leak from the matrix ETC chain at complex I-III react with oxygen to result in superoxide radicals. While it has long been held that these reactive molecules augment the immune response by attacking intracellular pathogens (54, 55), it is now established that their relevance to the immune response is significantly more extensive, as mROS can modify innate cellular responses both directly and indirectly. mROS has been shown to directly modify cellular functions through their interaction with the pivotal transcription factors HIF-1α (hypoxia-inducible factor-1α) and NFkB (nuclear factor kB) (56, 57). Their indirect effects are through their ability to interact with and modify other molecules, including mtDNA, that results in heightened sensing by innate immune receptors.

Just as the production of mROS is tightly regulated, so is the removal of mROS. Scavenging enzymes of the superoxide dismutase family convert superoxide to hydrogen peroxide, which is then further reduced to water by catalase, glutathione peroxidases, and peroxidredoxins (5860). Uncoupling proteins downregulate mROS production by decreasing the mitochondrial membrane potential (61). Finally, removal of mROS-producing damaged mitochondria by mitophagy reduces mROS and downregulates innate immune responses, while blockade of mitophagy is associated with an increase in inflammation (62).

Additional mitochondrial alarmins include ATP that is expelled extracellularly by apoptotic or necrotic cells and sensed through P2X7 purinergic receptors to trigger innate immune responses including the NLRP3 inflammasome (63, 64). Mitochondrial transcription factor A (TFAM) is structurally homologous to the nuclear alarmin molecule high mobility group box protein 1 (HMGB1), and is released from damaged mitochondria (65). TFAM have been reported to activate dendritic cell subsets, amplify TLR9 signaling, and trigger inflammatory cytokine release (66, 67).

Mitochondria and innate immune responses:

The many DAMPs of mitochondria can be released into the cytosol or extracellularly following infection, injury, or loss of cellular or mitochondrial homeostasis. These DAMPs can then be sensed by the numerous PRRs, which are germline encoded receptors that recognize the unique signatures of PAMPs and DAMPs and upon activation trigger the innate immune response (68). Based on their structures, locations, and functional specificities, PRRs are separated into discrete families that include the membrane bound Toll-like receptors (TLRs) and C-type lectin receptors (CLRs) and the cytosolic NOD (nucleotide-binding oligomerization domain)-like receptors (NLRs) and RIG (retinoic acid-inducible gene)-l-like receptors (RLRs). Upon activation, these receptors trigger the release of cytokines and chemokines that recruit and activate other immune cells as well as regulate organism-wide responses to the specific insult that lead to (69).

Mitochondria and TLR activation and signaling:

The TLRs are a PRR, first identified in Drosophila, that are activated by ligand binding to their carboxy-terminal leucine rich repeat (LRR). Ten TLRs have been identified in humans, with TLRs 1,2, 4, 5, 6, and 10 found on the cell surface and TRLs 3, 7, 8, and 9 spanning the endosomal membrane. Activation of TLRs results in signaling through the p38 and MAPK pathways with the resultant activation and nuclear localization of NFkB triggering the expression of pro-inflammatory genes. While TLRs are activated by pathogens and non-mitochondrial DAMPs, a number have been shown to be activated by mitochondrial components. For example, mtDNA, with its prokaryote like structure, can be recognized by TLR9 (Figure 1) (47). However, mitochondria are implicated in TLR signaling beyond acting as direct activators. Activated TLRs can signal through TRAF6 (tumor necrosis factor receptor-associated factor 6), which translocates to the mitochondrion and ubiquitinates the mitochondrial complex l-associated protein ECSIT (evolutionarily conserved signaling intermediate in Toll pathways). This causes the mitochondrion to both move to the phagosome and enhance mROS production, resulting in direct antimicrobial killing (55, 70). Additional studies have shown relevance for mitochondria in the signaling pathways downstream of TLR activation. Early studies linked LPS- (lipopolysaccharide) induced ROS to TLR4 signaling and dsRNA-induced ROS to TLR3 signaling, with loss of ROS associated with downregulation of NFkB activation (71). Subsequent studies linked ROS to p38 signaling activation, with the ROS generated by TLR4 activation necessary for TRAF6 and ASK1 (apoptosis signal-regulating kinase 1) to trigger p38, but not NFҡB signals (72). More recently, NFҡB signaling following TLR4 activation has been confirmed to be dependent upon mitochondria, although not specifically ROS, with TLR4 activation drives the interaction of mitochondrial ECSIT with both TRAF6 and TAK1 (transforming growth factor- β-activated kinase 1), and that this interaction is required for downstream NFkB signaling (73). TLR7 signaling by viral nucleic acid is also modulated by mitochondrial components as the mitochondrial outer membrane ubiquitin ligase MARCH5 (membrane-associated ring finger (C3HC4) 5) has been shown to interact with and ubiquitinate TANK (TRAF family member- associated NFҡB activator) in response to TLR7 stimuli. This interaction prevents TANK from inhibiting TRAF6 and leads to enhanced TLR7-induced responses (74).

An external file that holds a picture, illustration, etc.
Object name is nihms-1503027-f0001.jpg
Mitochondrial regulation of innate immune responses.

Mitochondrial DAMPs activate a number of innate immune pathways. Mitochondrial DNA that escapes to the cytosol from damaged mitochondria is recognized by cGAS and signals through cGAMP and STING to activate inflammatory gene transcription. Mitochondrial DNA leaked from a cell can be phagocytosed and bind endosomal TLR9, triggering MyD88-dependent signaling to interferons and pro-inflammatory cytokines. These inflammatory cascades also serve to prime the NLRP3 inflammasome and upregulate pro-lL-1 β. The NLRP3 inflammasome is then activated by oxidized mtDNA from dysfunctional mitochondria. Cardiolipin, which moves to the outer mitochondrial membrane in response to mitochondrial dysfunction, tethers the NLRP3 inflammasome to the mitochondrion and can also trigger its activation. ATP that leaks from a damaged cell can bind to the P2X7 receptor on adjacent cells, triggering NLRP3 inflammasome activation within those nearby cells. Similarly, formyl peptides are also released by damaged cells and are recognized by FPRs on neutrophils and result in neutrophil activation including chemotaxis and the respiratory burst. Modulation of innate immune signaling pathways also depend upon mitochondria. Mitochondria- independent activation of TLRs results in signals through the mitochondrial protein ECSIT, generating mROS and enhancing inflammatory gene output. Activation of the RLRs MDA5 and RIG-I by viral RNA is initiated in the cytosol but signaling depends upon the mitochondria, as both MDA5 and RIG-I must bind MAVS on the outer mitochondrial membrane to activate their downstream signaling pathways. This results in upregulation of interferons and other inflammatory genes. Further, mROS can enhance RLR signaling by upregulating MAVS expression on the outer membrane.

The interactions between TLR signals and mitochondria is bidirectional, as TLR activation impacts mitochondrial characteristics as well. Mitochondrial gene expression is upregulated downstream of both TLR2 and TLR4 through activity of members of the PPAR-γ coactivator family (75, 76). Separate studies have shown upregulation of mtDNA downstream of TLR2 and TLR4 activation through functions of the transcription factors NFҡB, CREB (cAMP response element binding protein), NRF2 (nuclear factor erythroid 2-related factor 2), IRF (interferon regulatory factor)3, and IRF7 (77, 78). Thus, mitochondria can both directly activate TLR pathways and also modulate their signaling to regulate the resulting innate immune response.

Mitochondria and the NLR family:

Unlike the membrane-associated TLRs, the NLR family of pattern recognition receptors are localized in the cytoplasm of the cell where they are activated by PAMPs and/or DAMPs. Most members of the NLR family have a tripartite domain structure: carboxy-terminal LRR, a central nucleotide binding domain (NBD), and a variable amino- terminal domain that is involved in protein-protein interactions. NLRs are divided into subfamilies by the class of amino-terminal domain: an acidic transactivation domain (AD), a baculoviral inhibitory repeat (BIR)-like domain, a caspase recruitment domain (CARD), a pyrin domain, or a domain of unknown function. A unique feature of some NLRs is, upon activation, their formation of a multi-protein complex termed an “inflammasome”. The core structure of an inflammasome is the NLR as the sensory component, bound at the amino-terminal domain to the adaptor ASC (apoptosis-associated speck-like protein containing a CARD), that forms a bridge to the CARD domain of pro-caspase-1, which is the inactive pro-enzyme of the effector cysteine protease caspase-1. Inflammasome assembly triggers auto-catalysis of pro-caspase-1 to active caspase-1 that in turn cleaves inactive pro-IL (interleukin)- 1β and pro-IL-18 to their active, secreted forms that have broad pro-inflammatory effects. Active caspase-1 also cleaves gasdermin D that can introduce pores in the cell membrane through which IL-1 β and IL-18 are secreted and that can also induce inflammasome-associated pyroptotic cell death (79, 80). A number of NLRs have been suggested to form inflammasomes that culminate in caspase-1 activation, while some have been reported to have inflammasome-independent functions. Thus far, a defined role for mitochondria in NLR activation and function has only been confirmed for NLRP3 and NLRX1 as detailed below.

NLRP3:

NLRP3 was first described when its mutation was found to be causative to a group of autoinflammatory disorders, now collectively known as CAPS (cryopyrin-associated periodic syndromes) (8183). Subsequently, activation of the NLRP3 inflammasome has been linked to a wide array of infectious and sterile inflammatory disorders, including but not limited to bacterial, viral, and fungal infections, metabolic syndrome, ischemia-reperfusion injury, atherosclerosis, Alzheimer’s disease, and gout (39). NLRP3 is expressed in macrophages, monocytes, dendritic cells, neutrophils, as well as in numerous non-hematopoietic cells (84). Although NLRP3 is the most studied member of the NLR family, significant gaps remain in our understanding of its activation. NLRP3 activation occurs in response to two discrete steps, termed priming and activation. While it has been established both steps must occur, with priming preceding activation, the precise events that occur during each step to allow inflammasome activation to proceed are defined incompletely. In priming, an inflammatory stimulus such as TLR activation or binding of a cytokine to a receptor signal through the adaptor molecule MyD88 (myeloid differentiation primary response 88) or TRIF (TIR-domain-containing adapter- interferon β) with subsequent activation of NFkB (85, 86). This induces the upregulation of expression of both NLRP3 and pro-1 L-1 β, although increased protein levels are neither sufficient nor required for priming to occur (85, 87, 88). Additional steps implicated in priming include multiple post- translational modifications to inflammasome components, including the de-ubiquitination and phosphorylation of NLRP3 and the ubiquitination and phosphorylation of ASC (87, 8995). The activation step is induced by a wide array of structurally diverse molecules, including endogenous and exogenous crystals that require phagocytosis, ATP via the cell surface P2X7 receptor, and pathogen-derived toxins. As none of these activators has been shown to directly associate with NLRP3, studies have focused on the identification of a final common pathway downstream of these divergent stimuli to determine the mechanism by which the NLRP3 inflammasome is activated. These studies have led directly to the mitochondrion, showing that activated NLRP3 inflammasomes co-localize with mitochondria (96). Recent work by our group has shown the use of mitochondria as a scaffold for inflammasome assembly is initiated at priming, as both NLRP3 and pro-caspase-1 move to the mitochondrion at priming, with the movement of the adaptor molecule ASC following in response to the activating stimulus (97). While additional studies have confirmed the co-localization of the NLRP3 inflammasome with mitochondria, the mechanism by which it occurs is less clear, with studies separately linking it to MAVS (mitochondrial antiviral signaling protein), cardiolipin, or c-FLIP (c-FLICE-like inhibitory protein) (39, 40, 98100). The role of MAVS is intriguing as instead of serving as a common mediator for all NLRP3 inflammasome activators, it appears to be necessary for only a subset of NLRP3 activators. The initial report describing a role for MAVS in NLRP3 activation showed that only the non-crystalline subset of NLRP3 activators required MAVS, as activation of NLRP3 by crystalline activators did not require or involve MAVS (98). A subsequent study confirmed a role for MAVS in NLRP3 activation by Sendai virus but did not explore further the type of agonist driving the response (99). Additional studies will be needed to determine what function MAVS provides specifically to the non-crystalline activators of NRLP3 and to determine if a separate agent serves a parallel role for the crystalline activators of NLRP3.

In the same study that first showed the co-localization of NLRP3 inflammasomes with mitochondria the authors also showed that the activation of NLRP3 was associated with both mitochondrial damage and the release of mROS, and that these were required for NLRP3 activation (96). Subsequent studies have confirmed mROS is induced by most but not all activators of NLRP3, although the induction of mitochondrial dysfunction has still been shown to be required for some of these ROS-independent activators (40, 101106). This is also consistent with studies showing inhibition of mitophagy, the removal of damaged or dysfunctional mitochondria, augments activation of NLPR3 (62, 96). While the source of the mitochondrial damage has not been determined, one possible mechanism is that the elevated cytosolic calcium required during NLRP3 inflammasome activation is taken up by mitochondria, overloading the mitochondria causing loss of negative potential in the mitochondrial matrix and resulting in the mitochondrial dysfunction that drives NLRP3 activation.

The relevance of mitochondria to NLRP3 activation extends beyond this role for mROS and mitochondrial dysfunction as well as its function as a scaffold upon which the NLRP3 inflammasome assembles. In contrast to the various extra-cellular activators the result in NLRP3 activation, specific components of mitochondria have been suggested to directly activate NLRP3. An early study by Dr. Choi’s group showed mtDNA was released into the cytosol during NLRP3 activation downstream of mROS release (101). This finding was built on by subsequent work by Dr. Arditi’s group that showed mtDNA undergoes oxidation during the mitochondrial dysfunction associated with NLRP3 activation and that this oxidized mtDNA bound to and activated the NLRP3 inflammasome, suggesting mtDNA may be the ligand that directly drives NLRP3 activation (102). Studies from our group showed a different mitochondrial component, cardiolipin, both tethers NLRP3 to the mitochondria and triggers the activation of NLRP3 (40). Future studies will be necessary to determine the relative roles of discrete mitochondrial factors in the subcellular localization and activation of NLRP3.

NLRX1:

That the NLR family member NLRX1 interacts with mitochondria was initially suggested based on its novel amino-terminal domain. Rather than the pyrin or CARD domains found in other NLR family members, the NLRX1 protein begins with a putative mitochondrial targeting sequence. NLRX1 was confirmed experimentally to localize to the mitochondria but its exact mitochondrial location and its function have been less straight forward to determine (107, 108). Initially, NLRX1 was shown to be a negative regulator of type I interferon production by binding to MAVS, preventing the interaction of MAVS and RIG-I (retinoic acid inducible gene-l), and thereby blocking the downstream activation of IRF3 and NFkB (107). This negative regulatory role was supported by two subsequent studies, the first of which showed NLRX1 blocked the inflammatory response to DNA viruses by binding to STING (stimulator of interferon gene) and preventing downstream cGAS (cyclic GAMP synthase) signaling (109). The second confirmatory study reported the loss of NLRX1 was associated with enhanced NFkB signals, consistent with NLRX1 being a negative regulator of inflammation (110). In contrast, a separate report using overexpressed NLRX1 showed that rather than inhibiting interferon signaling, NLRX1 increased NFkB signaling downstream of Shigella flexneri infection (108). In support of NLRX1 not downregulating the inflammatory response, NLRX1-deficient fibroblasts and macrophages had no defect in interferon responses to Sendai virus and NLRPXI-deficient mice had no identifiable defects in early signals to influenza A virus in vivo (111). A second study supported this independence of anti-viral signaling from NLRX1 (112). A recent study may have found the explanation for these seemingly conflicting functions, as NLRX1 was shown to simultaneously promote the upregulation of IRF1 but limit the formation of IRF3 dimers, consistent with a nuanced regulatory function for NLRX1 in innate immune responses with potentially both pro- and anti-inflammatory aspects (113). However, controversy relating to NLRX1 persists: while initial studies described the localization of NLRX1 on the outer mitochondrial membrane, NLRX1 has more recently been described as undergoing translocation into the mitochondrial matrix in a pathway dependent upon the negative potential within the mitochondrial matrix (107, 114). With this matrix localization NLRX1 was found to interact with a protein of the respiratory chain, potentially explaining the modulation of ROS described with NLRX1 overexpression, but was subsequently and separately shown instead to interact with mitochondrial TUFM (Tu translation elongation factor) to regulate viral-induced autophagy (108, 114116). In addition to these conflicting reports as to the location of NLRX1 and its impact on interferon signals, the relevance of NLRX1 in regulating ROS is similarly unclear. While several studies have shown NLRX1 driving increased ROS in response to inflammatory stimuli, a conflicting report described the loss of NLRX1 was associated with an increase in oxygen consumption, consistent with a downregulatory rather than upregulatory role for NLRX1 (108, 111, 115, 117, 118). One possible explanation for the divergent findings of NLRX1 in these studies is that NLRX1 may impact uncoupling of oxidative phosphorylation, resulting in a disconnect between oxygen consumption, ATP generation, and ROS production (119). Additional studies will be needed to confirm if a possible role in uncoupling explains these seemingly discrepant results, to expand on how NLRX1 balances up- and down-regulation of interferons, and also to determine how NLRX1 modulates outer mitochondrial membrane components like MAVS if its location is confirmed within the mitochondrial matrix.

Mitochondria and RLRs:

As viral genomes can undergo amplification in the cytoplasm of host cell, they are often inaccessible to detection by TLRs. Intact type I interferon responses to viral infections in cells lacking the sole TLR3 adaptor TRIF suggested a TLR-independent receptor pathway may exist that also generated these potent anti-viral cytokines (120). The quest to identify these sensors of viral infections lead to the identification of the RNA helicase retinoic acid inducible protein I (RIG-I). RIG-I, along with the other RLR (RIG-l-like receptor) family members, melanoma differentiation associated gene 5 (MDA5), and laboratory of genetics and physiology 2 (LGP2) are cytosolic proteins that sense viral RNA (121). The RLRs share a conserved domain structure with a pair of amino-terminal CARD domains (absent in LGP2), a central RNA helicase domain, and a carboxy-terminal regulatory domain. RIG-I and MDA5 bind cytosolic viral RNA and, through the adaptor MAVS, trigger the release of type I interferons (122125). RIG-I recognizes the unique 5’-phosphorylated blunt ends of viral genomic RNA while MDA5 binds long dsRNA (126128). The function of LGP2 seems to be to augment these RIG-I or MDA5-triggered responses (129133).

As noted, RLR signaling require MAVS, a 540 amino acid protein with three domains: an amino-terminal CARD, a proline rich central domain, and a carboxy-terminal transmembrane portion. The transmembrane domain anchors MAVS primarily to the mitochondrial outer membrane, but MAVS is also inserted within, and can signal from, the mitochondrial associated membranes, the endoplasmic reticulum, and peroxisomes (134, 135). RIG-I or MDA5, activated by their nucleic acid ligands, bind to and activate MAVS via CARD-CARD interactions, triggering polymerization of MAVS into prion-like structures that are required for it to signal downstream (136, 137). MAVS then complexes the cytoplasmic kinases TBK1 (Tank binding kinase 1) and ΙΚΚε (IKB kinase-ε) resulting in activation of the transcription factors IRF3, IRF7 and NFkB and upregulating the transcription of type I interferons, interferon inducible genes and proinflammatory cytokines (136, 138). This activation pathway is regulated through extensive post-translational modifications, including ubiquitination and phosphorylation at each step of activation (139143). The mitochondrion itself also significantly impacts and regulates the activation of MAVS. MFN2 (mitofusin 2), a mitochondrial GTPase that, along with MFN1 and OPA1, regulates mitochondrial fusion, was shown to inhibit RLR signaling by binding to MAVS and preventing CARD dimerization (144). Despite significant homology between MFN2 and MFN1, in contrast to MFN2, MFN1 was found to enhance MAVS signaling (145, 146). Additional proteins that have been suggested to serve as negative regulators of MAVS are NLRX1, gC1qR (globular head domain of complement component 1q receptor), and PLK1 (Polo-like kinase 1) (107, 147, 148). The negative membrane potential of the mitochondria has also been linked to MAVS signaling as dissipation of that potential abrogated RLR signaling (149). ROS have been shown in two studies to regulate RLR signaling, first that the increase in ROS resulting from the blockade of autophagy upregulates RLR signaling and subsequently that MAVS protein expression, and thereby MAVS signaling, is dependent upon ROS (150, 151). Despite these important findings, the mechanistic details as to the activation, regulation, and signaling of these pathways remain to be determined.

Mitochondria and cGAS-STING pathways:

Although DNA was hypothesized to be immunogenic over a century ago, the sensors and immunologic pathways activated by cytosolic DNA remain only partially characterized (152). STING (stimulator of interferon genes) was identified ten years ago as pivotal in the release of type I interferons (105, 153, 154). This activation of STING was found to be in response to cytosolic DNA, and while more commonly considered to be of nuclear or microbial sources, this activation is also induced by mitochondrial DNA that can be released during apoptosis or other conditions of mitochondrial damage and dysfunction (155161). STING is not the sensor of cytoplasmic DNA but rather an adaptor that links the activated receptor to downstream inflammatory gene upregulation. Cytosolic DNA binds to cGAS (cyclic GMP (guanosine monophosphate) synthase) and induces a change in cGAS conformation that activates cGAS to catalyze GTP and ATP to the second messenger cGAMP (cyclic GMP-AMP) that then binds to and activates STING (162167). This binding induces a conformational change in STING, triggering its translocation from the ER to the Golgi apparatus (153, 166, 168, 169). This movement to the Golgi activates STING, in turn activating TBK1 and IKK which phosphorylate and activate the transcription factors IRF3 and NFҡB. These transcription factors move to the nucleus and drive type I interferon and proinflammatory cytokine production (170, 171). Thus, any mitochondrial event that results in the leaking of mtDNA can stimulate the cGAS-STING pathway, resulting in a potent innate inflammatory response.

Mitochondria and FPR signaling:

Formyl peptide receptors (FPR) are a family of G protein coupled receptors expressed as transmembrane proteins on the surface of many hematopoietic as well as non-hematopoietic cells (172). They were first described on neutrophils and have been studied most extensively in regulating neutrophil migration and function (172, 173). These receptors are activated by n-FP, peptides produced by microbes or mitochondria that contain an amino-terminal formylated methionine (fMet) (27, 28). There are three family members in humans, FPR1, FPR2, and FPR3, with FPR1 the best characterized. FPR1 and FPR2 share significant sequence homology, particularly for the cytoplasmic signaling domains (174, 175). Both FPR1 and FPR2 bind n-FP, although with different binding specificity as FPR1 and FPR2 preferentially bind shorter and longer peptides, respectively (176, 177). While the other innate immune pathways induced by mitochondrial DAMPs discussed in this review predominantly result in inflammatory cytokine release, the interaction of n-FP with FPRs primarily triggers migration and neutrophil activation (178, 179). The relevance of FPRs to disease was confirmed by studies showing circulating mitochondrial DAMPs in patients with trauma-associated systemic inflammatory response syndrome with mechanistic animal studies confirming neutrophil migration to mitochondrial DAMPs required FPR1 (49, 180, 181). The intracellular signaling downstream of FPR1 activation starts with G-protein dissociation of the α and βγ subunits. Activation of phospholipases drives the generation of diacylglycerol and inositol triphosphate, triggering the release of calcium from the ER to the cytosol that culminates in ROS production (182). Parallel activation of kinase pathways results in signals necessary for cytoskeletal organization and contribute to ROS production and neutrophil activation (183). FPR1 has been shown to function in other innate immune cells, stimulating inflammatory cytokine release from monocytes and regulating the maturation and migration of dendritic cells (184, 185).

Thus, numerous components of the mitochondria can be released and trigger an array of inflammatory receptors and signaling pathways. These are summarized in Table 1.

Table 1.

Mitochondrial DAMPSensorResult
n-formyl peptidesFPR1, FPR2Chemotaxis, Oxidative burst
CardiolipinNLRP3Inflammasome activation, pro-inflammatory cytokines
mtDNATLR9, NLRP3, cGAS-STINGType I interferons, pro-inflammatory cytokines, inflammasome activation
ROSNLRP3Inflammasome activation, pro-inflammatory cytokines
ATPNLRP3Inflammasome activation, pro-inflammatory cytokines

Mitochondrial metabolism and innate immunity:

The focus of this section is the specific regulation of innate immune pathways by mitochondrial metabolism. For a broader discussion on, and insight into, immune regulation by metabolism the reader is referred to a recent in-depth review (186).

Mitochondria use oxidative phosphorylation via the electron transport chain (ETC) and the tricarboxylic acid (TCA) cycle to generate ATP for cellular functions. The electron transport chain consists of five multiprotein complexes on the inner mitochondrial membrane. Electrons are donated from NADH to complex I or FADH2 to complex II (also known as succinate dehydrogenase) and transferred through coenzyme Q, complex III, cytochrome C, complex IV, to complex V (also known as ATP synthase) which generates ATP from ADP. The TCA cycle, responsible for the majority of ATP production, progresses in concert, regenerating NADH and FADH2 to maintain electron transport. Under normal conditions a small amount of electrons escape the ETC to combine with oxygen and form oxygen radicals, collectively known as reactive oxygen species (ROS) (Figure 2A) (187189).

An external file that holds a picture, illustration, etc.
Object name is nihms-1503027-f0002.jpg
Mitochondrial metabolism in immune cell polarization and innate immune responses.

A. In resting macrophages, glucose and fatty acids are broken down to acetyl CoA that enters the TCA cycle. As the TCA cycle progresses, NADH and FADH2 are regenerated as electron donors for the ETC. Most electrons in the ETC are used to generate ATP although some leak off and combine with oxygen to create ROS. B. Activated M1 macrophages have modifications in their metabolism that drives their pro-inflammatory characteristics. (1) M1 macrophages downregulate isocitrate dehydrogenase, resulting in a block in the cycle moving forward from citrate. (2) Citrate accumulates. (3) The metabolic products of citrate are converted to itaconate by IRG1 (immune response gene 1 protein), which is markedly upregulated in M1 macrophages (4) In addition to direct antimicrobial effects, itaconate also inhibits complex II (also known as succinate dehydrogenase), preventing forward progression of the ETC. (5) Reverse transfer of electrons to complex 1 results in increased ROS generation. (6) increased ROS results in stabilization of the transcription factor HIF-1a with subsequent upregulation of inflammatory gene expression, including the potent inflammatory cytokine IL-1 β. C. The metabolic signature of activated M2 macrophages is necessary for their function. (1) M2 macrophages have enhanced fatty acid oxidation driving the TCA cycle to generate ATP. (2) M2 macrophages also require glycolysis, hydrolyzing glucose and using glutamine through the hexosamine pathway to generate uridine diphosphate (UDP)-N-acetylglucosamine (UDP- GIcNAc). This production of UDP-GIcNAc is necessary for (3) glycosolation of immune receptors and (4) the upregulation of expression of specific M2 markers.

In the TCA cycle, fatty acids and pyruvate are brought into the mitochondrion and oxidized to acetyl-CoA which enters the TCA cycle through combining with oxaloacetate to form citrate. Citrate is oxidized to continue the cycle, leading to regeneration of NADH and FADH2 for the ETC and then back to oxaloacetate to continue the cycle (187189). Perturbations to the ETC, TCA cycle, and TCA cycle metabolites have impacts upon wider immune cell function and respon.

Activated macrophanges manipulate the TCA cycle and ETC to modulate their immue function (190, 191). Inflammatory (M1), but not anti-inflammatory (M2) macrophages, have what is referred to as a “broken” TCA cycle (192). Down regulation of the TCA enzyme isocitrate dehydrogenase and upregulation of immune-responsive gene 1 protein (IRG1) blocks TCA progression and result in the conversion of accumulating citrate to itaconate (Figure 2B) (193, 194). Itaconate has direct antimicrobial functions that augment the innate immune response, but also inhibits complex II (also known as succinate dehydrogenase), preventing succinate oxidation (195197). This results in reverse electron transport back down the ETC; associated increased ROS stabilizes the transcription factor HIF-1α that enhances the inflammatory response (198, 199). Interference with the TCA cycle in M1 macrophages blocks the production of ATP, consistent with the finding that M1 macrophages rely on glycolysis for energy production while M2 macrophages generate ATP with the TCA cycle (194, 200202). Indeed, additional studies have shown that differential stimuli of innate immune cells trigger unique metabolic signatures necessary for subsequent immune function (203, 204).

Concluding remarks:

From their relatively humble beginnings as ancient bacteria, mitochondria have established themselves as regulators of sweeping aspects of mammalian function at both the cellular and organism level. This regulation extends well beyond simply providing critical bioenergetics and includes, but is not limited to, precise and nuanced control of innate immune activation and signaling. The advances in our understanding of the importance of mitochondria to the innate immune response along with technical advances in our ability to study the mitochondria are likely to reveal additional novel pathways through which these seemingly discrete systems are in fact intertwined.

Acknowledgements:

This work was supported by NIH grant AI104706 to Suzanne L. Cassel. We thank Fayyaz Sutterwala for discussing, reviewing, and editing the manuscript. The authors have read the journal’s policy on disclosure of potential conflicts of interest and have no conflicts of interest to disclose. The authors have read the journal’s authorship agreement and the manuscript has been reviewed by and approved by the authors.

Abbrevations:

PAMPPathogen Associated Molecular Pattern
DAMPDamage Associated Molecular Pattern
PRRsPathogen Recognition Receptors
ROSReactive Oxygen Species
TLRToll Like Receptor
RLRRetinoic acid Inducible Gene like Receptor
NODnucleotide-binding oligomerization domain
NLRNOD like receptor
ATPAdenosine Tri-Phosphate
MFN1/2Mitofusin
RIG-IRetinoic acid Inducible Gene I
MAVSMitochondrial Anti-Viral Signaling
ASCApoptosis associated Speck like protein containing CARD
ETCElectron Transport Chain
ESCITEvolutionarily Conserved Signaling Intermediate in the Toll pathway
NFҡBNuclear factor κ B
IRFInterferon Regulatory Factor
IFNInterferon
JNKcJUN NH2- terminal kinase
TANKTRAF family member-associated NFkB activator
TBK1TANK Binding Kinase 1
cGAScyclic GMP-AMP Synthase
STINGStimulator of Interferon Gene
CARDCaspase Activation and Recruitment Domain
ILInterleukin
CLCardiolipin
mROSmitochondrial ROS
mtDNAmitochondrial DNA
TCATri-carboxylic acid
NADHNicotinamide adenine dinucleotide
TFAMMitochondrial transcription factor A
HMGB1high mobility group box protein 1
fMetN-formylated methionine
n-fpn-formyl peptides
FPRformyl peptide receptor
CLRC-type lectin receptor
LRRleucine rich repeat
MAPKmitogen-activated protein kinase
TRAF6tumor necrosis factor receptor-associated factor 6
LPSlipopolysaccharide
ASK1apoptosis signal-regulating kinase 1
TAK1transforming growth factor-β-activated kinase 1
MARCH5membrane-associated ring finger (C3HC4) 5
PPARperoxisome proliferator-accelerated receptor
HIFhypoxia-inducible factor
CREBcAMP response element binding protein
NRF2nuclear factor erythroid 2-related factor 2
NBDnucleotide binding domain
CAPScryopyrin associated periodic syndromes
MyD88myeloid differentiation primary response 88
TRIFTIR-domain-containing adapter-inducing interferon β
MAVSmitochondrial antiviral signaling protein
TUFMTu translation elongation factor
MPTmitochondrial permeability transition
MOMPmitochondrial outer membrane permeabilization

References:

1. Archibald JM. Endosymbiosis and Eukaryotic Cell Evolution. Curr Biol 2015;25(19):R911–21. [PubMed] [Google Scholar]
2. Martin WF, Garg S, Zimorski V. Endosymbiotic theories for eukaryote origin. Philos Trans R Soc Lond B Biol Sci 2015;370(1678):20140330. [PMC free article] [PubMed] [Google Scholar]
3. Ryan MT, Hoogenraad NJ. Mitochondrial-nuclear communications. Annu Rev Biochem 2007;76:701–22. [PubMed] [Google Scholar]
4. Bonawitz ND, Clayton DA, Shadel GS. Initiation and beyond: multiple functions of the human mitochondrial transcription machinery. Mol Cell. 2006;24(6):813–25. [PubMed] [Google Scholar]
5. Bereiter-Hahn J Behavior of mitochondria in the living cell. Int Rev Cytol 1990;122:1–63. [PubMed] [Google Scholar]
6. Nunnari J, Marshall WF, Straight A, Murray A, Sedat JW, Walter P. Mitochondrial transmission during mating in Saccharomyces cerevisiae is determined by mitochondrial fusion and fission and the intramitochondrial segregation of mitochondrial DNA. Mol Biol Cell. 1997;8(7):1233–42. [PMC free article] [PubMed] [Google Scholar]
7. Vafai SB, Mootha VK. Mitochondrial disorders as windows into an ancient organelle. Nature. 2012;491(7424):374–83. [PubMed] [Google Scholar]
8. Berry BJ, Trewin AJ, Amitrano AM, Kim M, Wojtovich AP. Use the protonmotive force: Mitochondrial uncoupling and reactive oxygen species. J Mol Biol 2018. [PMC free article] [PubMed] [Google Scholar]
9. Kanaan GN, Harper ME. Cellular redox dysfunction in the development of cardiovascular diseases. Biochim Biophys Acta 2017;1861(11 Pt A):2822–9. [PubMed] [Google Scholar]
10. Nakahira K, Hisata S, Choi AM. The Roles of Mitochondrial Damage-Associated Molecular Patterns in Diseases. Antioxid Redox Signal. 2015;23(17):1329–50. [PMC free article] [PubMed] [Google Scholar]
11. Youle RJ, van der Bliek AM. Mitochondrial fission, fusion, and stress. Science. 2012;337(6098):1062–5. [PMC free article] [PubMed] [Google Scholar]
12. Janeway CA Jr., Medzhitov R. Innate immune recognition. Annu Rev Immunol 2002;20:197–216. [PubMed] [Google Scholar]
13. Medzhitov R. Innate immunity: quo vadis? Nat Immunol 2010;11(7):551–3. [PubMed] [Google Scholar]
14. Iwasaki A, Medzhitov R. Regulation of adaptive immunity by the innate immune system. Science. 2010;327(5963):291–5. [PMC free article] [PubMed] [Google Scholar]
15. Iwasaki A, Medzhitov R. Control of adaptive immunity by the innate immune system. Nat Immunol 2015;16(4):343–53. [PMC free article] [PubMed] [Google Scholar]
16. Antonelli M, Kushner I. It’s time to redefine inflammation. FASEB J 2017;31(5):1787–91. [PubMed] [Google Scholar]
17. Manthiram K, Zhou Q, Aksentijevich I, Kastner DL. The monogenic autoinflammatory diseases define new pathways in human innate immunity and inflammation. Nat Immunol 2017;18(8):832–42. [PubMed] [Google Scholar]
18. Slaats J, Ten Oever J, van de Veerdonk FL, Netea MG. IL-1beta/IL-6/CRP and IL-18/ferritin: Distinct Inflammatory Programs in Infections. PLoS Pathog 2016;12(12):e1005973. [PMC free article] [PubMed] [Google Scholar]
19. Bernardi P The mitochondrial permeability transition pore: a mystery solved? Front Physiol 2013;4:95. [PMC free article] [PubMed] [Google Scholar]
20. Kalkavan H, Green DR. MOMP, cell suicide as a BCL-2 family business. Cell Death Differ 2018;25(1):46–55. [PMC free article] [PubMed] [Google Scholar]
21. Hunter FE Jr., Ford L. Inactivation of oxidative and phosphorylative systems in mitochondria by preincubation with phosphate and other ions. J Biol Chem 1955;216(1):357–69. [PubMed] [Google Scholar]
22. Izzo V, Bravo-San Pedro JM, Sica V, Kroemer G, Galluzzi L. Mitochondrial Permeability Transition: New Findings and Persisting Uncertainties. Trends Cell Biol 2016;26(9):655–67. [PubMed] [Google Scholar]
23. Czabotar PE, Lessene G, Strasser A, Adams JM. Control of apoptosis by the BCL-2 protein family: implications for physiology and therapy. Nat Rev Mol Cell Biol 2014;15(1):49–63. [PubMed] [Google Scholar]
24. McArthur K, Whitehead LW, Heddleston JM, Li L, Padman BS, Oorschot V, et al. BAK/BAX macropores facilitate mitochondrial herniation and mtDNA efflux during apoptosis. Science. 2018;359(6378). [PubMed] [Google Scholar]
25. Tait SW, Parsons MJ, Llambi F, Bouchier-Hayes L, Connell S, Munoz-Pinedo C, et al. Resistance to caspase-independent cell death requires persistence of intact mitochondria. Dev Cell. 2010;18(5):802–13. [PMC free article] [PubMed] [Google Scholar]
26. Spencer AC, Spremulli LL. Interaction of mitochondrial initiation factor 2 with mitochondrial fMet-tRNA. Nucleic Acids Res 2004;32(18):5464–70. [PMC free article] [PubMed] [Google Scholar]
27. Schiffmann E, Corcoran BA, Wahl SM. N-formylmethionyl peptides as chemoattractants for leucocytes. Proc Natl Acad Sci U S A 1975;72(3):1059–62. [PMC free article] [PubMed] [Google Scholar]
28. Carp H. Mitochondrial N-formylmethionyl proteins as chemoattractants for neutrophils. J Exp Med 1982;155(1):264–75. [PMC free article] [PubMed] [Google Scholar]
29. Wenceslau CF, McCarthy CG, Goulopoulou S, Szasz T, NeSmith EG, Webb RC. Mitochondrial- derived N-formyl peptides: novel links between trauma, vascular collapse and sepsis. Med Hypotheses. 2013;81(4):532–5. [PMC free article] [PubMed] [Google Scholar]
30. Wenceslau CF, McCarthy CG, Webb RC. Formyl Peptide Receptor Activation Elicits Endothelial Cell Contraction and Vascular Leakage. Front Immunol 2016;7:297. [PMC free article] [PubMed] [Google Scholar]
31. Dorward DA, Lucas CD, Doherty MK, Chapman GB, Scholefield EJ, Conway Morris A, et al. Novel role for endogenous mitochondrial formylated peptide-driven formyl peptide receptor 1 signalling in acute respiratory distress syndrome. Thorax. 2017;72(10):928–36. [PMC free article] [PubMed] [Google Scholar]
32. Ren M, Phoon CK, Schlame M. Metabolism and function of mitochondrial cardiolipin. Prog Lipid Res 2014;55:1–16. [PubMed] [Google Scholar]
33. Shen Z, Ye C, McCain K, Greenberg ML. The Role of Cardiolipin in Cardiovascular Health. Biomed Res Int 2015;2015:891707. [PMC free article] [PubMed] [Google Scholar]
34. Gebert N, Joshi AS, Kutik S, Becker T, McKenzie M, Guan XL, et al. Mitochondrial cardiolipin involved in outer-membrane protein biogenesis: implications for Barth syndrome. Curr Biol 2009;19(24):2133–9. [PMC free article] [PubMed] [Google Scholar]
35. Schlame M, Greenberg ML. Biosynthesis, remodeling and turnover of mitochondrial cardiolipin. Biochim Biophys Acta. 2017;1862(1):3–7. [PMC free article] [PubMed] [Google Scholar]
36. Tatsuta T, Langer T. Intramitochondrial phospholipid trafficking. Biochim Biophys Acta. 2017;1862(1):81–9. [PubMed] [Google Scholar]
37. Schlame M, Ren M. The role of cardiolipin in the structural organization of mitochondrial membranes. Biochim Biophys Acta. 2009;1788(10):2080–3. [PMC free article] [PubMed] [Google Scholar]
38. Dudek J Role of Cardiolipin in Mitochondrial Signaling Pathways. Front Cell Dev Biol 2017;5:90. [PMC free article] [PubMed] [Google Scholar]
39. Elliott EI, Sutterwala FS. Initiation and perpetuation of NLRP3 inflammasome activation and assembly. Immunol Rev 2015;265(1):35–52. [PMC free article] [PubMed] [Google Scholar]
40. Iyer SS, He Q, Janczy JR, Elliott EI, Zhong Z, Olivier AK, et al. Mitochondrial cardiolipin is required for Nlrp3 inflammasome activation. Immunity. 2013;39(2):311–23. [PMC free article] [PubMed] [Google Scholar]
41. Toksoy A, Sennefelder H, Adam C, Hofmann S, Trautmann A, Goebeler M, et al. Potent NLRP3 Inflammasome Activation by the HIV Reverse Transcriptase Inhibitor Abacavir. J Biol Chem 2017;292(7):2805–14. [PMC free article] [PubMed] [Google Scholar]
42. Rossen RD, Michael LH, Hawkins HK, Youker K, Dreyer WJ, Baughn RE, et al. Cardiolipin-protein complexes and initiation of complement activation after coronary artery occlusion. Circ Res 1994;75(3):546–55. [PubMed] [Google Scholar]
43. Chu CT, Ji J, Dagda RK, Jiang JF, Tyurina YY, Kapralov AA, et al. Cardiolipin externalization to the outer mitochondrial membrane acts as an elimination signal for mitophagy in neuronal cells. Nat Cell Biol 2013;15(10):1197–205. [PMC free article] [PubMed] [Google Scholar]
44. Balasubramanian K, Maeda A, Lee JS, Mohammadyani D, Dar HH, Jiang JF, et al. Dichotomous roles for externalized cardiolipin in extracellular signaling: Promotion of phagocytosis and attenuation of innate immunity. Sci Signal. 2015;8(395):ra95. [PMC free article] [PubMed] [Google Scholar]
45. Thorslund T, Sunesen M, Bohr VA, Stevnsner T. Repair of 8-oxoG is slower in endogenous nuclear genes than in mitochondrial DNA and is without strand bias. DNA Repair (Amst) 2002;1(4):261–73. [PubMed] [Google Scholar]
46. Alexeyev M, Shokolenko I, Wilson G, LeDoux S. The maintenance of mitochondrial DNA integrity-critical analysis and update. Cold Spring Harb Perspect Biol 2013;5(5):a012641. [PMC free article] [PubMed] [Google Scholar]
47. Collins LV, Hajizadeh S, Holme E, Jonsson IM, Tarkowski A. Endogenously oxidized mitochondrial DNA induces in vivo and in vitro inflammatory responses. J Leukoc Biol 2004;75(6):995–1000. [PubMed] [Google Scholar]
48. Hauser CJ, Sursal T, Rodriguez EK, Appleton PT, Zhang Q, Itagaki K. Mitochondrial damage associated molecular patterns from femoral reamings activate neutrophils through formyl peptide receptors and P44/42 MAP kinase. J Orthop Trauma. 2010;24(9):534–8. [PMC free article] [PubMed] [Google Scholar]
49. Zhang Q, Raoof M, Chen Y, Sumi Y, Sursal T, Junger W, et al. Circulating mitochondrial DAMPs cause inflammatory responses to injury. Nature. 2010;464(7285):104–7. [PMC free article] [PubMed] [Google Scholar]
50. Zhang Q, Itagaki K, Hauser CJ. Mitochondrial DNA is released by shock and activates neutrophils via p38 map kinase. Shock. 2010;34(1):55–9. [PubMed] [Google Scholar]
51. Rodero MP, Crow YJ. Type I interferon-mediated monogenic autoinflammation: The type I interferonopathies, a conceptual overview. J Exp Med 2016;213(12):2527–38. [PMC free article] [PubMed] [Google Scholar]
52. West AP, Shadel GS, Ghosh S. Mitochondria in innate immune responses. Nat Rev Immunol 2011;11(6):389–402. [PMC free article] [PubMed] [Google Scholar]
53. Kroller-Schon S, Steven S, Kossmann S, Scholz A, Daub S, Oelze M, et al. Molecular mechanisms of the crosstalk between mitochondria and NADPH oxidase through reactive oxygen species-studies in white blood cells and in animal models. Antioxid Redox Signal. 2014;20(2):247–66. [PMC free article] [PubMed] [Google Scholar]
54. Lambeth JD. NOX enzymes and the biology of reactive oxygen. Nat Rev Immunol 2004;4(3):181–9. [PubMed] [Google Scholar]
55. West AP, Brodsky IE, Rahner C, Woo DK, Erdjument-Bromage H, Tempst P, et al. TLR signalling augments macrophage bactericidal activity through mitochondrial ROS. Nature. 2011;472(7344):476–U543. [PMC free article] [PubMed] [Google Scholar]
56. Brunelle JK, Bell EL, Quesada NM, Vercauteren K, Tiranti V, Zeviani M, et al. Oxygen sensing requires mitochondrial ROS but not oxidative phosphorylation. Cell Metab 2005;1(6):409–14. [PubMed] [Google Scholar]
57. Chandel NS, Trzyna WC, McClintock DS, Schumacker PT. Role of oxidants in NF-kappa B activation and TNF-alpha gene transcription induced by hypoxia and endotoxin. J Immunol 2000;165(2):1013–21. [PubMed] [Google Scholar]
58. Gottfredsen RH, Goldstrohm DA, Hartney JM, Larsen UG, Bowler RP, Petersen SV. The cellular distribution of extracellular superoxide dismutase in macrophages is altered by cellular activation but unaffected by the naturally occurring R213G substitution. Free Radic Biol Med 2014;69:348–56. [PMC free article] [PubMed] [Google Scholar]
59. Karnati S, Luers G, Pfreimer S, Baumgart-Vogt E. Mammalian SOD2 is exclusively located in mitochondria and not present in peroxisomes. Histochem Cell Biol 2013;140(2):105–17. [PubMed] [Google Scholar]
60. Kawamata H, Manfredi G. Import, maturation, and function of SOD1 and its copper chaperone CCS in the mitochondrial intermembrane space. Antioxid Redox Signal. 2010;13(9):1375–84. [PMC free article] [PubMed] [Google Scholar]
61. Vozza A, Parisi G, De Leonardis F, Lasorsa FM, Castegna A, Amorese D, et al. UCP2 transports C4 metabolites out of mitochondria, regulating glucose and glutamine oxidation. Proc Natl Acad Sci U S A 2014;111(3):960–5. [PMC free article] [PubMed] [Google Scholar]
62. Lupfer C, Thomas PG, Anand PK, Vogel P, Milasta S, Martinez J, et al. Receptor interacting protein kinase 2-mediated mitophagy regulates inflammasome activation during virus infection. Nat Immunol 2013;14(5):480–8. [PMC free article] [PubMed] [Google Scholar]
63. Iyer SS, Pulskens WP, Sadler JJ, Butter LM, Teske GJ, Ulland TK, et al. Necrotic cells trigger a sterile inflammatory response through the Nlrp3 inflammasome. Proc Natl Acad Sci U S A 2009;106(48):20388–93. [PMC free article] [PubMed] [Google Scholar]
64. Elliott MR, Chekeni FB, Trampont PC, Lazarowski ER, Kadl A, Walk SF, et al. Nucleotides released by apoptotic cells act as a find-me signal to promote phagocytic clearance. Nature. 2009;461(7261):282–6. [PMC free article] [PubMed] [Google Scholar]
65. Kang D, Kim SH, Hamasaki N. Mitochondrial transcription factor A (TFAM): roles in maintenance of mtDNA and cellular functions. Mitochondrion. 2007;7(1–2):39–44. [PubMed] [Google Scholar]
66. Chaung WW, Wu R, Ji Y, Dong W, Wang P. Mitochondrial transcription factor A is a proinflammatory mediator in hemorrhagic shock. Int J Mol Med 2012;30(1):199–203. [PMC free article] [PubMed] [Google Scholar]
67. Julian MW, Shao G, Vangundy ZC, Papenfuss TL, Crouser ED Mitochondrial transcription factor A, an endogenous danger signal, promotes TNFalpha release via RAGE- and TLR9-responsive plasmacytoid dendritic cells. PLoS One. 2013;8(8):e72354. [PMC free article] [PubMed] [Google Scholar]
68. West AP, Shadel GS. Mitochondrial DNA in innate immune responses and inflammatory pathology. Nat Rev Immunol 2017;17(6):363–75. [PMC free article] [PubMed] [Google Scholar]
69. Kumar H, Kawai T, Akira S. Pathogen recognition by the innate immune system. Int Rev Immunol 2011;30(1):16–34. [PubMed] [Google Scholar]
70. Carneiro FRG, Lepelley A, Seeley JJ, Hayden MS, Ghosh S. An Essential Role for ECSIT in Mitochondrial Complex I Assembly and Mitophagy in Macrophages. Cell Rep 2018;22(10):2654–66. [PMC free article] [PubMed] [Google Scholar]
71. Schreck R, Rieber P, Baeuerle PA. Reactive oxygen intermediates as apparently widely used messengers in the activation of the NF-kappa B transcription factor and HIV-1. EMBO J 1991;10(8):2247–58. [PMC free article] [PubMed] [Google Scholar]
72. Matsuzawa A, Saegusa K, Noguchi T, Sadamitsu C, Nishitoh H, Nagai S, et al. ROS-dependent activation of the TRAF6-ASK1-p38 pathway is selectively required for TLR4-mediated innate immunity. Nat Immunol 2005;6(6):587–92. [PubMed] [Google Scholar]
73. Wi SM, Moon G, Kim J, Kim ST, Shim JH, Chun E, et al. TAK1-ECSIT-TRAF6 complex plays a key role in the TLR4 signal to activate NF-kappaB. J Biol Chem 2014;289(51):35205–14. [PMC free article] [PubMed] [Google Scholar]
74. Shi HX, Liu X, Wang Q, Tang PP, Liu XY, Shan YF, et al. Mitochondrial ubiquitin ligase MARCH5 promotes TLR7 signaling by attenuating TANK action. PLoS Pathog 2011;7(5):e1002057. [PMC free article] [PubMed] [Google Scholar]
75. Suliman HB, Welty-Wolf KE, Carraway MS, Schwartz DA, Hollingsworth JW, Piantadosi CA. Tolllike receptor 4 mediates mitochondrial DNA damage and biogenic responses after heat-inactivated E. coli. FASEB J 2005;19(11):1531–3. [PubMed] [Google Scholar]
76. Sweeney TE, Suliman HB, Hollingsworth JW, Piantadosi CA. Differential regulation of the PGC family of genes in a mouse model of Staphylococcus aureus sepsis. PLoS One. 2010;5(7):e11606. [PMC free article] [PubMed] [Google Scholar]
77. Suliman HB, Sweeney TE, Withers CM, Piantadosi CA. Co-regulation of nuclear respiratory factor-1 by NFkappaB and CREB links LPS-induced inflammation to mitochondrial biogenesis. J Cell Sci 2010;123(Pt 15):2565–75. [PMC free article] [PubMed] [Google Scholar]
78. Sweeney TE, Suliman HB, Hollingsworth JW, Welty-Wolf KE, Piantadosi CA. A toll-like receptor 2 pathway regulates the Ppargc1a/b metabolic co-activators in mice with Staphylococcal aureus sepsis. PLoS One. 2011;6(9):e25249. [PMC free article] [PubMed] [Google Scholar]
79. Shi J, Zhao Y, Wang K, Shi X, Wang Y, Huang H, et al. Cleavage of GSDMD by inflammatory caspases determines pyroptotic cell death. Nature. 2015;526(7575):660–5. [PubMed] [Google Scholar]
80. Kayagaki N, Stowe IB, Lee BL, O’Rourke K, Anderson K, Warming S, et al. Caspase-11 cleaves gasdermin D for non-canonical inflammasome signalling. Nature. 2015;526(7575):666–71. [PubMed] [Google Scholar]
81. Hoffman HM, Mueller JL, Broide DH, Wanderer AA, Kolodner RD. Mutation of a new gene encoding a putative pyrin-like protein causes familial cold autoinflammatory syndrome and Muckle- Wells syndrome. Nat Genet 2001;29(3):301–5. [PMC free article] [PubMed] [Google Scholar]
82. Aganna E, Martinon F, Hawkins PN, Ross JB, Swan DC, Booth DR, et al. Association of mutations in the NALP3/CIAS1/PYPAF1 gene with a broad phenotype including recurrent fever, cold sensitivity, sensorineural deafness, and AA amyloidosis. Arthritis Rheum 2002;46(9):2445–52. [PubMed] [Google Scholar]
83. Aksentijevich I, Nowak M, Mallah M, Chae JJ, Watford WT, Hofmann SR, et al. De novo CIAS1 mutations, cytokine activation, and evidence for genetic heterogeneity in patients with neonatal-onset multisystem inflammatory disease (NOMID): a new member of the expanding family of pyrin-associated autoinflammatory diseases. Arthritis Rheum 2002;46(12):3340–8. [PMC free article] [PubMed] [Google Scholar]
84. Guarda G, Zenger M, Yazdi AS, Schroder K, Ferrero I, Menu P, et al. Differential expression of NLRP3 among hematopoietic cells. J Immunol 2011;186(4):2529–34. [PubMed] [Google Scholar]
85. Bauernfeind FG, Horvath G, Stutz A, Alnemri ES, MacDonald K, Speert D, et al. Cutting edge: NF- kappaB activating pattern recognition and cytokine receptors license NLRP3 inflammasome activation by regulating NLRP3 expression. J Immunol 2009;183(2):787–91. [PMC free article] [PubMed] [Google Scholar]
86. Franchi L, Eigenbrod T, Nunez G. Cutting edge: TNF-alpha mediates sensitization to ATP and silica via the NLRP3 inflammasome in the absence of microbial stimulation. J Immunol 2009;183(2):792–6. [PMC free article] [PubMed] [Google Scholar]
87. Juliana C, Fernandes-Alnemri T, Kang S, Farias A, Qin F, Alnemri ES. Non-transcriptional priming and deubiquitination regulate NLRP3 inflammasome activation. J Biol Chem 2012;287(43):36617–22. [PMC free article] [PubMed] [Google Scholar]
88. Schroder K, Sagulenko V, Zamoshnikova A, Richards AA, Cridland JA, Irvine KM, et al. Acute lipopolysaccharide priming boosts inflammasome activation independently of inflammasome sensor induction. Immunobiology. 2012;217(12):1325–9. [PubMed] [Google Scholar]
89. Py BF, Kim MS, Vakifahmetoglu-Norberg H, Yuan J. Deubiquitination of NLRP3 by BRCC3 critically regulates inflammasome activity. Mol Cell. 2013;49(2):331–8. [PubMed] [Google Scholar]
90. Stutz A, Kolbe CC, Stahl R, Horvath GL, Franklin BS, van Ray O, et al. NLRP3 inflammasome assembly is regulated by phosphorylation of the pyrin domain. J Exp Med 2017;214(6):1725–36. [PMC free article] [PubMed] [Google Scholar]
91. Song N, Liu ZS, Xue W, Bai ZF, Wang QY, Dai J, et al. NLRP3 Phosphorylation Is an Essential Priming Event for Inflammasome Activation. Mol Cell. 2017;68(1):185–97 e6. [PubMed] [Google Scholar]
92. Hara H, Tsuchiya K, Kawamura I, Fang R, Hernandez-Cuellar E, Shen Y, et al. Phosphorylation of the adaptor ASC acts as a molecular switch that controls the formation of speck-like aggregates and inflammasome activity. Nat Immunol 2013;14(12):1247–55. [PMC free article] [PubMed] [Google Scholar]
93. Lopez-Castejon G, Luheshi NM, Compan V, High S, Whitehead RC, Flitsch S, et al. Deubiquitinases regulate the activity of caspase-1 and interleukin-1beta secretion via assembly of the inflammasome. J Biol Chem 2013;288(4):2721–33. [PMC free article] [PubMed] [Google Scholar]
94. Rodgers MA, Bowman JW, Fujita H, Orazio N, Shi M, Liang Q, et al. The linear ubiquitin assembly complex (LUBAC) is essential for NLRP3 inflammasome activation. J Exp Med 2014;211(7):1333–47. [PMC free article] [PubMed] [Google Scholar]
95. Weng L, Mitoma H, Trichot C, Bao M, Liu Y, Zhang Z, et al. The E3 ubiquitin ligase tripartite motif 33 is essential for cytosolic RNA-induced NLRP3 inflammasome activation. J Immunol 2014;193(7):3676–82. [PMC free article] [PubMed] [Google Scholar]
96. Zhou R, Yazdi AS, Menu P, Tschopp J. A role for mitochondria in NLRP3 inflammasome activation. Nature. 2011;469(7329):221–5. [PubMed] [Google Scholar]
97. Elliott EI, Miller AN, Banoth B, Iyer SS, Stotland A, Weiss JP, et al. Cutting Edge: Mitochondrial Assembly of the NLRP3 Inflammasome Complex Is Initiated at Priming. J Immunol 2018. [PMC free article] [PubMed] [Google Scholar]
98. Subramanian N, Natarajan K, Clatworthy MR, Wang Z, Germain RN. The adaptor MAVS promotes NLRP3 mitochondrial localization and inflammasome activation. Cell. 2013;153(2):348–61. [PMC free article] [PubMed] [Google Scholar]
99. Park S, Juliana C, Hong S, Datta P, Hwang I, Fernandes-Alnemri T, et al. The mitochondrial antiviral protein MAVS associates with NLRP3 and regulates its inflammasome activity. J Immunol 2013;191(8):4358–66. [PMC free article] [PubMed] [Google Scholar]
100. Wu YH, Kuo WC, Wu YJ, Yang KT, Chen ST, Jiang ST, et al. Participation of c-FLIP in NLRP3 and AIM2 inflammasome activation. Cell Death Differ 2014;21(3):451–61. [PMC free article] [PubMed] [Google Scholar]
101. Nakahira K, Haspel JA, Rathinam VA, Lee SJ, Dolinay T, Lam HC, et al. Autophagy proteins regulate innate immune responses by inhibiting the release of mitochondrial DNA mediated by the NALP3 inflammasome. Nat Immunol 2011;12(3):222–30. [PMC free article] [PubMed] [Google Scholar]
102. Shimada K, Crother TR, Karlin J, Dagvadorj J, Chiba N, Chen S, et al. Oxidized mitochondrial DNA activates the NLRP3 inflammasome during apoptosis. Immunity. 2012;36(3):401–14. [PMC free article] [PubMed] [Google Scholar]
103. Misawa T, Takahama M, Kozaki T, Lee H, Zou J, Saitoh T, et al. Microtubule-driven spatial arrangement of mitochondria promotes activation of the NLRP3 inflammasome. Nat Immunol 2013;14(5):454–60. [PubMed] [Google Scholar]
104. Murakami T, Ockinger J, Yu J, Byles V, McColl A, Hofer AM, et al. Critical role for calcium mobilization in activation of the NLRP3 inflammasome. Proc Natl Acad Sci U S A 2012;109(28):11282–7. [PMC free article] [PubMed] [Google Scholar]
105. Zhong Z, Zhai Y, Liang S, Mori Y, Han R, Sutterwala F, et al. TRPM2 links oxidative stress to the NLRP3 inflammasome activation (P1268). J Immunol 2013;190. [PMC free article] [PubMed] [Google Scholar]
106. Ichinohe T, Yamazaki T, Koshiba T, Yanagi Y. Mitochondrial protein mitofusin 2 is required for NLRP3 inflammasome activation after RNA virus infection. Proc Natl Acad Sci U S A 2013;110(44):17963–8. [PMC free article] [PubMed] [Google Scholar]
107. Moore CB, Bergstralh DT, Duncan JA, Lei Y, Morrison TE, Zimmermann AG, et al. NLRX1 is a regulator of mitochondrial antiviral immunity. Nature. 2008;451(7178):573–7. [PubMed] [Google Scholar]
108. Tattoli I, Carneiro LA, Jehanno M, Magalhaes JG, Shu Y, Philpott DJ, et al. NLRX1 is a mitochondrial NOD-like receptor that amplifies NF-kappaB and JNK pathways by inducing reactive oxygen species production. EMBO Rep 2008;9(3):293–300. [PMC free article] [PubMed] [Google Scholar]
109. Guo H, Konig R, Deng M, Riess M, Mo J, Zhang L, et al. NLRX1 Sequesters STING to Negatively Regulate the Interferon Response, Thereby Facilitating the Replication of HIV-1 and DNA Viruses. Cell Host Microbe 2016;19(4):515–28. [PMC free article] [PubMed] [Google Scholar]
110. Xia X, Cui J, Wang HY, Zhu L, Matsueda S, Wang Q, et al. NLRX1 negatively regulates TLR-induced NF-kappaB signaling by targeting TRAF6 and IKK. Immunity. 2011;34(6):843–53. [PMC free article] [PubMed] [Google Scholar]
111. Soares F, Tattoli I, Wortzman ME, Arnoult D, Philpott DJ, Girardin SE. NLRX1 does not inhibit MAVS-dependent antiviral signalling. Innate Immun 2013;19(4):438–48. [PubMed] [Google Scholar]
112. Rebsamen M, Vazquez J, Tardivel A, Guarda G, Curran J, Tschopp J. NLRX1/NOD5 deficiency does not affect MAVS signalling. Cell Death Differ 2011;18(8):1387. [PMC free article] [PubMed] [Google Scholar]
113. Feng H, Lenarcic EM, Yamane D, Wauthier E, Mo J, Guo H, et al. NLRX1 promotes immediate IRF1-directed antiviral responses by limiting dsRNA-activated translational inhibition mediated by PKR. Nat Immunol 2017;18(12):1299–309. [PMC free article] [PubMed] [Google Scholar]
114. Arnoult D, Soares F, Tattoli I, Castanier C, Philpott DJ, Girardin SE. An N-terminal addressing sequence targets NLRX1 to the mitochondrial matrix. J Cell Sci. 2009;122(Pt 17):3161–8. [PMC free article] [PubMed] [Google Scholar]
115. Abdul-Sater AA, Said-Sadier N, Lam VM, Singh B, Pettengill MA, Soares F, et al. Enhancement of reactive oxygen species production and chlamydial infection by the mitochondrial Nod-like family member NLRX1. J Biol Chem 2010;285(53):41637–45. [PMC free article] [PubMed] [Google Scholar]
116. Lei Y, Wen H, Yu Y, Taxman DJ, Zhang L, Widman DG, et al. The mitochondrial proteins NLRX1 and TUFM form a complex that regulates type I interferon and autophagy. Immunity. 2012;36(6):933–46. [PMC free article] [PubMed] [Google Scholar]
117. Unger BL, Ganesan S, Comstock AT, Faris AN, Hershenson MB, Sajjan US. Nod-like receptor X-1 is required for rhinovirus-induced barrier dysfunction in airway epithelial cells. J Virol 2014;88(7):3705–18. [PMC free article] [PubMed] [Google Scholar]
118. Stokman G, Kors L, Bakker PJ, Rampanelli E, Claessen N, Teske GJD, et al. NLRX1 dampens oxidative stress and apoptosis in tissue injury via control of mitochondrial activity. J Exp Med 2017;214(8):2405–20. [PMC free article] [PubMed] [Google Scholar]
119. Costford SR, Tattoli I, Duan FT, Volchuk A, Klip A, Philpott DJ, et al. Male Mice Lacking NLRX1 Are Partially Protected From High-Fat Diet-Induced Hyperglycemia. J Endocr Soc 2018;2(4):336–47. [PMC free article] [PubMed] [Google Scholar]
120. Kato H, Sato S, Yoneyama M, Yamamoto M, Uematsu S, Matsui K, et al. Cell type-specific involvement of RIG-I in antiviral response. Immunity. 2005;23(1):19–28. [PubMed] [Google Scholar]
121. Yoneyama M, Kikuchi M, Natsukawa T, Shinobu N, Imaizumi T, Miyagishi M, et al. The RNA helicase RIG-I has an essential function in double-stranded RNA-induced innate antiviral responses. Nat Immunol 2004;5(7):730–7. [PubMed] [Google Scholar]
122. Seth RB, Sun L, Ea CK, Chen ZJ. Identification and characterization of MAVS, a mitochondrial antiviral signaling protein that activates NF-kappaB and IRF 3. Cell. 2005;122(5):669–82. [PubMed] [Google Scholar]
123. Meylan E, Curran J, Hofmann K, Moradpour D, Binder M, Bartenschlager R, et al. Cardif is an adaptor protein in the RIG-I antiviral pathway and is targeted by hepatitis C virus. Nature. 2005;437(7062):1167–72. [PubMed] [Google Scholar]
124. Xu LG, Wang YY, Han KJ, Li LY, Zhai Z, Shu HB. VISA is an adapter protein required for virus- triggered IFN-beta signaling. Mol Cell. 2005;19(6):727–40. [PubMed] [Google Scholar]
125. Kawai T, Takahashi K, Sato S, Coban C, Kumar H, Kato H, et al. I PS-1, an adaptor triggering RIG-I- and Mda5-mediated type I interferon induction. Nat Immunol 2005;6(10):981–8. [PubMed] [Google Scholar]
126. Hornung V, Ellegast J, Kim S, Brzozka K, Jung A, Kato H, et al. 5’-Triphosphate RNA is the ligand for RIG-I. Science. 2006;314(5801):994–7. [PubMed] [Google Scholar]
127. Kato H, Takeuchi O, Sato S, Yoneyama M, Yamamoto M, Matsui K, et al. Differential roles of MDA5 and RIG-I helicases in the recognition of RNA viruses. Nature. 2006;441(7089):101–5. [PubMed] [Google Scholar]
128. Pichlmair A, Schulz O, Tan CP, Naslund TI, Liljestrom P, Weber F, et al. RIG-I-mediated antiviral responses to single-stranded RNA bearing 5’-phosphates. Science. 2006;314(5801):997–1001. [PubMed] [Google Scholar]
129. Satoh T, Kato H, Kumagai Y, Yoneyama M, Sato S, Matsushita K, et al. LGP2 is a positive regulator of RIG-I- and MDA5-mediated antiviral responses. Proc Natl Acad Sci U S A 2010;107(4):1512–7. [PMC free article] [PubMed] [Google Scholar]
130. Childs KS, Randall RE, Goodbourn S. LGP2 plays a critical role in sensitizing mda-5 to activation by double-stranded RNA. PLoS One. 2013;8(5):e64202. [PMC free article] [PubMed] [Google Scholar]
131. Deddouche S, Goubau D, Rehwinkel J, Chakravarty P, Begum S, Maillard PV, et al. Identification of an LGP2-associated MDA5 agonist in picornavirus-infected cells. Elife. 2014;3:e01535. [PMC free article] [PubMed] [Google Scholar]
132. Rothenfusser S, Goutagny N, DiPerna G, Gong M, Monks BG, Schoenemeyer A, et al. The RNA helicase Lgp2 inhibits TLR-independent sensing of viral replication by retinoic acid-inducible gene-I. J Immunol 2005;175(8):5260–8. [PubMed] [Google Scholar]
133. Yoneyama M, Kikuchi M, Matsumoto K, Imaizumi T, Miyagishi M, Taira K, et al. Shared and unique functions of the DExD/H-box helicases RIG-I, MDA5, and LGP2 in antiviral innate immunity. J Immunol 2005;175(5):2851–8. [PubMed] [Google Scholar]
134. Dixit E, Boulant S, Zhang Y, Lee AS, Odendall C, Shum B, et al. Peroxisomes are signaling platforms for antiviral innate immunity. Cell. 2010;141(4):668–81. [PMC free article] [PubMed] [Google Scholar]
135. Odendall C, Dixit E, Stavru F, Bierne H, Franz KM, Durbin AF, et al. Diverse intracellular pathogens activate type III interferon expression from peroxisomes. Nat Immunol 2014;15(8):717–26. [PMC free article] [PubMed] [Google Scholar]
136. Hou F, Sun L, Zheng H, Skaug B, Jiang QX, Chen ZJ. MAVS forms functional prion-like aggregates to activate and propagate antiviral innate immune response. Cell. 2011;146(3):448–61. [PMC free article] [PubMed] [Google Scholar]
137. Peisley A, Wu B, Xu H, Chen ZJ, Hur S. Structural basis for ubiquitin-mediated antiviral signal activation by RIG-I. Nature. 2014;509(7498):110–4. [PMC free article] [PubMed] [Google Scholar]
138. Takeuchi O, Akira S. Pattern recognition receptors and inflammation. Cell. 2010;140(6):805–20. [PubMed] [Google Scholar]
139. Gack MU, Shin YC, Joo CH, Urano T, Liang C, Sun L, et al. TRIM25 RING-finger E3 ubiquitin ligase is essential for RIG-I-mediated antiviral activity. Nature. 2007;446(7138):916–20. [PubMed] [Google Scholar]
140. Oshiumi H, Matsumoto M, Seya T. Ubiquitin-mediated modulation of the cytoplasmic viral RNA sensor RIG-I. J Biochem 2012;151(1):5–11. [PubMed] [Google Scholar]
141. Wies E, Wang MK, Maharaj NP, Chen K, Zhou S, Finberg RW, et al. Dephosphorylation of the RNA sensors RIG-I and MDA5 by the phosphatase PP1 is essential for innate immune signaling. Immunity. 2013;38(3):437–49. [PMC free article] [PubMed] [Google Scholar]
142. Zeng W, Sun L, Jiang X, Chen X, Hou F, Adhikari A, et al. Reconstitution of the RIG-I pathway reveals a signaling role of unanchored polyubiquitin chains in innate immunity. Cell. 2010;141(2):315–30. [PMC free article] [PubMed] [Google Scholar]
143. Zhong B, Zhang Y, Tan B, Liu TT, Wang YY, Shu HB. The E3 ubiquitin ligase RNF5 targets virus- induced signaling adaptor for ubiquitination and degradation. J Immunol 2010;184(11):6249–55. [PubMed] [Google Scholar]
144. Yasukawa K, Oshiumi H, Takeda M, Ishihara N, Yanagi Y, Seya T, et al. Mitofusin 2 inhibits mitochondrial antiviral signaling. Sci Signal 2009;2(84):ra47. [PubMed] [Google Scholar]
145. Castanier C, Garcin D, Vazquez A, Arnoult D. Mitochondrial dynamics regulate the RIG-I-like receptor antiviral pathway. EMBO Rep 2010;11(2):133–8. [PMC free article] [PubMed] [Google Scholar]
146. Onoguchi K, Onomoto K, Takamatsu S, Jogi M, Takemura A, Morimoto S, et al. Virus-infection or 5’ppp-RNA activates antiviral signal through redistribution of I PS-1 mediated by MFN1. PLoS Pathog 2010;6(7):e1001012. [PMC free article] [PubMed] [Google Scholar]
147. Vitour D, Dabo S, Ahmadi Pour M, Vilasco M, Vidalain PO, Jacob Y, et al. Polo-like kinase 1 (PLK1) regulates interferon (IFN) induction by MAVS. J Biol Chem 2009;284(33):21797–809. [PMC free article] [PubMed] [Google Scholar]
148. Xu L, Xiao N, Liu F, Ren H, Gu J. Inhibition of RIG-I and MDA5-dependent antiviral response by gC1qR at mitochondria. Proc Natl Acad Sci U S A 2009;106(5):1530–5. [PMC free article] [PubMed] [Google Scholar]
149. Koshiba T, Yasukawa K, Yanagi Y, Kawabata S. Mitochondrial membrane potential is required for MAVS-mediated antiviral signaling. Sci Signal 2011;4(158):ra7. [PubMed] [Google Scholar]
150. Tal MC, Sasai M, Lee HK, Yordy B, Shadel GS, Iwasaki A. Absence of autophagy results in reactive oxygen species-dependent amplification of RLR signaling. Proc Natl Acad Sci U S A 2009;106(8):2770–5. [PMC free article] [PubMed] [Google Scholar]
151. Soucy-Faulkner A, Mukawera E, Fink K, Martel A, Jouan L, Nzengue Y, et al. Requirement of NOX2 and reactive oxygen species for efficient RIG-I-mediated antiviral response through regulation of MAVS expression. PLoS Pathog 2010;6(6):e1000930. [PMC free article] [PubMed] [Google Scholar]
152. O’Neill LA. DNA makes RNA makes innate immunity. Cell. 2009;138(3):428–30. [PubMed] [Google Scholar]
153. Ishikawa H, Barber GN. STING is an endoplasmic reticulum adaptor that facilitates innate immune signalling. Nature. 2008;455(7213):674–8. [PMC free article] [PubMed] [Google Scholar]
154. Sun W, Li Y, Chen L, Chen H, You F, Zhou X, et al. ERIS, an endoplasmic reticulum IFN stimulator, activates innate immune signaling through dimerization. Proc Natl Acad Sci U S A 2009;106(21):8653–8. [PMC free article] [PubMed] [Google Scholar]
155. Carroll EC, Jin L, Mori A, Munoz-Wolf N, Oleszycka E, Moran HBT, et al. The Vaccine Adjuvant Chitosan Promotes Cellular Immunity via DNA Sensor cGAS-STING-Dependent Induction of Type I Interferons. Immunity. 2016;44(3):597–608. [PMC free article] [PubMed] [Google Scholar]
156. West AP, Khoury-Hanold W, Staron M, Tal MC, Pineda CM, Lang SM, et al. Mitochondrial DNA stress primes the antiviral innate immune response. Nature. 2015;520(7548):553–7. [PMC free article] [PubMed] [Google Scholar]
157. Rongvaux A, Jackson R, Harman CC, Li T, West AP, de Zoete MR, et al. Apoptotic caspases prevent the induction of type I interferons by mitochondrial DNA. Cell. 2014;159(7):1563–77. [PMC free article] [PubMed] [Google Scholar]
158. White MJ, McArthur K, Metcalf D, Lane RM, Cambier JC, Herold MJ, et al. Apoptotic caspases suppress mtDNA-induced STING-mediated type I IFN production. Cell. 2014;159(7):1549–62. [PMC free article] [PubMed] [Google Scholar]
159. Collins AC, Cai H, Li T, Franco LH, Li XD, Nair VR, et al. Cyclic GMP-AMP Synthase Is an Innate Immune DNA Sensor for Mycobacterium tuberculosis. Cell Host Microbe 2015;17(6):820–8. [PMC free article] [PubMed] [Google Scholar]
160. Storek KM, Gertsvolf NA, Ohlson MB, Monack DM. cGAS and Ifi204 cooperate to produce type I IFNs in response to Francisella infection. J Immunol 2015;194(7):3236–45. [PMC free article] [PubMed] [Google Scholar]
161. Marinho FV, Benmerzoug S, Oliveira SC, Ryffel B, Quesniaux VFJ. The Emerging Roles of STING in Bacterial Infections. Trends Microbiol 2017;25(11):906–18. [PMC free article] [PubMed] [Google Scholar]
162. Wu J, Sun L, Chen X, Du F, Shi H, Chen C, et al. Cyclic GMP-AMP is an endogenous second messenger in innate immune signaling by cytosolic DNA. Science. 2013;339(6121):826–30. [PMC free article] [PubMed] [Google Scholar]
163. Ablasser A, Goldeck M, Cavlar T, Deimling T, Witte G, Rohl I, et al. cGAS produces a 2’−5’-linked cyclic dinucleotide second messenger that activates STING. Nature. 2013;498(7454):380–4. [PMC free article] [PubMed] [Google Scholar]
164. Diner EJ, Burdette DL, Wilson SC, Monroe KM, Kellenberger CA, Hyodo M, et al. The innate immune DNA sensor cGAS produces a noncanonical cyclic dinucleotide that activates human STING. Cell Rep 2013;3(5):1355–61. [PMC free article] [PubMed] [Google Scholar]
165. Gao P, Ascano M, Wu Y, Barchet W, Gaffney BL, Zillinger T, et al. Cyclic [G(2’,5’)pA(3’,5’)p] is the metazoan second messenger produced by DNA-activated cyclic GMP-AMP synthase. Cell. 2013;153(5):1094–107. [PMC free article] [PubMed] [Google Scholar]
166. Zhang X, Shi H, Wu J, Zhang X, Sun L, Chen C, et al. Cyclic GMP-AMP containing mixed phosphodiester linkages is an endogenous high-affinity ligand for STING. Mol Cell. 2013;51(2):226–35. [PMC free article] [PubMed] [Google Scholar]
167. Sun L, Wu J, Du F, Chen X, Chen ZJ. Cyclic GMP-AMP synthase is a cytosolic DNA sensor that activates the type I interferon pathway. Science. 2013;339(6121):786–91. [PMC free article] [PubMed] [Google Scholar]
168. Gao P, Ascano M, Zillinger T, Wang W, Dai P, Serganov AA, et al. Structure-function analysis of STING activation by c[G(2’,5’)pA(3’,5’)p] and targeting by antiviral DMXAA. Cell. 2013;154(4):748–62. [PMC free article] [PubMed] [Google Scholar]
169. Saitoh T, Fujita N, Hayashi T, Takahara K, Satoh T, Lee H, et al. Atg9a controls dsDNA-driven dynamic translocation of STING and the innate immune response. Proc Natl Acad Sci U S A. 2009;106(49):20842–6. [PMC free article] [PubMed] [Google Scholar]
170. Liu S, Cai X, Wu J, Cong Q, Chen X, Li T, et al. Phosphorylation of innate immune adaptor proteins MAVS, STING, and TRIF induces IRF3 activation. Science. 2015;347(6227):aaa2630. [PubMed] [Google Scholar]
171. Tanaka Y, Chen ZJ. STING specifies IRF3 phosphorylation by TBK1 in the cytosolic DNA signaling pathway. Sci Signal. 2012;5(214):ra20. [PMC free article] [PubMed] [Google Scholar]
172. Boulay F, Tardif M, Brouchon L, Vignais P. Synthesis and use of a novel N-formyl peptide derivative to isolate a human N-formyl peptide receptor cDNA. Biochem Biophys Res Commun 1990;168(3):1103–9. [PubMed] [Google Scholar]
173. Le Y, Oppenheim JJ, Wang JM. Pleiotropic roles of formyl peptide receptors. Cytokine Growth Factor Rev 2001;12(1):91–105. [PubMed] [Google Scholar]
174. Bao L, Gerard NP, Eddy RL, Jr., Shows TB, Gerard C. Mapping of genes for the human C5a receptor (C5AR), human FMLP receptor (FPR), and two FMLP receptor homologue orphan receptors (FPRH1, FPRH2) to chromosome 19. Genomics 1992;13(2):437–40. [PubMed] [Google Scholar]
175. Zabel BA, Rott A, Butcher EC. Leukocyte chemoattractant receptors in human disease pathogenesis. Annu Rev Pathol. 2015;10:51–81. SL, Boulay I the ligand and of the receptor alters the receptor preference for neutrophil activating peptides starting with a formylmethionyl group. Biochim Biophys Acta. 2015;1853(1):192–200. [PubMed] [Google Scholar]
176. Forsman H, Winther M, Gabl M, Skovbakke SL, Boulay F, Rabiet MJ, et al. Structural changes of the ligand and of the receptor alters the receptor preference for neutrophil activating peptides starting with a formylumethionyl group. Biochim Biophy Acta. 2015; 1853(1):192–200. [PubMed] [Google Scholar]
177. He HQ, Troksa EL, Caltabiano G, Pardo L, Ye RD. Structural determinants for the interaction of formyl peptide receptor 2 with peptide ligands. J Biol Chem 2014;289(4):2295–306. [PMC free article] [PubMed] [Google Scholar]
178. Ye RD, Boulay F. Structure and function of leukocyte chemoattractant receptors. Adv Pharmacol 1997;39:221–89. [PubMed] [Google Scholar]
179. Ye RD, Boulay F, Wang JM, Dahlgren C, Gerard C, Parmentier M, et al. International Union of Basic and Clinical Pharmacology. LXXIII. Nomenclature for the formyl peptide receptor (FPR) family. Pharmacol Rev 2009;61(2):119–61. [PMC free article] [PubMed] [Google Scholar]
180. Marques PE, Amaral SS, Pires DA, Nogueira LL, Soriani FM, Lima BH, et al. Chemokines and mitochondrial products activate neutrophils to amplify organ injury during mouse acute liver failure. Hepatology. 2012;56(5):1971–82. [PubMed] [Google Scholar]
181. McDonald B, Pittman K, Menezes GB, Hirota SA, Slaba I, Waterhouse CC, et al. Intravascular danger signals guide neutrophils to sites of sterile inflammation. Science. 2010;330(6002):362–6. [PubMed] [Google Scholar]
182. O’Flaherty JT, Jacobson DP, Redman JF, Rossi AG. Translocation of protein kinase C in human polymorphonuclear neutrophils. Regulation by cytosolic Ca2(+)-independent and Ca2(+)-dependent mechanisms. J Biol Chem 1990;265(16):9146–52. [PubMed] [Google Scholar]
183. Rabiet MJ, Huet E, Boulay F. The N-formyl peptide receptors and the anaphylatoxin C5a receptors: an overview. Biochimie 2007;89(9):1089–106. [PMC free article] [PubMed] [Google Scholar]
184. Crouser ED, Shao G, Julian MW, Macre JE, Shadel GS, Tridandapani S, et al. Monocyte activation by necrotic cells is promoted by mitochondrial proteins and formyl peptide receptors. Crit Care Med 2009;37(6):2000–9. [PMC free article] [PubMed] [Google Scholar]
185. Kang HK, Lee HY, Kim MK, Park KS, Park YM, Kwak JY, et al. The synthetic peptide Trp-Lys-Tyr- Met-Val-D-Met inhibits human monocyte-derived dendritic cell maturation via formyl peptide receptor and formyl peptide receptor-like 2. J Immunol 2005;175(2):685–92. [PubMed] [Google Scholar]
186. O’Neill LA, Kishton RJ, Rathmell J. A guide to immunometabolism for immunologists. Nat Rev Immunol 2016;16(9):553–65. [PMC free article] [PubMed] [Google Scholar]
187. Milenkovic D, Blaza JN, Larsson NG, Hirst J. The Enigma of the Respiratory Chain Supercomplex. Cell Metab 2017;25(4):765–76. [PubMed] [Google Scholar]
188. Letts JA, Sazanov LA. Clarifying the supercomplex: the higher-order organization of th mitochondrial electron transport chain. Nat Struct Mol Biol 2017;24(10):800–8. [PubMed] [Google Scholar]
189. Guo R, Zong S, Wu M, Gu J, Yang M. Architecture of Human Mitochondrial Respir Megacomplex I2MI2IV2. Cell. 2017;170(6):1247–57 e12. [PubMed] [Google Scholar]
190. Garaude J, Acin-Perez R, Martinez-Cano S, Enamorado M, Ugolini M, Nistal-Villan E, et al. Mitochondrial respiratory-chain adaptations in macrophages contribute to antibacterial host defense. Nat Immunol 2016;17(9):1037–45. [PMC free article] [PubMed] [Google Scholar]
191. Mehta MM, Weinberg SE, Chandel NS. Mitochondrial control of immunity: beyond ATP. Nat Rev Immunol 2017;17(10):608–20. [PubMed] [Google Scholar]
192. O’Neill LA. A broken krebs cycle in macrophages. Immunity. 2015;42(3):393–4. [PubMed] [Google Scholar]
193. Strelko CL, Lu W, Dufort FJ, Seyfried TN, Chiles TC, Rabinowitz JD, et al. Itaconic acid is a mammalian metabolite induced during macrophage activation. J Am Chem Soc 2011;133(41):16386–9. [PMC free article] [PubMed] [Google Scholar]
194. Jha AK, Huang SC, Sergushichev A, Lampropoulou V, Ivanova Y, Loginicheva E, et al. Network integration of parallel metabolic and transcriptional data reveals metabolic modules that regulate macrophage polarization. Immunity. 2015;42(3):419–30. [PubMed] [Google Scholar]
195. Lampropoulou V, Sergushichev A, Bambouskova M, Nair S, Vincent EE, Loginicheva E, et al. Itaconate Links Inhibition of Succinate Dehydrogenase with Macrophage Metabolic Remodeling and Regulation of Inflammation. Cell Metab 2016;24(1):158–66. [PMC free article] [PubMed] [Google Scholar]
196. Naujoks J, Tabeling C, Dill BD, Hoffmann C, Brown AS, Kunze M, et al. IFNs Modify the Proteome of Legionella-Containing Vacuoles and Restrict Infection Via IRG1-Derived Itaconic Acid. PLoS Pathog 2016;12(2):e1005408. [PMC free article] [PubMed] [Google Scholar]
197. Michelucci A, Cordes T, Ghelfi J, Pailot A, Reiling N, Goldmann O, et al. Immune-responsive gene 1 protein links metabolism to immunity by catalyzing itaconic acid production. Proc Natl Acad Sci U S A 2013;110(19):7820–5. [PMC free article] [PubMed] [Google Scholar]
198. Mills EL, Kelly B, Logan A, Costa ASH, Varma M, Bryant CE, et al. Succinate Dehydrogenase Supports Metabolic Repurposing of Mitochondria to Drive Inflammatory Macrophages. Cell. 2016;167(2):457–70 e13. [PMC free article] [PubMed] [Google Scholar]
199. Tannahill GM, Curtis AM, Adamik J, Palsson-McDermott EM, McGettrick AF, Goel G, et al. Succinate is an inflammatory signal that induces IL-1beta through HIF-1alpha. Nature. 2013;496(7444):238–42. [PMC free article] [PubMed] [Google Scholar]
200. Anthony RM, Urban JF Jr., Alem F, Hamed HA, Rozo CT, Boucher JL, et al. Memory T(H)2 cells induce alternatively activated macrophages to mediate protection against nematode parasites. Nat Med 2006;12(8):955–60. [PMC free article] [PubMed] [Google Scholar]
201. Vats D, Mukundan L, Odegaard JI, Zhang L, Smith KL, Morel CR, et al. Oxidative metabolism and PGC-1beta attenuate macrophage-mediated inflammation. Cell Metab 2006;4(1):13–24. [PMC free article] [PubMed] [Google Scholar]
202. Rodriguez-Prados JC, Traves PG, Cuenca J, Rico D, Aragones J, Martin-Sanz P, et al. Substrate fate in activated macrophages: a comparison between innate, classic, and alternative activation. J Immunol 2010;185(1):605–14. [PubMed] [Google Scholar]
203. Wu D, Sanin DE, Everts B, Chen Q, Qiu J, Buck MD, et al. Type 1 Interferons Induce Changes in Core Metabolism that Are Critical for Immune Function. Immunity. 2016;44(6):1325–36. [PMC free article] [PubMed] [Google Scholar]
204. Lachmandas E, Boutens L, Ratter JM, Hijmans A, Hooiveld GJ, Joosten LA, et al. Microbial stimulation of different Toll-like receptor signalling pathways induces diverse metabolic programmes in human monocytes. Nat Microbiol 2016; 2:16246. [PubMed] [Google Scholar]
-