Mitochondria have long been appreciated as the metabolic hub of cells. Emerging evidence also posits these organelles as hubs for innate immune signalling and activation, particularly in macrophages. Macrophages are front-line cellular defenders against endogenous and exogenous threats in mammals. These cells use an array of receptors and downstream signalling molecules to respond to a diverse range of stimuli, with mitochondrial biology implicated in many of these responses. Mitochondria have the capacity to both divide through mitochondrial fission and coalesce through mitochondrial fusion. Mitochondrial dynamics, the balance between fission and fusion, regulate many cellular functions, including innate immune pathways in macrophages. In these cells, mitochondrial fission has primarily been associated with pro-inflammatory responses and metabolic adaptation, so can be considered as a combative strategy utilised by immune cells. In contrast, mitochondrial fusion has a more protective role in limiting cell death under conditions of nutrient starvation. Hence, fusion can be viewed as a cellular survival strategy. Here we broadly review the role of mitochondria in macrophage functions, with a focus on how regulated mitochondrial dynamics control different functional responses in these cells.

Macrophages are innate immune cells with central roles in host defence in mammals. These cells constantly survey their surroundings, using pattern recognition receptors (PRRs) and other detection systems to sense and respond to indicators of danger, for example, infection or injury [1,2]. This results in the engagement of antimicrobial defence systems, coordination of inflammatory responses, priming of adaptive immunity, and initiation of repair processes. Macrophage-expressed PRRs recognise both exogenous pathogen-associated molecular patterns (PAMPs) such as components of microorganisms, as well as endogenous danger signals such as products released from dead or dying cells, tumour cells, and certain mitochondrial components that are collectively referred to as danger-associated molecular patterns (DAMPs). The innate immune system is equipped with diverse families of PRRs, including the toll-like receptors (TLRs), C-type lectin receptors, retinoic acid-inducible gene 1 (RIG-1)-like helicase receptors (RLRs), and nucleotide-binding oligomerization domain-like receptors (NLRs), with each family being comprised of several different receptors [3]. Despite this diversity, there is often overlap in the downstream biological responses that are generated upon sensing PAMPs and/or DAMPs. This may partly reflect the involvement of the mitochondrion, a key organelle integrating extracellular signals, cell metabolism, and biological outputs in macrophages.

The conception of mitochondria as a signalling organelle began with the discovery that the release of cytochrome c from mitochondria initiates a signalling cascade that leads to apoptotic cell death [4,5]. Since then, a vast literature has revealed that mitochondria have central roles in cell activation, cell survival, and many forms of cell death, with these organelles profoundly influencing numerous biological processes [6,7]. This has been extensively studied in immune responses, where mitochondria regulate both host defence [8] and sterile inflammation [9]. Mitochondria are dynamic organelles that can exist within a spectrum of morphological states within cells. This is governed by the cellular processes of mitochondrial fission and fusion, with the balance between fission and fusion often referred to as mitochondrial dynamics. Mitochondrial dynamics control many cellular pathways, including metabolism [10,11] and inflammatory responses [12–15]. In this review, we briefly describe the role of mitochondria in innate immunity, before focusing on how mitochondrial dynamics influence the metabolic status of macrophages, as well as the functional responses of these cells.

Mitochondria are double membrane energy-generating organelles. The endosymbiont theory of mitochondrial origin proposes that a free-living α-proteobacterium was engulfed by an eukaryotic precursor cell ∼2 billion years ago, resulting in a mutually beneficial relationship [16]. During evolution, mitochondria lost most of the proteobacterial genomic materials and transferred many genes to the nuclear genome via endosymbiotic gene transfer [17]. Thus, most mitochondrial components are encoded by the nuclear genome. The small circular mitochondrial genome (mtDNA) mostly encodes translation machinery and components of respiratory chain complexes I, III, IV, and V for carrying out the key mitochondrial function of oxidative phosphorylation (OXPHOS), via the co-ordinated actions of the tricarboxylic acid (TCA) cycle and the electron transport chain (ETC). ETC complexes I, II, and III also generate mitochondrial reactive oxygen species (mROS) which contribute to various functions in innate immunity (see ahead).

Beyond their roles in energy generation, mitochondria control diverse cellular processes. In innate immune cells, mitochondria serve as signalling platforms for some PRR pathways, control PRR-inducible metabolic reprogramming, generate free radicals and metabolites that contribute to host defence and inflammation, and provide a reservoir of DAMPs for cellular activation upon disruption of homeostasis (Figure 1). Below we briefly describe examples of each of these.

The multifaceted roles of mitochondria in innate immunity.

Figure 1.
The multifaceted roles of mitochondria in innate immunity.

Mitochondria have diverse functions in innate immune cells, including: (1) cell signalling, as exemplified by RLR-mediated engagement of MAVS for antiviral gene expression and TLR-inducible activation of ECSIT via TRAF6, as well as mROS and mitochondria-derived metabolites acting as signalling molecules; (2) metabolic reprogramming, as is apparent during TLR activation in which there is a metabolic shift from OXPHOS to glycolysis, as well as increased production of succinate, itaconate, fumarate, and mROS, all of which have inflammatory and/or antimicrobial roles; (3) generation of antimicrobial responses, with the antimicrobial effector molecule mROS and antibacterial metabolites all being produced downstream of TLR activation; and (4) DAMP-mediated cellular activation, in which mitochondrial DAMPs, such as TFAM, n-FP, and mtDNA, can all trigger innate immune activation. mtDNA, mitochondrial DNA; n-FP, N-formyl peptides; TFAM, mitochondrial transcription factor A. Created with BioRender.com.

Figure 1.
The multifaceted roles of mitochondria in innate immunity.

Mitochondria have diverse functions in innate immune cells, including: (1) cell signalling, as exemplified by RLR-mediated engagement of MAVS for antiviral gene expression and TLR-inducible activation of ECSIT via TRAF6, as well as mROS and mitochondria-derived metabolites acting as signalling molecules; (2) metabolic reprogramming, as is apparent during TLR activation in which there is a metabolic shift from OXPHOS to glycolysis, as well as increased production of succinate, itaconate, fumarate, and mROS, all of which have inflammatory and/or antimicrobial roles; (3) generation of antimicrobial responses, with the antimicrobial effector molecule mROS and antibacterial metabolites all being produced downstream of TLR activation; and (4) DAMP-mediated cellular activation, in which mitochondrial DAMPs, such as TFAM, n-FP, and mtDNA, can all trigger innate immune activation. mtDNA, mitochondrial DNA; n-FP, N-formyl peptides; TFAM, mitochondrial transcription factor A. Created with BioRender.com.

Close modal

Mitochondria are intimately connected to many innate immune signalling pathways. One of the most intensely studied examples of this involves the adaptor protein mitochondrial antiviral signalling protein (MAVS) that initiates antiviral responses upon RLR-mediated sensing of cytosolic viral RNA. MAVS is positioned at the mitochondrial outer membrane (OMM) where it forms complexes with the RLRs RIG-I [18] and MDA5 [19] upon activation by viral RNA. This interaction consequently triggers antiviral responses via the transcription factors interferon (IFN) regulatory factor (IRF) 3, IRF7, and nuclear factor-κB (NF-κB), leading to inducible expression of type I IFNs and other antiviral genes [20,21].

The burgeoning field of immunometabolism encompasses the role of mitochondria-regulated metabolism and metabolites in modulating the immune functions of cells, such as macrophages. Mitochondria-mediated metabolic changes alter macrophage functions, particularly their inflammatory and antimicrobial status. In response to lipopolysaccharide (LPS) and other inflammatory stimuli, cells rewire their metabolism from OXPHOS towards aerobic glycolysis, leading to a metabolic shift. For instance, LPS-inducible TLR4-activation redirects metabolic fluxes to generate acetyl-coenzyme A from glucose and increases ATP-citrate lyase activity, thus facilitating inducible histone acetylation in macrophages [22]. Moreover, Jha et al. [23] showed that the metabolites succinate and itaconate accumulate in activated macrophages due to a TLR-inducible break in the TCA cycle. Intriguingly, several studies have revealed that these metabolites have immunomodulatory and/or antimicrobial properties [24–28], though further studies are required to understand the in vivo relevance of some of these effects. One possible mechanism underlying metabolic reprogramming could be the translocation of TLR signalling molecules such as ECSIT [29] and STAT3 [30] to mitochondria in macrophages.

Under steady-state conditions, the mitochondrial ETC generates a small amount of ROS; however, this is amplified during cell stress and/or during metabolic adaptations. In macrophages, for example, LPS-inducible metabolic reprogramming leads to succinate accumulation that drives mROS production [25]. The increased mROS can activate pro-inflammatory signalling pathways [31,32], with this linked to many inflammatory conditions, for example, chronic obstructive pulmonary disease [33], chronic kidney disease [34], and type-1 diabetes-associated vascular inflammation [35]. Furthermore, TLR-inducible mROS also contributes to macrophage antibacterial responses [29,36]. Such studies have established mROS as an effector molecule of innate immunity.

Owing to their bacterial origin, mitochondria contain DAMPs, such as mtDNA, N-formyl peptides (n-FP), and mitochondrial transcription factor A (TFAM). Release of mitochondrial contents from damaged or necrotic cells can thus initiate sterile inflammation. For example, the concomitant release of n-FP and TFAM from necrotic cells activates monocytes [37] and promotes immune cell recruitment [38], while circulating mtDNA can trigger TLR9-mediated inflammatory responses [39,40] in cardiovascular-related conditions [41,42]. In this way, mitochondrial components can drive innate immune inflammatory responses.

Mitochondria are dynamic organelles that exist in a continuum of states ranging from long filamentous to small spherical structures. The opposing processes of mitochondrial fission and fusion, referred to as mitochondrial dynamics, co-ordinate, and determine the overall mitochondrial morphology in a cell at any given time [43]. Mitochondrial dynamics play a vital role in mitochondrial quality control, cell division, and cellular stress responses. Underlying the importance of this process, the genetic deletion of essential regulators of mitochondrial dynamics results in embryonic lethality in mice [44,45]. For example, mice defective in genes required for mitochondrial fusion die in mid-gestation [44], while Wakabayashi et al. [45] demonstrated genetically that fission is essential for mouse embryonic and brain development, as well as mitochondrial morphogenesis, mitotic division, and cell death. In a healthy undisturbed cell, the balance in mitochondrial dynamics is generally skewed more towards a fused interconnected network of mitochondria, although fragmented spherical mitochondria are also normally present. When nutrients are limiting, the mitochondrial pool becomes hyperfused to enable functional cooperativity between mitochondria and cellular protection [46,47]. Conversely, excess nutrients and other stress signals lead to a hyperfragmented mitochondrial population, with fission exceeding fusion. This can have various functional consequences, including initiating apoptosis [48], aiding in metabolic adaptations [49], and regulating energy expenditure [50].

Cells use a specialised set of mechanical GTPases to control mitochondrial dynamics. One such GTPase, dynamin-related protein 1 (DRP1), encoded by DNM1L, is essential for mitochondrial fission [51,52]. DRP1 is a cytosolic protein that localises to mitochondria, forming an oligomeric complex upon activation. The act of fission occurs in two sequential steps. First, the endoplasmic reticulum (ER) and actin collaborate to mark a scission site where DRP1 assembles on the OMM. Next, DRP1 monomers form a large oligomer encircling this site, with the GTPase activity of DRP1 then facilitating membrane scission [53–56]. A recent study showed that the ER transmembrane protein CTRP1 directly interacts with DRP1 and facilitates its recruitment to mitochondria, suggesting a mechanism of ER–mitochondrial interaction during the initial stages of fission [57]. Several OMM-localised adaptor proteins have also been implicated in regulating DRP1-dependent fission. These include mitochondrial fission factor (MFF), mitochondrial dynamics of 51 kDa protein (MiD51), MiD49, and mitochondrial fission protein 1 (FIS1) [58–61]. DRP1 can bind to each of these adaptor proteins on the OMM, with the exact mechanisms by which they act being an intense area of current investigation.

MFF can directly bind to DRP1 to facilitate its recruitment, with the absence of MFF in HeLa cells skewing cells towards fusion [60,62]. There are contrasting studies on MiD49- and MiD51-mediated control of mitochondrial dynamics, with evidence that they promote both fission and fusion in different cell types [61,63,64]. Similarly, there may be context-dependent roles for FIS1 in mitochondrial fission. Zhang et al. [65] showed that FIS1 competitively binds to MiD51, suppressing its inhibitory effect on DRP1 to promote mitochondrial fission in a human lung-adenocarcinoma cell line. In contrast, Otera et al. [62] reported that FIS1 was dispensable for fission in HeLa cells. Kleele et al. [66] recently provided key insights into how different adaptor proteins regulate mitochondrial fission in different contexts to enable distinct functional outputs. Specifically, two distinct forms of DRP1-dependent fission were reported, one occurring at the periphery and another at the midzone of mitochondria. Peripheral fission occurs during mitochondrial stress and requires the establishment of FIS1-mediated lysosomal–mitochondrial contact sites. In contrast, midzone fission occurs during mitochondrial proliferation and requires MFF, along with ER-and actin-mediated pre-constriction of mitochondria. In this way, different DRP1 adaptor proteins can engage fission for distinct biological responses, namely quality control of mitochondria and cell division.

Both recruitment of DRP1 to mitochondria, along with its activation, are controlled by several post-translational modifications (PTMs). These include phosphorylation, S-nitrosylation, sumoylation, acetylation, and ubiquitination of specific residues. The contributions of specific PTMs on DRP1 to its activation and functional responses in different cell types are summarised in Table 1. This summary table highlights the diversity in DRP1 PTM sites, as well as in the enzymes involved in mediating these effects in different cell types. It is likely that different PTMs on DRP1 may influence its interactions with different adaptors for initiating or constraining fission, an area of investigation that is still evolving. For example, UV-stimulation of human lung-adenocarcinoma cells decreased phosphorylation of DRP1 at serine (S) 637, thus promoting a DRP1-MFF interaction and enhancing fission during apoptosis [65].

Table 1.
PTM sites on DRP1, along with mechanisms involved (serine, S; alanine, A; threonine, T; cysteine, C; lysine, K; aspartic acid, D; glutamic acid, E; arginine, R)
Type of PTMPTM siteResponsible enzymeEffect on DRP1 activitySpecific DRP1 point mutations assessedCell typeReferences
Phosphorylation S616 CDK1/cyclin Activation S to A HeLa cells, human liver cells [125,126
PKCδ Activation — Mouse cardiomyocytes [127
ERK2 (also known as MAPK1) Activation S to A HEK-TtH cells [128
S to A Huntington's disease mouse striatal cells [129
PINK1 Activation S to A
S to D 
HEK293 cells [130
S to A
S to D 
Mouse primary neurons [131
CDK5 Activation S to A
S to E 
Glioblastoma cells [132,133
Inhibition S to A
S to D 
Mouse primary neurons [134
Ca2+/calmodulin-dependent kinase II (CaMKII) Activation S to A Rat cardiomyocytes [135
S412 S684 TBK1 Inhibition S to A
S to D 
HEK293T cells [136
S637 PKA Inhibition S to A
S to D 
Rat PC12 cells, African green monkey kidney fibroblast cells [137
CaMKIa Inhibition S to A
S to D 
Rat primary neurons, HeLa cells [138
T595 LRRK2 Activation T to A
T to D 
HeLa cells, HEK293T cells [139
Dephosphorylation S637 Calcineurin (also known as PP2B) Activation S to A
S to D 
Rat PC12 cells, African green monkey kidney fibroblast cells [137
Neuron-specific PP2A/Bβ2 phosphatase Activation — Mouse hippocampal neurons [140
S-nitrosylation C644 Redox-mediated catalysis (donor is nitric oxide) Activation C to A Mouse cerebrocortical neurons [141
Not indicated Protein disulphide isomerase Facilitates DRP1 S616 phosphorylation and activation — Mouse hippocampal neurons [142
Sumoylation Not indicated SUMO E3 ligase, MAPL Activation — HeLa cells [143
De-sumoylation Not indicated SENP5 Inhibition — COS-7 murine fibroblast like cells [144
K557, K560, K569 or K571 SENP3 Activation K557, 560, 569, and 571 to R Mouse primary cortical neurons [145
Enhanced DRP1-MFF binding K557, 560, 569, and 571 to R HEK293 cells [146
Ubiquitination Not indicated E3 ubiquitin ligase, MARCH Activation — HeLa cells [147
Inhibition — COS-7 murine fibroblast like cells, HeLa cells [148,149
Acetylation K642 Not identified yet Activation K to R Mouse cardiomyocytes [150
Type of PTMPTM siteResponsible enzymeEffect on DRP1 activitySpecific DRP1 point mutations assessedCell typeReferences
Phosphorylation S616 CDK1/cyclin Activation S to A HeLa cells, human liver cells [125,126
PKCδ Activation — Mouse cardiomyocytes [127
ERK2 (also known as MAPK1) Activation S to A HEK-TtH cells [128
S to A Huntington's disease mouse striatal cells [129
PINK1 Activation S to A
S to D 
HEK293 cells [130
S to A
S to D 
Mouse primary neurons [131
CDK5 Activation S to A
S to E 
Glioblastoma cells [132,133
Inhibition S to A
S to D 
Mouse primary neurons [134
Ca2+/calmodulin-dependent kinase II (CaMKII) Activation S to A Rat cardiomyocytes [135
S412 S684 TBK1 Inhibition S to A
S to D 
HEK293T cells [136
S637 PKA Inhibition S to A
S to D 
Rat PC12 cells, African green monkey kidney fibroblast cells [137
CaMKIa Inhibition S to A
S to D 
Rat primary neurons, HeLa cells [138
T595 LRRK2 Activation T to A
T to D 
HeLa cells, HEK293T cells [139
Dephosphorylation S637 Calcineurin (also known as PP2B) Activation S to A
S to D 
Rat PC12 cells, African green monkey kidney fibroblast cells [137
Neuron-specific PP2A/Bβ2 phosphatase Activation — Mouse hippocampal neurons [140
S-nitrosylation C644 Redox-mediated catalysis (donor is nitric oxide) Activation C to A Mouse cerebrocortical neurons [141
Not indicated Protein disulphide isomerase Facilitates DRP1 S616 phosphorylation and activation — Mouse hippocampal neurons [142
Sumoylation Not indicated SUMO E3 ligase, MAPL Activation — HeLa cells [143
De-sumoylation Not indicated SENP5 Inhibition — COS-7 murine fibroblast like cells [144
K557, K560, K569 or K571 SENP3 Activation K557, 560, 569, and 571 to R Mouse primary cortical neurons [145
Enhanced DRP1-MFF binding K557, 560, 569, and 571 to R HEK293 cells [146
Ubiquitination Not indicated E3 ubiquitin ligase, MARCH Activation — HeLa cells [147
Inhibition — COS-7 murine fibroblast like cells, HeLa cells [148,149
Acetylation K642 Not identified yet Activation K to R Mouse cardiomyocytes [150

In comparison with fission, fusion requires more stringent regulation by multiple GTPases, both at the OMM and the IMM [67]. The IMM lipid cardiolipin interacts with the GTPase optic atrophy 1 (OPA1) to promote its GTPase activity [68,69], enabling it to initiate IMM fusion. In contrast, the GTPases mitofusin 1 (MFN1) and MFN2 drive OMM fusion [44,68,70]. In addition, two OMM proteins, FAM73a and FAM73b, facilitate fusion downstream of MFNs via the mitochondrial phospholipase D [71]. The fusion-promoting GTPases are also regulated via distinct PTMs. For example, MFN1 and OPA1 deacetylation by the lysine deacetylases HDAC6 [72] and sirtuin 3 [73], respectively, activate these GTPases to promote mitochondrial fusion. In contrast, MFN1 phosphorylation results in its ubiquitin-mediated proteasomal degradation, thus inhibiting fusion [74]. Given the diverse regulatory mechanisms that control each GTPase involved in fission and fusion, it is evident that complex mechanisms connect cell signalling to mitochondrial dynamics, with much yet to be understood about how mitochondrial dynamics are regulated.

Several innate immune stimuli and pathogens modulate and/or disrupt mitochondrial dynamics (Table 2), with this having many consequences for cellular functions (Figure 2). Given the range of stimuli that can affect fission and fusion, it seems likely that multiple PRRs and PRR signalling pathways may converge to modulate mitochondrial dynamics. The consequences of this modulation on macrophage metabolism, inflammatory outputs, phagocytosis, and the host–pathogen dynamic, are discussed below.

Modulation of mitochondrial dynamics by inflammatory stimuli and infectious agents.

Figure 2.
Modulation of mitochondrial dynamics by inflammatory stimuli and infectious agents.

Pathogens and inflammatory stimuli can regulate mitochondrial dynamics, driving either mitochondrial fusion or fission depending on the pathogen/stimulus and cellular context. Specific examples of viruses, bacteria, and inflammatory stimuli that drive either mitochondrial fusion or fission are shown. These can affect mitochondrial dynamics through a variety of mechanisms, with modulation of DRP1 being common to many stimuli/pathogens (with the exception of IL-4, which skews towards fusion via the mitochondrial outer membrane protein FAM73b). Pathogen-driven manipulation of mitochondrial dynamics to either fusion or fission can favour pathogen persistence, replication, and/or survival, depending on the nature of the pathogen and cellular context. Red arrows indicate inhibition, green arrows indicate activation, dotted green arrows indicate positive effect on either fusion or fission. OCR, oxygen consumption rate; OXPHOS, oxidative phosphorylation; SRC, spare respiratory capacity. Created with BioRender.com.

Figure 2.
Modulation of mitochondrial dynamics by inflammatory stimuli and infectious agents.

Pathogens and inflammatory stimuli can regulate mitochondrial dynamics, driving either mitochondrial fusion or fission depending on the pathogen/stimulus and cellular context. Specific examples of viruses, bacteria, and inflammatory stimuli that drive either mitochondrial fusion or fission are shown. These can affect mitochondrial dynamics through a variety of mechanisms, with modulation of DRP1 being common to many stimuli/pathogens (with the exception of IL-4, which skews towards fusion via the mitochondrial outer membrane protein FAM73b). Pathogen-driven manipulation of mitochondrial dynamics to either fusion or fission can favour pathogen persistence, replication, and/or survival, depending on the nature of the pathogen and cellular context. Red arrows indicate inhibition, green arrows indicate activation, dotted green arrows indicate positive effect on either fusion or fission. OCR, oxygen consumption rate; OXPHOS, oxidative phosphorylation; SRC, spare respiratory capacity. Created with BioRender.com.

Close modal
Table 2.
Modulation of mitochondrial dynamics by innate immune stimuli
StimuliCell typeEffect on fission or fusionFunctional consequencesEvidenceReferences
Extracellular signals 
LPS Murine macrophages.
Murine microglial cells 
Fission  Inflammatory cytokines  Drp1 silencing or treatment with Mdivi1 [10,12,81
Succinate Rat cardiomyocytes Fission  Cell apoptosis, myocardial ischaemia injury DRP1 recruitment to mitochondria and activation of MFF [127
IL-4 Murine macrophages Fusion  OXPHOS Mitochondrial morphology, MFN1 and MFN2 ↑ [81
TNF H9C2 cardiomyocytes Fission  Cell death during sepsis  Inhibition of DRP1 by Rho-associated kinases inhibitor [151
Poly(I:C) HEK293T cells Fusion  Cell survival TBK1 inhibits mitochondrial aggregation of DRP1 [136
Bacterial infections 
Shigella flexneri HeLa cells Fission  Cell death
Cell-to-cell spreading  
DRP1 silencing or treatment with Mdivi1 [152
Legionella pneumophila Human macrophages Fission  Glycolysis
Bacterial survival  
DRP1 inhibition with Mdivi1 [112
Chlamydia trachomatis HUVECS, HeLa cells Fusion  OXPHOS
Bacterial survival  
DRP1 levels  [153
Vibrio cholerae HEK cells, CHO cells, HeLa cells Fission  Host inflammatory responses  Bacterial VopE interacts with Miro GTPases at mitochondria [154
Listeria monocytogenes HeLa cells Fission  ATP production
Bacterial survival  
Genetic silencing of DRP1, MFN1 and MFN2 [111,155
Helicobacter pylori Human epithelial AZ-521 cells Fission  Cell apoptosis  DRP1 inhibition with Mdivi1 [156
Viral Infections 
Dengue virus Human hepatoma 7 cells Fusion  Viral replication  DRP1 expression  [157,158
Sendai virus HEK293T cells, HeLa cells Fusion  Viral persistence, virus detection and signalling  DRP1, FIS1, OPA1 and MFN1 silencing [159
Venezuelan equine encephalitis virus U87MG (human glioblastoma cell line) Fission  Mitophagy, autophagy and cell death  Inhibition of fission with Mdivi1 [160
Epstein–Barr virus Gastric and breast cancer cells Fission  Cell apoptosis and migration  DRP1 levels  [161
SARS coronavirus Pulmonary epithelial cells, HEK cells, THP-1 cells Fusion  Innate immune signalling
Viral persistence  
DRP1 levels  [162
Influenza A virus HEK cells, murine macrophages Fission  Antiviral response  Influenza A viral protein PB1-F2 localises to mitochondria [163,164
Hepatitis B virus Human hepatoma 7 cells Fission  Mitophagy
Apoptosis
Viral persistence  
DRP1 S616 phosphorylation, MFN2 ubiquitination and degradation [165
Hepatitis C virus Human hepatoma 7 cells Fission  Viral persistence
Apoptosis 
DRP1 S616 phosphorylation and translocation to mitochondria [166
StimuliCell typeEffect on fission or fusionFunctional consequencesEvidenceReferences
Extracellular signals 
LPS Murine macrophages.
Murine microglial cells 
Fission  Inflammatory cytokines  Drp1 silencing or treatment with Mdivi1 [10,12,81
Succinate Rat cardiomyocytes Fission  Cell apoptosis, myocardial ischaemia injury DRP1 recruitment to mitochondria and activation of MFF [127
IL-4 Murine macrophages Fusion  OXPHOS Mitochondrial morphology, MFN1 and MFN2 ↑ [81
TNF H9C2 cardiomyocytes Fission  Cell death during sepsis  Inhibition of DRP1 by Rho-associated kinases inhibitor [151
Poly(I:C) HEK293T cells Fusion  Cell survival TBK1 inhibits mitochondrial aggregation of DRP1 [136
Bacterial infections 
Shigella flexneri HeLa cells Fission  Cell death
Cell-to-cell spreading  
DRP1 silencing or treatment with Mdivi1 [152
Legionella pneumophila Human macrophages Fission  Glycolysis
Bacterial survival  
DRP1 inhibition with Mdivi1 [112
Chlamydia trachomatis HUVECS, HeLa cells Fusion  OXPHOS
Bacterial survival  
DRP1 levels  [153
Vibrio cholerae HEK cells, CHO cells, HeLa cells Fission  Host inflammatory responses  Bacterial VopE interacts with Miro GTPases at mitochondria [154
Listeria monocytogenes HeLa cells Fission  ATP production
Bacterial survival  
Genetic silencing of DRP1, MFN1 and MFN2 [111,155
Helicobacter pylori Human epithelial AZ-521 cells Fission  Cell apoptosis  DRP1 inhibition with Mdivi1 [156
Viral Infections 
Dengue virus Human hepatoma 7 cells Fusion  Viral replication  DRP1 expression  [157,158
Sendai virus HEK293T cells, HeLa cells Fusion  Viral persistence, virus detection and signalling  DRP1, FIS1, OPA1 and MFN1 silencing [159
Venezuelan equine encephalitis virus U87MG (human glioblastoma cell line) Fission  Mitophagy, autophagy and cell death  Inhibition of fission with Mdivi1 [160
Epstein–Barr virus Gastric and breast cancer cells Fission  Cell apoptosis and migration  DRP1 levels  [161
SARS coronavirus Pulmonary epithelial cells, HEK cells, THP-1 cells Fusion  Innate immune signalling
Viral persistence  
DRP1 levels  [162
Influenza A virus HEK cells, murine macrophages Fission  Antiviral response  Influenza A viral protein PB1-F2 localises to mitochondria [163,164
Hepatitis B virus Human hepatoma 7 cells Fission  Mitophagy
Apoptosis
Viral persistence  
DRP1 S616 phosphorylation, MFN2 ubiquitination and degradation [165
Hepatitis C virus Human hepatoma 7 cells Fission  Viral persistence
Apoptosis 
DRP1 S616 phosphorylation and translocation to mitochondria [166

Alterations in mitochondrial dynamics are interwoven with changes in the metabolic phenotype of a cell. When fusion is favoured over fission, cells generally occupy a catabolic state and generate ATP through OXPHOS [49]. In fibroblasts, fusion was shown to have a causative role in promoting OXPHOS, with this required for cell proliferation [75]. In contrast, a hyperfragmented mitochondrial pool portrays an anabolic state and a shift towards aerobic glycolysis [76]. In cancer cells, one of the rate-limiting enzymes of glycolysis, pyruvate kinase isoform M2, directly binds to MFN2. This interaction results in augmented mitochondrial fusion and a subsequent metabolic shift towards OXPHOS, in this case leading to suppression of cancer cell growth [77]. On the contrary, Nair et al. [10] showed that LPS induces mitochondrial fission and skews metabolism from OXPHOS to glycolysis in primary microglia. They demonstrated that pharmacological inhibition of mitochondrial fission with Mdivi1 [78] reversed this metabolic reprogramming and attenuated LPS-induced pro-inflammatory cytokine and chemokine production in these cells. Similarly, Zhang et al. [11] showed that genetic silencing of DRP1 inhibited LPS-inducible glycolysis in airway smooth muscle cells, as well as cell proliferation. Thus, growing evidence connects TLR-inducible mitochondrial fission to metabolic reprogramming.

As discussed above, mROS and mitochondrial metabolites regulate macrophage inflammatory responses. Given the intricate link between mitochondrial dynamics and metabolism, current research in this area is dissecting the role of mitochondrial dynamics in inflammation. Most studies in this area have primarily focused on neuroinflammation [13] and neurodegenerative diseases [79] (see ahead). However, several in vitro studies using the primary mouse or human macrophages have investigated specific molecular pathways and inflammatory outputs. For example, LPS triggered mitochondrial fission in both primary human and mouse macrophages, with genetic or pharmacological targeting of DRP1 in mouse macrophages and embryonic fibroblasts inhibiting the LPS-inducible production of a subset inflammatory mediators including IL-12p40, IL-6, and TNF [12]. Gao et al. [80] also established that LPS- or Staphylococcus aureus-mediated activation of DRP1 in mouse macrophages facilitated the production of the pro-inflammatory cytokine TNF. Furthermore, depletion of the fusion-promoting protein FAM73b skewed towards fission, impaired OXPHOS and promoted specific TLR-induced pro-inflammatory responses in murine macrophages and dendritic cells [81]. This resulted in increased Il12a expression, as well as decreased Il10 and Il23a expression, enhancing macrophage-mediated anti-tumour immune responses [81]. Similarly, genetic silencing of MFN2 in primary human macrophages enhanced TLR2-mediated pro-inflammatory outputs [82]. However, MFN2-silenced cells showed only a mild mitochondrial fragmentation, with this attributed to compensatory expression of MFN1 in the absence of MFN2. In contrast, Tur et al. demonstrated that Mfn2-deficient mouse macrophages were defective in LPS-inducible production of pro-inflammatory cytokines and nitric oxide (NO). However, they did not ascribe this phenotype to defective mitochondrial fusion, rather reduced ROS production. Interestingly, MFN2 was also shown to be essential for inflammasome activation upon RNA virus infection in mouse macrophages [83], suggestive of a role for mitochondrial fusion in this PRR pathway. In line with this, skewing towards fusion by silencing Drp1 in murine macrophages increased ERK signalling, leading to subsequent activation of the NLRP3 inflammasome pathway and IL-1β release [84]. These studies on MFN2 are suggestive of pro-inflammatory functions for fusion, contrasting with the general view that fission and fusion are linked to pro- and anti-inflammatory responses, respectively. However, it is also possible that MFN2 may have an additional mitochondrial fusion-independent function that may account for these phenotypes. Overall, a growing body of literature has demonstrated that TLR agonists and other inflammatory stimuli alter mitochondrial dynamics (Table 2), with consequent initiation of specific inflammatory responses in macrophages. It should be noted that much of the existing literature on TLR-regulated mitochondrial dynamics has focused on TLR4, however, with additional studies now being required to ascertain whether other TLRs influence this cellular process and downstream biological effects.

As noted above, much of the literature on mitochondrial dynamics and inflammation has focused on neuroinflammation, particularly with respect to microglia. These tissue-resident macrophages of the central nervous system regulate neuronal survival [85], tissue-repair [86], and immunity [87]. However, during infection or injury, microglia may adopt a pro-inflammatory phenotype, releasing cytokines, ROS, and NO [88]. Sustained and chronic release of these inflammatory mediators in the central nervous system is neurotoxic, and may promote neuronal damage [89]. For example, activated microglia are associated with initiating pro-inflammatory signalling to promote neuronal damage in several neurodegenerative diseases, including Parkinson's disease (PD) [90–92] and Alzheimer's disease [93]. Mounting evidence implicates pro-inflammatory microglia in neuroinflammation and neurodegenerative pathology.

Exactly how microglia drive neuroinflammation remains elusive, but several lines of evidence support a role for an axis involving TLR4 and mitochondrial fission. Intraperitoneal injection of LPS initiated microglial activation, as well as dopaminergic neuron degeneration in mice [94,95]. This suggests that microglial TLR4-mediated pro-inflammatory pathways can drive neurodegeneration. Several studies also showed that LPS drives fission in microglia, with this linked to increased ROS, NO, and pro-inflammatory cytokines (IL-1β, IL-6, TNF) [14,96,97]. Furthermore, inhibition of DRP1 function dampened inducible LPS-induced mRNA expression of Il1b, Il6, and Tnf, as well as intracellular ROS production, in a mouse microglial cell line [98]. Metabolic reprogramming from oxidative phosphorylation to glycolysis is required for microglia to adopt a pro-inflammatory phenotype [99], and as noted above, mitochondrial fission was required for this metabolic switch in microglia [10]. These data thus suggest that TLR4-mediated mitochondrial fission may enhance pro-inflammatory phenotypes in microglia. Interestingly, increased mitochondrial fission has also been observed in pro-inflammatory astrocytes in vitro [15], suggesting a conserved role for mitochondrial fission across multiple cell types during neuroinflammation.

Another possible mechanism of mitochondrial fission perpetuating neuroinflammation is via enhanced microglial NLRP3 signalling. It is established that mitochondrial dysfunction primes and/or engages the NLRP3 inflammasome. For example, mROS and mtDNA trigger assembly and activation of the cytosolic NLRP3 inflammasome, as well as pro-inflammatory responses via IL-1β release and cell death [100–102]. In mouse macrophages, skewing towards fusion suppressed the release of the inflammasome-dependent cytokine IL-1β [12]. Furthermore, antagonising mitochondrial fission in PD models reduced brain tissue expression of NLRP3 and NLRP3 signalling components, which were otherwise elevated in the brain tissue of rats with a PD-like phenotype [103]. Similarly, intraperitoneal administration of the fission-inhibiting compound Mdivi1 in an acute kidney injury model in mice significantly down-regulated the expression of NLRP3 and inflammasome-related proteins in kidney tissue [104]. This suggests that mitochondrial fission may contribute to the priming of inflammasome responses during neuroinflammation, as well as other inflammatory conditions. Moreover, the administration of mitochondrial fission inhibitors in vivo was neuroprotective in several animal models of neurodegenerative disease. For example, intraperitoneal injection of Mdivi1 protected against dopaminergic neuron damage in a rat model of PD [105]. Similarly, another mitochondrial fission inhibitor, P110 [106], prevented the loss of dopaminergic neurons and improved motor ability in a PD mouse model [107]. The specific mechanisms involved are not well understood, but collectively these data suggest that mitochondrial fission may contribute to neuroinflammation and progressive neurodegenerative disease.

A few studies have documented key roles for mitochondrial dynamics in macrophage phagocytic responses [108–110]. Wang et al. [108] demonstrated that initial apoptotic cell uptake triggers DRP1-dependent fission in murine macrophages, with this facilitating continued clearance of the apoptotic cells. The importance of fission in this efferocytosis response was validated in vivo using myeloid-specific Drp1-knockout mice. Consistent with these findings, tumour cells resist phagocytosis by human macrophages by inhibiting mitochondrial fission in these cells, and this pathway can be targeted for effective antibody therapy against several malignancies [109]. In contrast with the pro-phagocytic activity of fission, the fusion-mediating protein MFN2 was also required for phagocytosis, as demonstrated using myeloid-specific Mfn2-knockout mice [110].

As evident in Table 2, a wide array of pathogens can modulate mitochondrial dynamics, with the functional consequences of this being either detrimental or beneficial for the pathogen. This may reflect different roles for mitochondrial dynamics in different cell types, different kinetics, and/or the specific pathogen being studied. For example, Listeria monocytogenes skewed mitochondrial dynamics towards fission in HeLa cells transiently, with mitochondria shifting back towards a more fused state over time [111]. Depleting MFN1 and MFN2 in these cells prolonged fission and impaired Listeria survival, while depleting DRP1 skewed towards fusion and favoured bacterial survival. Hence, it was postulated that the transient nature of mitochondrial fission in these cells may reflect pathogen subversion to support intracellular survival. In contrast with this study in HeLa cells, myeloid-specific Mfn2-knockout mice have fission-skewed macrophages and were more vulnerable to septic shock, as well as L. monocytogenes and Mycobacterium tuberculosis infection [110]. Similarly, Legionella pneumophila triggered mitochondrial fission and a shift towards aerobic glycolysis in human macrophages [112], with pharmacological targeting of DRP1 decreasing intracellular survival of this bacterial pathogen in these cells [78]. Such studies suggest that regulated mitochondrial dynamics may influence the host–pathogen dynamic and antimicrobial defence; however, there are major knowledge gaps regarding the underlying mechanistic details of this pathway and how it applies to different pathogens.

Although various inflammatory stimuli can modulate mitochondrial dynamics, a detailed molecular understanding of how different PRR signalling pathways exert these effects in macrophages is yet to emerge. Based on the variety of regulated PTMs on DRP1 alone, it can be speculated that altered mitochondrial dynamics is rather a universal response to many stimuli; however, the precise mechanisms involved may depend on the specific PAMP-PRR signalling pathway and/or cell type. Deconvoluting these mechanisms will be an interesting area of future research. Furthermore, different mechanisms of DRP1 activation, for example through distinct PTMs, may alter mitochondrial dynamics in different ways to elicit distinct functional outcomes. This may also be achieved through regulated or cell type-specific expression of different DRP1 transcriptional variants, of which there are many [113]. This gene regulation-mediated mechanism could also enable isoform-specific PTMs and/or functions of DRP1 [114], including differential interactions with OMM adaptor proteins such as MFF [115]. Of note, DRP1 can also shape and fragment other organelles, such as the ER and peroxisomes [116,117]. Thus, careful consideration should be taken before attributing specific biological effects to mitochondrial dynamics, based on DRP1 manipulation alone.

Another interesting research direction for the future involves the potential control of macrophage functions by intercellular transfer of mitochondria [118]. Tunnelling nanotubes for intercellular mitochondrial transfer have been studied in different contexts, such as between cancer and immune cells [119], as well as between mesenchymal stem cells and macrophages during acute respiratory distress syndrome [120]. Brestoff et al. [121] also reported immunometabolic cross-talk between adipocytes and macrophages to regulate metabolic homeostasis in obesity. A subsequent study showed that macrophages transfer mitochondria from white adipose tissue to distant organs, such as the heart, via the circulation to facilitate metabolic adaptation during nutrient stress [122]. An intriguing question in this regard is whether mitochondrial dynamics are affected during such intercellular mitochondrial transfer, both in the recipient and donor cells. Whether there is interplay between mitochondrial dynamics and mitochondrial nanotunnels [123] and/or inter-mitochondrial junctions [124] to regulate organelle behaviour and cell–cell communication will also be interesting areas of future investigation. Finally, the continued assessment of myeloid-specific knockouts of Drp1, Mfn1, Mfn2, and/or other genes controlling mitochondrial dynamics in different animal models of inflammatory and infectious diseases will be informative for understanding the in vivo functions of mitochondrial dynamics in macrophages in health and disease.

  • Mitochondria have multifaceted roles in innate immunity. Many inflammatory stimuli and pathogens regulate mitochondrial dynamics in macrophages.

  • Regulated mitochondrial dynamics control metabolic and inflammatory responses in macrophages. In myeloid cells, mitochondrial fission drives inducible glycolysis, production of specific inflammatory mediators, neuroinflammation, and phagocytosis.

  • Future investigations into mitochondrial dynamics in macrophages should focus on defining the precise molecular mechanisms by which innate immune stimuli modulate mitochondrial dynamics, the downstream mechanisms that link regulated mitochondrial dynamics to biological effects, and the contributions of mitochondrial fission and fusion to homeostatic and disease processes in vivo.

The authors declare that there are no competing interests associated with the manuscript.

M.J.S. and K.S. acknowledge the support of Australian National Health and Medical Research Council (NHMRC) Investigator Grants [APP1194406 to M.J.S.; APP2009075 to K.S.]. S.F.A., K.D.R. and G.M.E.P.L. were supported by RTP scholarships from The University of Queensland. R.K. was supported by the European Union's Horizon 2020 research and innovation programme under the Marie Sklodowska-Curie grant agreement no. [894690].

Open access for this article was enabled by the participation of the University of Queensland in an all-inclusive Read & Publish agreement with Portland Press and the Biochemical Society under a transformative agreement with CAUL.

S.F.A., M.J.S., K.D.R. and G.M.E.P.L. contributed to the conception and drafting of the manuscript. R.K. and K.S. reviewed and edited the manuscript.

DAMP

danger-associated molecular pattern

DRP1

dynamin-related protein 1

ECSIT

evolutionarily conserved signalling intermediate in Toll pathways

ETC

electron transport chain

FIS1

mitochondrial fission protein 1

IKK

inhibitor of nuclear factor-κB (IκB) kinase

IMM

inner mitochondrial membrane

LPS

lipopolysaccharide

MAL

MyD88-adapter-like

MAVS

mitochondrial antiviral signalling protein

MDA-5

melanoma differentiation-associated gene 5

MFF

mitochondrial fission factor

MFN

mitofusin

NLRP3

NOD-, LRR- and pyrin domain-containing protein 3

OMM

outer mitochondrial membrane

OXPHOS

oxidative phosphorylation

PAMP

pathogen-associated molecular pattern

PRR

pattern recognition receptor

PTM

post-translational modification

RLR

retinoic acid-inducible gene 1 (RIG-1)-like helicase receptor

ROS

reactive oxygen species

TBK1

TANK-binding kinase 1

TCA

tricarboxylic acid

TFAM

mitochondrial transcription factor A

TLR

Toll-like receptor

TRAF6

tumour necrosis factor receptor-associated factor 6

1
Hirayama
,
D.
,
Iida
,
T.
and
Nakase
,
H.
(
2017
)
The phagocytic function of macrophage-enforcing innate immunity and tissue homeostasis
.
Int. J. Mol. Sci.
19
,
92
2
Wynn
,
T.A.
and
Vannella
,
K.M.
(
2016
)
Macrophages in tissue repair, regeneration, and fibrosis
.
Immunity
44
,
450
462
3
Mogensen
,
T.H.
(
2009
)
Pathogen recognition and inflammatory signaling in innate immune defenses
.
Clin. Microbiol. Rev.
22
,
240
273
4
Chandel
,
N.S.
(
2014
)
Mitochondria as signaling organelles
.
BMC Biol.
12
,
34
5
Yang
,
J.
,
Liu
,
X.
,
Bhalla
,
K.
,
Kim
,
C.N.
,
Ibrado
,
A.M.
,
Cai
,
J.
et al (
1997
)
Prevention of apoptosis by Bcl-2: release of cytochrome c from mitochondria blocked
.
Science
275
,
1129
1132
6
Aon
,
M.A.
and
Camara
,
A.K.
(
2015
)
Mitochondria: hubs of cellular signaling, energetics and redox balance. A rich, vibrant, and diverse landscape of mitochondrial research
.
Front. Physiol.
6
,
94
7
Aguilar-Lopez
,
B.A.
,
Moreno-Altamirano
,
M.M.B.
,
Dockrell
,
H.M.
,
Duchen
,
M.R.
and
Sanchez-Garcia
,
F.J.
(
2020
)
Mitochondria: an integrative hub coordinating circadian rhythms, metabolism, the microbiome, and immunity
.
Front. Cell Dev. Biol.
8
,
51
8
Tiku
,
V.
,
Tan
,
M.W.
and
Dikic
,
I.
(
2020
)
Mitochondrial functions in infection and immunity
.
Trends Cell Biol.
30
,
263
275
9
Grazioli
,
S.
and
Pugin
,
J.
(
2018
)
Mitochondrial damage-associated molecular patterns: from inflammatory signaling to human diseases
.
Front. Immunol.
9
,
832
10
Nair
,
S.
,
Sobotka
,
K.S.
,
Joshi
,
P.
,
Gressens
,
P.
,
Fleiss
,
B.
,
Thornton
,
C.
et al (
2019
)
Lipopolysaccharide-induced alteration of mitochondrial morphology induces a metabolic shift in microglia modulating the inflammatory response in vitro and in vivo
.
Glia
67
,
1047
1061
11
Zhang
,
L.
,
Ma
,
C.
,
Wang
,
X.
,
He
,
S.
,
Li
,
Q.
,
Zhou
,
Y.
et al (
2019
)
Lipopolysaccharide-induced proliferation and glycolysis in airway smooth muscle cells via activation of Drp1
.
J. Cell Physiol.
234
,
9255
9263
12
Kapetanovic
,
R.
,
Afroz
,
S.F.
,
Ramnath
,
D.
,
Lawrence
,
G.M.
,
Okada
,
T.
,
Curson
,
J.E.
et al (
2020
)
Lipopolysaccharide promotes Drp1-dependent mitochondrial fission and associated inflammatory responses in macrophages
.
Immunol. Cell Biol.
98
,
528
539
13
de Oliveira
,
L.G.
,
Angelo
,
Y.S.
,
Iglesias
,
A.H.
and
Peron
,
J.P.S.
(
2021
)
Unraveling the link between mitochondrial dynamics and neuroinflammation
.
Front. Immunol.
12
,
624919
14
Park
,
J.
,
Choi
,
H.
,
Min
,
J.S.
,
Park
,
S.J.
,
Kim
,
J.H.
,
Park
,
H.J.
et al (
2013
)
Mitochondrial dynamics modulate the expression of pro-inflammatory mediators in microglial cells
.
J. Neurochem.
127
,
221
232
15
Joshi
,
A.U.
,
Minhas
,
P.S.
,
Liddelow
,
S.A.
,
Haileselassie
,
B.
,
Andreasson
,
K.I.
,
Dorn
, II,
G.W.
et al (
2019
)
Fragmented mitochondria released from microglia trigger A1 astrocytic response and propagate inflammatory neurodegeneration
.
Nat. Neurosci.
22
,
1635
1648
16
Lane
,
N.
and
Martin
,
W.
(
2010
)
The energetics of genome complexity
.
Nature
467
,
929
934
17
Gray
,
M.W.
(
2015
)
Mosaic nature of the mitochondrial proteome: implications for the origin and evolution of mitochondria
.
Proc. Natl Acad. Sci. U.S.A.
112
,
10133
10138
18
Yoneyama
,
M.
,
Kikuchi
,
M.
,
Natsukawa
,
T.
,
Shinobu
,
N.
,
Imaizumi
,
T.
,
Miyagishi
,
M.
et al (
2004
)
The RNA helicase RIG-I has an essential function in double-stranded RNA-induced innate antiviral responses
.
Nat. Immunol.
5
,
730
737
19
Andrejeva
,
J.
,
Childs
,
K.S.
,
Young
,
D.F.
,
Carlos
,
T.S.
,
Stock
,
N.
,
Goodbourn
,
S.
et al (
2004
)
The V proteins of paramyxoviruses bind the IFN-inducible RNA helicase, mda-5, and inhibit its activation of the IFN-beta promoter
.
Proc. Natl Acad. Sci. U.S.A.
101
,
17264
17269
20
Fitzgerald
,
K.A.
,
McWhirter
,
S.M.
,
Faia
,
K.L.
,
Rowe
,
D.C.
,
Latz
,
E.
,
Golenbock
,
D.T.
et al (
2003
)
IKKepsilon and TBK1 are essential components of the IRF3 signaling pathway
.
Nat. Immunol.
4
,
491
496
21
Guo
,
B.
and
Cheng
,
G.
(
2007
)
Modulation of the interferon antiviral response by the TBK1/IKKi adaptor protein TANK
.
J. Biol. Chem.
282
,
11817
11826
22
Lauterbach
,
M.A.
,
Hanke
,
J.E.
,
Serefidou
,
M.
,
Mangan
,
M.S.J.
,
Kolbe
,
C.C.
,
Hess
,
T.
et al (
2019
)
Toll-like receptor signaling rewires macrophage metabolism and promotes histone acetylation via ATP-citrate lyase
.
Immunity
51
,
997
1011.e7
23
Jha
,
A.K.
,
Huang
,
S.C.
,
Sergushichev
,
A.
,
Lampropoulou
,
V.
,
Ivanova
,
Y.
,
Loginicheva
,
E.
et al (
2015
)
Network integration of parallel metabolic and transcriptional data reveals metabolic modules that regulate macrophage polarization
.
Immunity
42
,
419
430
24
Tannahill
,
G.M.
,
Curtis
,
A.M.
,
Adamik
,
J.
,
Palsson-McDermott
,
E.M.
,
McGettrick
,
A.F.
,
Goel
,
G.
et al (
2013
)
Succinate is an inflammatory signal that induces IL-1beta through HIF-1alpha
.
Nature
496
,
238
242
25
Mills
,
E.L.
,
Kelly
,
B.
,
Logan
,
A.
,
Costa
,
A.S.H.
,
Varma
,
M.
,
Bryant
,
C.E.
et al (
2016
)
Succinate dehydrogenase supports metabolic repurposing of mitochondria to drive inflammatory macrophages
.
Cell
167
,
457
470.e13
26
Michelucci
,
A.
,
Cordes
,
T.
,
Ghelfi
,
J.
,
Pailot
,
A.
,
Reiling
,
N.
,
Goldmann
,
O.
et al (
2013
)
Immune-responsive gene 1 protein links metabolism to immunity by catalyzing itaconic acid production
.
Proc. Natl Acad. Sci. U.S.A.
110
,
7820
7825
27
Mills
,
E.L.
,
Ryan
,
D.G.
,
Prag
,
H.A.
,
Dikovskaya
,
D.
,
Menon
,
D.
,
Zaslona
,
Z.
et al (
2018
)
Itaconate is an anti-inflammatory metabolite that activates Nrf2 via alkylation of KEAP1
.
Nature
556
,
113
117
28
Lampropoulou
,
V.
,
Sergushichev
,
A.
,
Bambouskova
,
M.
,
Nair
,
S.
,
Vincent
,
E.E.
,
Loginicheva
,
E.
et al (
2016
)
Itaconate links inhibition of succinate dehydrogenase with macrophage metabolic remodeling and regulation of inflammation
.
Cell Metab.
24
,
158
166
29
West
,
A.P.
,
Brodsky
,
I.E.
Rahner
,
C.
,
Woo
,
D.K.
,
Erdjument-Bromage
,
H.
,
Tempst
,
P.
et al (
2011
)
TLR signalling augments macrophage bactericidal activity through mitochondrial ROS
.
Nature
472
,
476
480
30
Balic
,
J.J.
,
Albargy
,
H.
,
Luu
,
K.
,
Kirby
,
F.J.
,
Jayasekara
,
W.S.N.
,
Mansell
,
F.
et al (
2020
)
STAT3 serine phosphorylation is required for TLR4 metabolic reprogramming and IL-1beta expression
.
Nat. Commun.
11
,
3816
31
Wu
,
J.
,
Yan
,
Z.
,
Schwartz
,
D.E.
,
Yu
,
J.
,
Malik
,
A.B.
and
Hu
,
G.
(
2013
)
Activation of NLRP3 inflammasome in alveolar macrophages contributes to mechanical stretch-induced lung inflammation and injury
.
J. Immunol.
190
,
3590
3599
32
Zhou
,
R.
,
Yazdi
,
A.S.
,
Menu
,
P.
and
Tschopp
,
J.
(
2011
)
A role for mitochondria in NLRP3 inflammasome activation
.
Nature
469
,
221
225
33
Wiegman
,
C.H.
,
Michaeloudes
,
C.
,
Haji
,
G.
,
Narang
,
P.
,
Clarke
,
C.J.
,
Russell
,
K.E.
et al (
2015
)
Oxidative stress-induced mitochondrial dysfunction drives inflammation and airway smooth muscle remodeling in patients with chronic obstructive pulmonary disease
.
J. Allergy Clin. Immunol.
136
,
769
780
34
Tirichen
,
H.
,
Yaigoub
,
H.
,
Xu
,
W.
,
Wu
,
C.
,
Li
,
R.
and
Li
,
Y.
(
2021
)
Mitochondrial reactive oxygen species and their contribution in chronic kidney disease progression through oxidative stress
.
Front. Physiol.
12
,
627837
35
Pereira
,
C.A.
,
Carlos
,
D.
,
Ferreira
,
N.S.
,
Silva
,
J.F.
,
Zanotto
,
C.Z.
,
Zamboni
,
D.S.
et al (
2019
)
Mitochondrial DNA promotes NLRP3 inflammasome activation and contributes to endothelial dysfunction and inflammation in type 1 diabetes
.
Front. Physiol.
10
,
1557
36
Patoli
,
D.
,
Mignotte
,
F.
,
Deckert
,
V.
,
Dusuel
,
A.
,
Dumont
,
A.
,
Rieu
,
A.
et al (
2020
)
Inhibition of mitophagy drives macrophage activation and antibacterial defense during sepsis
.
J. Clin. Invest.
130
,
5858
5874
37
Crouser
,
E.D.
,
Shao
,
G.
,
Julian
,
M.W.
,
Macre
,
J.E.
,
Shadel
,
G.S.
,
Tridandapani
,
S.
et al (
2009
)
Monocyte activation by necrotic cells is promoted by mitochondrial proteins and formyl peptide receptors
.
Crit. Care Med.
37
,
2000
2009
38
Pittman
,
K.
and
Kubes
,
P.
(
2013
)
Damage-associated molecular patterns control neutrophil recruitment
.
J. Innate Immun.
5
,
315
323
39
Collins
,
L.V.
,
Hajizadeh
,
S.
,
Holme
,
E.
,
Jonsson
,
I.M.
and
Tarkowski
,
A.
(
2004
)
Endogenously oxidized mitochondrial DNA induces in vivo and in vitro inflammatory responses
.
J. Leukoc. Biol.
75
,
995
1000
40
Zhang
,
Q.
,
Raoof
,
M.
,
Chen
,
Y.
,
Sumi
,
Y.
,
Sursal
,
T.
,
Junger
,
W.
et al (
2010
)
Circulating mitochondrial DAMPs cause inflammatory responses to injury
.
Nature
464
,
104
107
41
Oka
,
T.
,
Hikoso
,
S.
,
Yamaguchi
,
O.
,
Taneike
,
M.
,
Takeda
,
T.
,
Tamai
,
T.
et al (
2012
)
Mitochondrial DNA that escapes from autophagy causes inflammation and heart failure
.
Nature
485
,
251
255
42
Nakayama
,
H.
and
Otsu
,
K.
(
2018
)
Mitochondrial DNA as an inflammatory mediator in cardiovascular diseases
.
Biochem. J.
475
,
839
852
43
Sesaki
,
H.
and
Jensen
,
R.E.
(
1999
)
Division versus fusion: Dnm1p and Fzo1p antagonistically regulate mitochondrial shape
.
J. Cell Biol.
147
,
699
706
44
Chen
,
H.
,
Detmer
,
S.A.
,
Ewald
,
A.J.
,
Griffin
,
E.E.
,
Fraser
,
S.E.
and
Chan
,
D.C.
(
2003
)
Mitofusins Mfn1 and Mfn2 coordinately regulate mitochondrial fusion and are essential for embryonic development
.
J. Cell Biol.
160
,
189
200
45
Wakabayashi
,
J.
,
Zhang
,
Z.
,
Wakabayashi
,
N.
,
Tamura
,
Y.
,
Fukaya
,
M.
,
Kensler
,
T.W.
et al (
2009
)
The dynamin-related GTPase Drp1 is required for embryonic and brain development in mice
.
J. Cell Biol.
186
,
805
816
46
Rambold
,
A.S.
,
Kostelecky
,
B.
,
Elia
,
N.
and
Lippincott-Schwartz
,
J.
(
2011
)
Tubular network formation protects mitochondria from autophagosomal degradation during nutrient starvation
.
Proc. Natl Acad. Sci. U.S.A.
108
,
10190
10195
47
Blackstone
,
C.
and
Chang
,
C.R.
(
2011
)
Mitochondria unite to survive
.
Nat. Cell Biol.
13
,
521
522
48
Molina
,
A.J.
,
Wikstrom
,
J.D.
,
Stiles
,
L.
,
Las
,
G.
,
Mohamed
,
H.
,
Elorza
,
A.
et al (
2009
)
Mitochondrial networking protects beta-cells from nutrient-induced apoptosis
.
Diabetes
58
,
2303
2315
49
Liu
,
X.
and
Hajnoczky
,
G.
(
2011
)
Altered fusion dynamics underlie unique morphological changes in mitochondria during hypoxia-reoxygenation stress
.
Cell Death Differ.
18
,
1561
1572
50
Wikstrom
,
J.D.
,
Mahdaviani
,
K.
,
Liesa
,
M.
,
Sereda
,
S.B.
,
Si
,
Y.
,
Las
,
G.
et al (
2014
)
Hormone-induced mitochondrial fission is utilized by brown adipocytes as an amplification pathway for energy expenditure
.
EMBO J.
33
,
418
436
51
Bleazard
,
W.
,
McCaffery
,
J.M.
,
King
,
E.J.
,
Bale
,
S.
,
Mozdy
,
A.
,
Tieu
,
Q.
et al (
1999
)
The dynamin-related GTPase Dnm1 regulates mitochondrial fission in yeast
.
Nat. Cell Biol.
1
,
298
304
52
Fonseca
,
T.B.
,
Sanchez-Guerrero
,
A.
,
Milosevic
,
I.
and
Raimundo
,
N.
(
2019
)
Mitochondrial fission requires DRP1 but not dynamins
.
Nature
570
,
E34
E42
53
Legesse-Miller
,
A.
,
Massol
,
R.H.
and
Kirchhausen
,
T.
(
2003
)
Constriction and Dnm1p recruitment are distinct processes in mitochondrial fission
.
Mol. Biol. Cell
14
,
1953
1963
54
Friedman
,
J.R.
,
Lackner
,
L.L.
,
West
,
M.
,
DiBenedetto
,
J.R.
,
Nunnari
,
J.
and
Voeltz
,
G.K.
(
2011
)
ER tubules mark sites of mitochondrial division
.
Science
334
,
358
362
55
Prudent
,
J.
and
McBride
,
H.M.
(
2016
)
Mitochondrial dynamics: ER actin tightens the Drp1 noose
.
Curr. Biol.
26
,
R207
R209
56
Ji
,
W.K.
,
Hatch
,
A.L.
,
Merrill
,
R.A.
,
Strack
,
S.
and
Higgs
,
H.N.
(
2015
)
Actin filaments target the oligomeric maturation of the dynamin GTPase Drp1 to mitochondrial fission sites
.
eLife
4
,
e11553
57
Sonn
,
S.K.
,
Seo
,
S.
,
Yang
,
J.
,
Oh
,
K.S.
,
Chen
,
H.
,
Chan
,
D.C.
et al (
2021
)
ER-associated CTRP1 regulates mitochondrial fission via interaction with DRP1
.
Exp. Mol. Med.
53
,
1769
1780
58
Smirnova
,
E.
,
Griparic
,
L.
,
Shurland
,
D.L.
and
van der Bliek
,
A.M.
(
2001
)
Dynamin-related protein Drp1 is required for mitochondrial division in mammalian cells
.
Mol. Biol. Cell
12
,
2245
2256
59
Yoon
,
Y.
,
Krueger
,
E.W.
,
Oswald
,
B.J.
and
McNiven
,
M.A.
(
2003
)
The mitochondrial protein hFis1 regulates mitochondrial fission in mammalian cells through an interaction with the dynamin-like protein DLP1
.
Mol. Cell. Biol.
23
,
5409
5420
60
Gandre-Babbe
,
S.
and
van der Bliek
,
A.M.
(
2008
)
The novel tail-anchored membrane protein Mff controls mitochondrial and peroxisomal fission in mammalian cells
.
Mol. Biol. Cell
19
,
2402
2412
61
Zhao
,
J.
,
Liu
,
T.
,
Jin
,
S.
,
Wang
,
X.
,
Qu
,
M.
,
Uhlen
,
P.
et al (
2011
)
Human MIEF1 recruits Drp1 to mitochondrial outer membranes and promotes mitochondrial fusion rather than fission
.
EMBO J.
30
,
2762
2778
62
Otera
,
H.
,
Wang
,
C.
,
Cleland
,
M.M.
,
Setoguchi
,
K.
,
Yokota
,
S.
,
Youle
,
R.J.
et al (
2010
)
Mff is an essential factor for mitochondrial recruitment of Drp1 during mitochondrial fission in mammalian cells
.
J. Cell Biol.
191
,
1141
1158
63
Loson
,
O.C.
,
Song
,
Z.
,
Chen
,
H.
and
Chan
,
D.C.
(
2013
)
Fis1, Mff, MiD49, and MiD51 mediate Drp1 recruitment in mitochondrial fission
.
Mol. Biol. Cell
24
,
659
667
64
Liu
,
T.
,
Yu
,
R.
,
Jin
,
S.B.
,
Han
,
L.
,
Lendahl
,
U.
,
Zhao
,
J.
et al (
2013
)
The mitochondrial elongation factors MIEF1 and MIEF2 exert partially distinct functions in mitochondrial dynamics
.
Exp. Cell Res.
319
,
2893
2904
65
Zhang
,
Z.
,
Liu
,
L.
,
Wu
,
S.
and
Xing
,
D.
(
2016
)
Drp1, Mff, Fis1, and MiD51 are coordinated to mediate mitochondrial fission during UV irradiation-induced apoptosis
.
FASEB J.
30
,
466
476
66
Kleele
,
T.
,
Rey
,
T.
,
Winter
,
J.
,
Zaganelli
,
S.
,
Mahecic
,
D.
,
Perreten Lambert
,
H.
et al (
2021
)
Distinct fission signatures predict mitochondrial degradation or biogenesis
.
Nature
593
,
435
439
67
Scott
,
I.
and
Youle
,
R.J.
(
2010
)
Mitochondrial fission and fusion
.
Essays Biochem.
47
,
85
98
68
Ban
,
T.
,
Ishihara
,
T.
,
Kohno
,
H.
,
Saita
,
S.
,
Ichimura
,
A.
,
Maenaka
,
K.
et al (
2017
)
Molecular basis of selective mitochondrial fusion by heterotypic action between OPA1 and cardiolipin
.
Nat. Cell Biol.
19
,
856
863
69
Ban
,
T.
,
Kohno
,
H.
,
Ishihara
,
T.
and
Ishihara
,
N.
(
2018
)
Relationship between OPA1 and cardiolipin in mitochondrial inner-membrane fusion
.
Biochim. Biophys. Acta Bioenerg.
1859
,
951
957
70
Cipolat
,
S.
,
de Brito O
,
M.
,
Zilio B
,
D.
and
Scorrano
,
L.
(
2004
)
OPA1 requires mitofusin 1 to promote mitochondrial fusion
.
Proc. Natl Acad. Sci. U.S.A.
101
,
15927
15932
71
Zhang
,
Y.
,
Liu
,
X.
,
Bai
,
J.
,
Tian
,
X.
,
Zhao
,
X.
,
Liu
,
W.
et al (
2016
)
Mitoguardin regulates mitochondrial fusion through MitoPLD and is required for neuronal homeostasis
.
Mol. Cell
61
,
111
124
72
Lee
,
J.Y.
,
Kapur
,
M.
,
Li
,
M.
,
Choi
,
M.C.
,
Choi
,
S.
,
Kim
,
H.J.
et al (
2014
)
MFN1 deacetylation activates adaptive mitochondrial fusion and protects metabolically challenged mitochondria
.
J. Cell Sci.
127
,
4954
4963
73
Samant
,
S.A.
,
Zhang
,
H.J.
,
Hong
,
Z.
,
Pillai
,
V.B.
,
Sundaresan
,
N.R.
,
Wolfgeher
,
D.
et al (
2014
)
SIRT3 deacetylates and activates OPA1 to regulate mitochondrial dynamics during stress
.
Mol. Cell. Biol.
34
,
807
819
74
Leboucher
,
G.P.
,
Tsai
,
Y.C.
,
Yang
,
M.
,
Shaw
,
K.C.
,
Zhou
,
M.
,
Veenstra
,
T.D.
et al (
2012
)
Stress-induced phosphorylation and proteasomal degradation of mitofusin 2 facilitates mitochondrial fragmentation and apoptosis
.
Mol. Cell
47
,
547
557
75
Yao
,
C.H.
,
Wang
,
R.
,
Wang
,
Y.
,
Kung
,
C.P.
,
Weber
,
J.D.
and
Patti
,
G.J.
(
2019
)
Mitochondrial fusion supports increased oxidative phosphorylation during cell proliferation
.
eLife
8
,
e41351
76
Buck
,
M.D.
,
O'Sullivan
,
D.
,
Klein Geltink
,
R.I.
,
Curtis
,
J.D.
,
Chang
,
C.H.
,
Sanin
,
D.E.
et al (
2016
)
Mitochondrial dynamics controls T cell fate through metabolic programming
.
Cell
166
,
63
76
77
Li
,
T.
,
Han
,
J.
,
Jia
,
L.
,
Hu
,
X.
,
Chen
,
L.
and
Wang
,
Y.
(
2019
)
PKM2 coordinates glycolysis with mitochondrial fusion and oxidative phosphorylation
.
Protein Cell
10
,
583
594
78
Cassidy-Stone
,
A.
,
Chipuk
,
J.E.
,
Ingerman
,
E.
,
Song
,
C.
,
Yoo
,
C.
,
Kuwana
,
T.
et al (
2008
)
Chemical inhibition of the mitochondrial division dynamin reveals its role in Bax/Bak-dependent mitochondrial outer membrane permeabilization
.
Dev. Cell
14
,
193
204
79
Pantiya
,
P.
,
Thonusin
,
C.
,
Chattipakorn
,
N.
and
Chattipakorn
,
S.C.
(
2020
)
Mitochondrial abnormalities in neurodegenerative models and possible interventions: focus on Alzheimer's disease, Parkinson's disease, Huntington's disease
.
Mitochondrion
55
,
14
47
80
Gao
,
F.
,
Reynolds
,
M.B.
,
Passalacqua
,
K.D.
,
Sexton
,
J.Z.
,
Abuaita
,
B.H.
and
O'Riordan
,
M.X.D.
(
2020
)
The mitochondrial fission regulator DRP1 controls post-transcriptional regulation of TNF-alpha
.
Front. Cell Infect. Microbiol.
10
,
593805
81
Gao
,
Z.
,
Li
,
Y.
,
Wang
,
F.
,
Huang
,
T.
,
Fan
,
K.
,
Zhang
,
Y.
et al (
2017
)
Mitochondrial dynamics controls anti-tumour innate immunity by regulating CHIP-IRF1 axis stability
.
Nat. Commun.
8
,
1805
82
Khodzhaeva
,
V.
,
Schreiber
,
Y.
,
Geisslinger
,
G.
,
Brandes
,
R.P.
,
Brune
,
B.
and
Namgaladze
,
D.
(
2021
)
Mitofusin 2 deficiency causes pro-inflammatory effects in human primary macrophages
.
Front. Immunol.
12
,
723683
83
Ichinohe
,
T.
,
Yamazaki
,
T.
,
Koshiba
,
T.
and
Yanagi
,
Y.
(
2013
)
Mitochondrial protein mitofusin 2 is required for NLRP3 inflammasome activation after RNA virus infection
.
Proc. Natl Acad. Sci. U.S.A.
110
,
17963
17968
84
Park
,
S.
,
Won
,
J.H.
,
Hwang
,
I.
,
Hong
,
S.
,
Lee
,
H.K.
and
Yu
,
J.W.
(
2015
)
Defective mitochondrial fission augments NLRP3 inflammasome activation
.
Sci. Rep.
5
,
15489
85
Salter
,
M.W.
and
Beggs
,
S.
(
2014
)
Sublime microglia: expanding roles for the guardians of the CNS
.
Cell
158
,
15
24
86
Jin
,
X.
and
Yamashita
,
T.
(
2016
)
Microglia in central nervous system repair after injury
.
J. Biochem.
159
,
491
496
87
Yang
,
I.
,
Han
,
S.J.
,
Kaur
,
G.
,
Crane
,
C.
and
Parsa
,
A.T.
(
2010
)
The role of microglia in central nervous system immunity and glioma immunology
.
J. Clin. Neurosci.
17
,
6
10
88
Smith
,
J.A.
,
Das
,
A.
,
Ray
,
S.K.
and
Banik
,
N.L.
(
2012
)
Role of pro-inflammatory cytokines released from microglia in neurodegenerative diseases
.
Brain Res. Bull.
87
,
10
20
89
Block
,
M.L.
,
Zecca
,
L.
and
Hong
,
J.S.
(
2007
)
Microglia-mediated neurotoxicity: uncovering the molecular mechanisms
.
Nat. Rev. Neurosci.
8
,
57
69
90
Gordon
,
R.
,
Albornoz
,
E.A.
,
Christie
,
D.C.
,
Langley
,
M.R.
,
Kumar
,
V.
,
Mantovani
,
S.
et al (
2018
)
Inflammasome inhibition prevents alpha-synuclein pathology and dopaminergic neurodegeneration in mice
.
Sci. Transl. Med.
10
,
eaah4066
91
Langston
,
J.W.
,
Forno
,
L.S.
,
Tetrud
,
J.
,
Reeves
,
A.G.
,
Kaplan
,
J.A.
and
Karluk
,
D.
(
1999
)
Evidence of active nerve cell degeneration in the substantia nigra of humans years after 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine exposure
.
Ann. Neurol.
46
,
598
605
92
Ferrari
,
C.C.
,
Pott Godoy
,
M.C.
,
Tarelli
,
R.
,
Chertoff
,
M.
,
Depino
,
A.M.
and
Pitossi
,
F.J.
(
2006
)
Progressive neurodegeneration and motor disabilities induced by chronic expression of IL-1β in the substantia nigra
.
Neurobiol. Dis.
24
,
183
193
93
Bornemann
,
K.D.
,
Wiederhold
,
K.H.
,
Pauli
,
C.
,
Ermini
,
F.
,
Stalder
,
M.
,
Schnell
,
L.
et al (
2001
)
Abeta-induced inflammatory processes in microglia cells of APP23 transgenic mice
.
Am. J. Pathol.
158
,
63
73
94
Beier
,
E.E.
,
Neal
,
M.
,
Alam
,
G.
,
Edler
,
M.
,
Wu
,
L.-J.
and
Richardson
,
J.R.
(
2017
)
Alternative microglial activation is associated with cessation of progressive dopamine neuron loss in mice systemically administered lipopolysaccharide
.
Neurobiol. Dis.
108
,
115
127
95
Qin
,
L.
,
Wu
,
X.
,
Block
,
M.L.
,
Liu
,
Y.
,
Breese
,
G.R.
,
Hong
,
J.-S.
et al (
2007
)
Systemic LPS causes chronic neuroinflammation and progressive neurodegeneration
.
Glia
55
,
453
462
96
Katoh
,
M.
,
Wu
,
B.
,
Nguyen
,
H.B.
,
Thai
,
T.Q.
,
Yamasaki
,
R.
,
Lu
,
H.
et al (
2017
)
Polymorphic regulation of mitochondrial fission and fusion modifies phenotypes of microglia in neuroinflammation
.
Sci. Rep.
7
,
4942
97
Park
,
J.
,
Min
,
J.-S.
,
Chae
,
U.
,
Lee
,
J.Y.
,
Song
,
K.-S.
,
Lee
,
H.-S.
et al (
2017
)
Anti-inflammatory effect of oleuropein on microglia through regulation of Drp1-dependent mitochondrial fission
.
J. Neuroimmunol.
306
,
46
52
98
Chae
,
U.
,
Min
,
J.-S.
,
Lee
,
H.
,
Song
,
K.-S.
,
Lee
,
H.-S.
,
Lee
,
H.J.
et al (
2017
)
Chrysophanol suppresses pro-inflammatory response in microglia via regulation of Drp1-dependent mitochondrial fission
.
Immunopharmacol. Immunotoxicol.
39
,
268
275
99
Cheng
,
J.
,
Zhang
,
R.
,
Xu
,
Z.
,
Ke
,
Y.
,
Sun
,
R.
,
Yang
,
H.
et al (
2021
)
Early glycolytic reprogramming controls microglial inflammatory activation
.
J. Neuroinflammation
18
,
129
100
Shimada
,
K.
,
Crother Timothy
,
R.
,
Karlin
,
J.
,
Dagvadorj
,
J.
,
Chiba
,
N.
,
Chen
,
S.
et al (
2012
)
Oxidized mitochondrial DNA activates the NLRP3 inflammasome during apoptosis
.
Immunity
36
,
401
414
101
Christina J
,
G.
,
Mishra
,
R.
,
Schneider Katharina
,
S.
,
Médard
,
G.
,
Wettmarshausen
,
J.
,
Dittlein Daniela
,
C.
et al (
2016
)
K+ efflux-independent NLRP3 inflammasome activation by small molecules targeting mitochondria
.
Immunity
45
,
761
773
102
Holley
,
C.L.
and
Schroder
,
K.
(
2020
)
The rOX-stars of inflammation: links between the inflammasome and mitochondrial meltdown
.
Clin. Transl. Immunol.
9
,
e01109
103
Zhang
,
X.-L.
,
Huang
,
W.-M.
,
Tang
,
P.-C.
,
Sun
,
Y.
,
Zhang
,
X.
,
Qiu
,
L.
et al (
2021
)
Anti-inflammatory and neuroprotective effects of natural cordycepin in rotenone-induced PD models through inhibiting Drp1-mediated mitochondrial fission
.
Neurotoxicology
84
,
1
13
104
Liu
,
R.
,
Wang
,
S.-C.
,
Li
,
M.
,
Ma
,
X.-H.
,
Jia
,
X.-N.
,
Bu
,
Y.
et al (
2020
)
An inhibitor of DRP1 (Mdivi-1) alleviates LPS-induced septic AKI by inhibiting NLRP3 inflammasome activation
.
Biomed. Res. Int.
2020
,
2398420-11
105
Bido
,
S.
,
Soria
,
F.N.
,
Fan
,
R.Z.
,
Bezard
,
E.
and
Tieu
,
K.
(
2017
)
Mitochondrial division inhibitor-1 is neuroprotective in the A53T-α-synuclein rat model of Parkinson's disease
.
Sci. Rep.
7
,
7495
7413
106
Qi
,
X.
,
Qvit
,
N.
,
Su
,
Y.C.
and
Mochly-Rosen
,
D.
(
2013
)
A novel Drp1 inhibitor diminishes aberrant mitochondrial fission and neurotoxicity
.
J. Cell Sci.
126
,
789
802
107
Filichia
,
E.
,
Hoffer
,
B.
,
Qi
,
X.
and
Luo
,
Y.
(
2016
)
Inhibition of Drp1 mitochondrial translocation provides neural protection in dopaminergic system in a Parkinson's disease model induced by MPTP
.
Sci. Rep.
6
,
32656
108
Wang
,
Y.
,
Subramanian
,
M.
,
Yurdagul
, Jr,
A.
,
Barbosa-Lorenzi
,
V.C.
,
Cai
,
B.
,
de Juan-Sanz
,
J.
et al (
2017
)
Mitochondrial fission promotes the continued clearance of apoptotic cells by macrophages
.
Cell
171
,
331
345.e22
109
Li
,
J.
,
Ye
,
Y.
,
Liu
,
Z.
,
Zhang
,
G.
,
Dai
,
H.
,
Li
,
J.
et al (
2022
)
Macrophage mitochondrial fission improves cancer cell phagocytosis induced by therapeutic antibodies and is impaired by glutamine competition
.
Nat. Cancer
3
,
453
470
110
Tur
,
J.
,
Pereira-Lopes
,
S.
,
Vico
,
T.
,
Marin
,
E.A.
,
Munoz
,
J.P.
,
Hernandez-Alvarez
,
M.
et al (
2020
)
Mitofusin 2 in macrophages links mitochondrial ROS production, cytokine release, phagocytosis, autophagy, and bactericidal activity
.
Cell Rep.
32
,
108079
111
Stavru
,
F.
,
Bouillaud
,
F.
,
Sartori
,
A.
,
Ricquier
,
D.
and
Cossart
,
P.
(
2011
)
Listeria monocytogenes transiently alters mitochondrial dynamics during infection
.
Proc. Natl Acad. Sci. U.S.A.
108
,
3612
3617
112
Escoll
,
P.
,
Song
,
O.R.
,
Viana
,
F.
,
Steiner
,
B.
,
Lagache
,
T.
,
Olivo-Marin
,
J.C.
et al (
2017
)
Legionella pneumophila modulates mitochondrial dynamics to trigger metabolic repurposing of infected macrophages
.
Cell Host Microbe
22
,
302
316.e7
113
Uo
,
T.
,
Dworzak
,
J.
,
Kinoshita
,
C.
,
Inman
,
D.M.
,
Kinoshita
,
Y.
,
Horner
,
P.J.
et al (
2009
)
Drp1 levels constitutively regulate mitochondrial dynamics and cell survival in cortical neurons
.
Exp. Neurol.
218
,
274
285
114
Strack
,
S.
,
Wilson
,
T.J.
and
Cribbs
,
J.T.
(
2013
)
Cyclin-dependent kinases regulate splice-specific targeting of dynamin-related protein 1 to microtubules
.
J. Cell Biol.
201
,
1037
1051
115
Macdonald
,
P.J.
,
Francy
,
C.A.
,
Stepanyants
,
N.
,
Lehman
,
L.
,
Baglio
,
A.
,
Mears
,
J.A.
et al (
2016
)
Distinct splice variants of dynamin-related protein 1 differentially utilize mitochondrial fission factor as an effector of cooperative GTPase activity
.
J. Biol. Chem.
291
,
493
507
116
Kamerkar
,
S.C.
,
Kraus
,
F.
,
Sharpe
,
A.J.
,
Pucadyil
,
T.J.
and
Ryan
,
M.T.
(
2018
)
Dynamin-related protein 1 has membrane constricting and severing abilities sufficient for mitochondrial and peroxisomal fission
.
Nat. Commun.
9
,
5239
117
Schrader
,
M.
(
2006
)
Shared components of mitochondrial and peroxisomal division
.
Biochim. Biophys. Acta
1763
,
531
541
118
Pang
,
Y.
,
Zhang
,
C.
and
Gao
,
J.
(
2021
)
Macrophages as emerging key players in mitochondrial transfers
.
Front. Cell Dev. Biol.
9
,
747377
119
Saha
,
T.
,
Dash
,
C.
,
Jayabalan
,
R.
,
Khiste
,
S.
,
Kulkarni
,
A.
,
Kurmi
,
K.
et al (
2022
)
Intercellular nanotubes mediate mitochondrial trafficking between cancer and immune cells
.
Nat. Nanotechnol.
17
,
98
106
120
Jackson
,
M.V.
,
Morrison
,
T.J.
,
Doherty
,
D.F.
,
McAuley
,
D.F.
,
Matthay
,
M.A.
,
Kissenpfennig
,
A.
et al (
2016
)
Mitochondrial transfer via tunneling nanotubes is an important mechanism by which mesenchymal stem cells enhance macrophage phagocytosis in the in vitro and in vivo models of ARDS
.
Stem Cells
34
,
2210
2223
121
Brestoff
,
J.R.
,
Wilen
,
C.B.
,
Moley
,
J.R.
,
Li
,
Y.
,
Zou
,
W.
,
Malvin
,
N.P.
et al (
2021
)
Intercellular mitochondria transfer to macrophages regulates white adipose tissue homeostasis and is impaired in obesity
.
Cell Metab.
33
,
270
282.e8
122
Borcherding
,
N.
,
Jia
,
W.
,
Giwa
,
R.
,
Field
,
R.L.
,
Moley
,
J.R.
,
Kopecky
,
B.J.
et al (
2022
)
Dietary lipids inhibit mitochondria transfer to macrophages to divert adipocyte-derived mitochondria into the blood
.
Cell Metab.
34
,
1499
1513.e8
123
Vincent
,
A.E.
,
Turnbull
,
D.M.
,
Eisner
,
V.
,
Hajnoczky
,
G.
and
Picard
,
M.
(
2017
)
Mitochondrial nanotunnels
.
Trends Cell Biol.
27
,
787
799
124
Picard
,
M.
,
McManus
,
M.J.
,
Csordas
,
G.
,
Varnai
,
P.
,
Dorn
, II,
G.W.
,
Williams
,
D.
et al (
2015
)
Trans-mitochondrial coordination of cristae at regulated membrane junctions
.
Nat. Commun.
6
,
6259
125
Taguchi
,
N.
,
Ishihara
,
N.
,
Jofuku
,
A.
,
Oka
,
T.
and
Mihara
,
K.
(
2007
)
Mitotic phosphorylation of dynamin-related GTPase Drp1 participates in mitochondrial fission
.
J. Biol. Chem.
282
,
11521
11529
126
Mukherjee
,
A.
,
Patra
,
U.
,
Bhowmick
,
R.
and
Chawla-Sarkar
,
M.
(
2018
)
Rotaviral nonstructural protein 4 triggers dynamin-related protein 1-dependent mitochondrial fragmentation during infection
.
Cell Microbiol.
20
,
e12831
127
Lu
,
Y.T.
,
Li
,
L.Z.
,
Yang
,
Y.L.
,
Yin
,
X.
,
Liu
,
Q.
,
Zhang
,
L.
et al (
2018
)
Succinate induces aberrant mitochondrial fission in cardiomyocytes through GPR91 signaling
.
Cell Death Dis.
9
,
672
128
Kashatus
,
J.A.
,
Nascimento
,
A.
,
Myers
,
L.J.
,
Sher
,
A.
,
Byrne
,
F.L.
,
Hoehn
,
K.L.
et al (
2015
)
Erk2 phosphorylation of Drp1 promotes mitochondrial fission and MAPK-driven tumor growth
.
Mol. Cell
57
,
537
551
129
Roe
,
A.J.
and
Qi
,
X.
(
2018
)
Drp1 phosphorylation by MAPK1 causes mitochondrial dysfunction in cell culture model of Huntington's disease
.
Biochem. Biophys. Res. Commun.
496
,
706
711
130
Han
,
H.
,
Tan
,
J.
,
Wang
,
R.
,
Wan
,
H.
,
He
,
Y.
,
Yan
,
X.
et al (
2020
)
PINK1 phosphorylates Drp1(S616) to regulate mitophagy-independent mitochondrial dynamics
.
EMBO Rep.
21
,
e48686
131
Gao
,
Q.
,
Tian
,
R.
,
Han
,
H.
,
Slone
,
J.
,
Wang
,
C.
,
Ke
,
X.
et al (
2022
)
PINK1-mediated drp1(S616) phosphorylation modulates synaptic development and plasticity via promoting mitochondrial fission
.
Signal. Transduct. Target. Ther.
7
,
103
132
Xie
,
Q.
,
Wu
,
Q.
,
Horbinski
,
C.M.
,
Flavahan
,
W.A.
,
Yang
,
K.
,
Zhou
,
W.
et al (
2015
)
Mitochondrial control by DRP1 in brain tumor initiating cells
.
Nat. Neurosci.
18
,
501
510
133
Rong
,
R.
,
Xia
,
X.
,
Peng
,
H.
,
Li
,
H.
,
You
,
M.
,
Liang
,
Z.
et al (
2020
)
Cdk5-mediated Drp1 phosphorylation drives mitochondrial defects and neuronal apoptosis in radiation-induced optic neuropathy
.
Cell Death Dis.
11
,
720
134
Cho
,
B.
,
Cho
,
H.M.
,
Kim
,
H.J.
,
Jeong
,
J.
,
Park
,
S.K.
,
Hwang
,
E.M.
et al (
2014
)
CDK5-dependent inhibitory phosphorylation of Drp1 during neuronal maturation
.
Exp. Mol. Med.
46
,
e105
135
Xu
,
S.
,
Wang
,
P.
,
Zhang
,
H.
,
Gong
,
G.
,
Gutierrez Cortes
,
N.
,
Zhu
,
W.
et al (
2016
)
CaMKII induces permeability transition through Drp1 phosphorylation during chronic beta-AR stimulation
.
Nat. Commun.
7
,
13189
136
Chen
,
S.
,
Liu
,
S.
,
Wang
,
J.
,
Wu
,
Q.
,
Wang
,
A.
,
Guan
,
H.
et al (
2020
)
TBK1-mediated DRP1 targeting confers nucleic acid sensing to reprogram mitochondrial dynamics and physiology
.
Mol. Cell
80
,
810
827.e7
137
Cribbs
,
J.T.
and
Strack
,
S.
(
2007
)
Reversible phosphorylation of Drp1 by cyclic AMP-dependent protein kinase and calcineurin regulates mitochondrial fission and cell death
.
EMBO Rep.
8
,
939
944
138
Han
,
X.J.
,
Lu
,
Y.F.
,
Li
,
S.A.
,
Kaitsuka
,
T.
,
Sato
,
Y.
,
Tomizawa
,
K.
et al (
2008
)
Cam kinase I alpha-induced phosphorylation of Drp1 regulates mitochondrial morphology
.
J. Cell Biol.
182
,
573
585
139
Su
,
Y.C.
and
Qi
,
X.
(
2013
)
Inhibition of excessive mitochondrial fission reduced aberrant autophagy and neuronal damage caused by LRRK2 G2019S mutation
.
Hum. Mol. Genet.
22
,
4545
4561
140
Merrill
,
R.A.
,
Slupe
,
A.M.
and
Strack
,
S.
(
2013
)
N-terminal phosphorylation of protein phosphatase 2A/Bbeta2 regulates translocation to mitochondria, dynamin-related protein 1 dephosphorylation, and neuronal survival
.
FEBS J.
280
,
662
673
141
Cho
,
D.H.
,
Nakamura
,
T.
,
Fang
,
J.
,
Cieplak
,
P.
,
Godzik
,
A.
,
Gu
,
Z.
et al (
2009
)
S-nitrosylation of Drp1 mediates beta-amyloid-related mitochondrial fission and neuronal injury
.
Science
324
,
102
105
142
Lee
,
D.S.
and
Kim
,
J.E.
(
2018
)
PDI-mediated S-nitrosylation of DRP1 facilitates DRP1-S616 phosphorylation and mitochondrial fission in CA1 neurons
.
Cell Death Dis.
9
,
869
143
Braschi
,
E.
,
Zunino
,
R.
and
McBride
,
H.M.
(
2009
)
MAPL is a new mitochondrial SUMO E3 ligase that regulates mitochondrial fission
.
EMBO Rep.
10
,
748
754
144
Zunino
,
R.
,
Schauss
,
A.
,
Rippstein
,
P.
,
Andrade-Navarro
,
M.
and
McBride
,
H.M.
(
2007
)
The SUMO protease SENP5 is required to maintain mitochondrial morphology and function
.
J. Cell Sci.
120
,
1178
1188
145
Guo
,
C.
,
Hildick
,
K.L.
,
Luo
,
J.
,
Dearden
,
L.
,
Wilkinson
,
K.A.
and
Henley
,
J.M.
(
2013
)
SENP3-mediated deSUMOylation of dynamin-related protein 1 promotes cell death following ischaemia
.
EMBO J.
32
,
1514
1528
146
Guo
,
C.
,
Wilkinson
,
K.A.
,
Evans
,
A.J.
,
Rubin
,
P.P.
and
Henley
,
J.M.
(
2017
)
SENP3-mediated deSUMOylation of Drp1 facilitates interaction with Mff to promote cell death
.
Sci. Rep.
7
,
43811
147
Karbowski
,
M.
,
Neutzner
,
A.
and
Youle
,
R.J.
(
2007
)
The mitochondrial E3 ubiquitin ligase MARCH5 is required for Drp1 dependent mitochondrial division
.
J. Cell Biol.
178
,
71
84
148
Nakamura
,
N.
,
Kimura
,
Y.
,
Tokuda
,
M.
,
Honda
,
S.
and
Hirose
,
S.
(
2006
)
MARCH-V is a novel mitofusin 2- and Drp1-binding protein able to change mitochondrial morphology
.
EMBO Rep.
7
,
1019
1022
149
Yonashiro
,
R.
,
Ishido
,
S.
,
Kyo
,
S.
,
Fukuda
,
T.
,
Goto
,
E.
,
Matsuki
,
Y.
et al (
2006
)
A novel mitochondrial ubiquitin ligase plays a critical role in mitochondrial dynamics
.
EMBO J.
25
,
3618
3626
150
Hu
,
Q.
,
Zhang
,
H.
,
Gutierrez Cortes
,
N.
,
Wu
,
D.
,
Wang
,
P.
,
Zhang
,
J.
et al (
2020
)
Increased Drp1 acetylation by lipid overload induces cardiomyocyte death and heart dysfunction
.
Circ. Res.
126
,
456
470
151
Shen
,
Y.L.
,
Shi
,
Y.Z.
,
Chen
,
G.G.
,
Wang
,
L.L.
,
Zheng
,
M.Z.
,
Jin
,
H.F.
et al (
2018
)
TNF-alpha induces Drp1-mediated mitochondrial fragmentation during inflammatory cardiomyocyte injury
.
Int. J. Mol. Med.
41
,
2317
2327
152
Lum
,
M.
and
Morona
,
R.
(
2014
)
Dynamin-related protein Drp1 and mitochondria are important for Shigella flexneri infection
.
Int. J. Med. Microbiol.
304
,
530
541
153
Chowdhury
,
S.R.
,
Reimer
,
A.
,
Sharan
,
M.
,
Kozjak-Pavlovic
,
V.
,
Eulalio
,
A.
,
Prusty
,
B.K.
et al (
2017
)
Chlamydia preserves the mitochondrial network necessary for replication via microRNA-dependent inhibition of fission
.
J. Cell Biol.
216
,
1071
1089
154
Suzuki
,
M.
,
Danilchanka
,
O.
and
Mekalanos
,
J.J.
(
2014
)
Vibrio cholerae T3SS effector VopE modulates mitochondrial dynamics and innate immune signaling by targeting Miro GTPases
.
Cell Host Microbe
16
,
581
591
155
Stavru
,
F.
,
Palmer
,
A.E.
,
Wang
,
C.
,
Youle
,
R.J.
and
Cossart
,
P.
(
2013
)
Atypical mitochondrial fission upon bacterial infection
.
Proc. Natl Acad. Sci. U.S.A.
110
,
16003
16008
156
Jain
,
P.
,
Luo
,
Z.Q.
and
Blanke
,
S.R.
(
2011
)
Helicobacter pylori vacuolating cytotoxin A (VacA) engages the mitochondrial fission machinery to induce host cell death
.
Proc. Natl Acad. Sci. U.S.A.
108
,
16032
16037
157
Chatel-Chaix
,
L.
,
Cortese
,
M.
,
Romero-Brey
,
I.
,
Bender
,
S.
,
Neufeldt
,
C.J.
,
Fischl
,
W.
et al (
2016
)
Dengue virus perturbs mitochondrial morphodynamics to dampen innate immune responses
.
Cell Host Microbe
20
,
342
356
158
Barbier
,
V.
,
Lang
,
D.
,
Valois
,
S.
,
Rothman
,
A.L.
and
Medin
,
C.L.
(
2017
)
Dengue virus induces mitochondrial elongation through impairment of Drp1-triggered mitochondrial fission
.
Virology
500
,
149
160
159
Castanier
,
C.
,
Garcin
,
D.
,
Vazquez
,
A.
and
Arnoult
,
D.
(
2010
)
Mitochondrial dynamics regulate the RIG-I-like receptor antiviral pathway
.
EMBO Rep.
11
,
133
138
160
Keck
,
F.
,
Brooks-Faulconer
,
T.
,
Lark
,
T.
,
Ravishankar
,
P.
,
Bailey
,
C.
,
Salvador-Morales
,
C.
et al (
2017
)
Altered mitochondrial dynamics as a consequence of Venezuelan equine encephalitis virus infection
.
Virulence
8
,
1849
1866
161
Pal
,
A.D.
,
Basak
,
N.P.
,
Banerjee
,
A.S.
and
Banerjee
,
S.
(
2014
)
Epstein-Barr virus latent membrane protein-2A alters mitochondrial dynamics promoting cellular migration mediated by Notch signaling pathway
.
Carcinogenesis
35
,
1592
1601
162
Shi
,
C.S.
,
Qi
,
H.Y.
,
Boularan
,
C.
,
Huang
,
N.N.
,
Abu-Asab
,
M.
,
Shelhamer
,
J.H.
et al (
2014
)
SARS-coronavirus open reading frame-9b suppresses innate immunity by targeting mitochondria and the MAVS/TRAF3/TRAF6 signalosome
.
J. Immunol.
193
,
3080
3089
163
Varga
,
Z.T.
,
Grant
,
A.
,
Manicassamy
,
B.
and
Palese
,
P.
(
2012
)
Influenza virus protein PB1-F2 inhibits the induction of type I interferon by binding to MAVS and decreasing mitochondrial membrane potential
.
J. Virol.
86
,
8359
8366
164
Yoshizumi
,
T.
,
Ichinohe
,
T.
,
Sasaki
,
O.
,
Otera
,
H.
,
Kawabata
,
S.
,
Mihara
,
K.
et al (
2014
)
Influenza A virus protein PB1-F2 translocates into mitochondria via Tom40 channels and impairs innate immunity
.
Nat. Commun.
5
,
4713
165
Kim
,
S.J.
,
Khan
,
M.
,
Quan
,
J.
,
Till
,
A.
,
Subramani
,
S.
and
Siddiqui
,
A.
(
2013
)
Hepatitis B virus disrupts mitochondrial dynamics: induces fission and mitophagy to attenuate apoptosis
.
PLoS Pathog.
9
,
e1003722
166
Kim
,
S.J.
,
Syed
,
G.H.
,
Khan
,
M.
,
Chiu
,
W.W.
,
Sohail
,
M.A.
,
Gish
,
R.G.
et al (
2014
)
Hepatitis C virus triggers mitochondrial fission and attenuates apoptosis to promote viral persistence
.
Proc. Natl Acad. Sci. U.S.A.
111
,
6413
6418
This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY). Open access for this article was enabled by the participation of the University of Queensland in an all-inclusive Read & Publish agreement with Portland Press and the Biochemical Society under a transformative agreement with CAUL.