Next Article in Journal
Additive Role of Immune System Infiltration and Angiogenesis in Uveal Melanoma Progression
Next Article in Special Issue
Macrophage Polarization in Cardiac Tissue Repair Following Myocardial Infarction
Previous Article in Journal
Spinal Inhibitory Interneurons: Gatekeepers of Sensorimotor Pathways
Previous Article in Special Issue
The Cross-Talk between Mesenchymal Stem Cells and Immune Cells in Tissue Repair and Regeneration
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Mitochondria in Immune-Cell-Mediated Tissue Regeneration and Ageing

1
Department of Internal Medicine, Kaohsiung Chang Gung Memorial Hospital, Kaohsiung 833, Taiwan
2
College of Medicine, Chang Gung University, Kaohsiung 833, Taiwan
3
Center for Mitochondrial Research and Medicine, Kaohsiung Chang Gung Memorial Hospital, 123, Dapi Road, Niaosong District, Kaohsiung 833, Taiwan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(5), 2668; https://doi.org/10.3390/ijms22052668
Submission received: 15 February 2021 / Accepted: 3 March 2021 / Published: 6 March 2021
(This article belongs to the Special Issue Immune Cell Mediated Tissue Regeneration)

Abstract

:
During tissue injury events, the innate immune system responds immediately to alarms sent from the injured cells, and the adaptive immune system subsequently joins in the inflammatory reaction. The control mechanism of each immune reaction relies on the orchestration of different types of T cells and the activators, antigen-presenting cells, co-stimulatory molecules, and cytokines. Mitochondria are an intracellular signaling organelle and energy plant, which supply the energy requirement of the immune system and maintain the system activation with the production of reactive oxygen species (ROS). Extracellular mitochondria can elicit regenerative effects or serve as an activator of the immune cells to eliminate the damaged cells. Recent clarification of the cytosolic escape of mitochondrial DNA triggering innate immunity underscores the pivotal role of mitochondria in inflammation-related diseases. Human mesenchymal stem cells could transfer mitochondria through nanotubular structures to defective mitochondrial DNA cells. In recent years, mitochondrial therapy has shown promise in treating heart ischemic events, Parkinson’s disease, and fulminating hepatitis. Taken together, these results emphasize the emerging role of mitochondria in immune-cell-mediated tissue regeneration and ageing.

1. Introduction

Organ injuries require tissue repair and regeneration, while subtle tissue injury also leads to macrophage activation, and fibroblast or stromal cell proliferation, such as plaque formation in atherosclerosis. Three phases are involved in the healing process: modifying the inflammatory process, cell proliferation, and remodeling the extracellular matrix. Tissue inflammation is clearly defined by the presence of either inflammatory cells or cytokines in the tissue. However, the differences between tissue repair, tissue regeneration, and tissue proliferation are subtle. Natural tissue repair occurs after tissue inflammation or damage, and the transition from inflammation to repair occurs automatically and sequentially. Importantly, the switch between wound repair, and the regeneration and inflammation pathways lead to differing disease outcomes [1]. In general, tissue repair and tissue regeneration are different steps of a continuum. Tissue repair is an early stage of functional restoration, but lacks patterned three-dimensional tissue reconstruction, which is expressed in tissue regeneration [2]. However, unhindered tissue regeneration, occurring under unknown mechanisms, leads to tissue proliferation, which could further cause unwanted insults, such as fibrosis in ageing.
Both innate and adaptive immunity against environmental stimulus needs energy supplies. The interaction between immune activation and the energy supplies is based on immuno-metabolomics. In short, the energy requirements for fast responses, such as the neutrophils, the dendritic cell, or the macrophages, mostly come from available glucose, i.e., glycolytic response, or glycogen storage inside the cell [3]. After the first rapid response finishes, the immune systems require energy supplies for further continuous reactions, such as differentiations, transformation, phagocytosis, or killing extracellular pathogens. At the same time, the energy supply mostly shifts from the glycolysis toward either the tricarboxylic acid (TCA) cycle or fatty acid beta-oxidation [3,4].
The mitochondrion has an indispensable role in the development and function of immune cells, from energy supply to the activation, proliferation, and phagocytosis, and ending with repairs and regeneration. Besides energy supply, with the help of mitochondrial reactive oxygen species (ROS), the immune cells execute complex functions, such as apoptosis, pyroptosis, netosis, and adaptive immunity activation. The mitochondrial ROS helps T cell activation and B cell antibody production [3,5].
In this review article, we focus on the role of immune cells in various tissue stages, from inflammation, through repair and regeneration, to tissue proliferation, with an emphasis on the emerging role of mitochondria in immune-cell-mediated tissue regeneration and ageing.

2. Immune Cells in Tissues under a Homeostasis Condition

In homeostasis, the immune cells’ continued surveillance involves the innate immunity and adaptive immunity systems working collaboratively to maintain balance between environmental stimuli, and prevent unwanted pathogen invasion. Collectively, innate immunity involves the stromal cells forming the skin barrier, endothelial cells forming the vessel barrier, and epithelial cells forming the mucosal barrier, as well as the leukocytes, natural killer cells, dendritic cells, and macrophages. Meanwhile, adaptive immunity including T cells and B cells is on regular immune surveillance, with the plasma B cells producing immunoglobulin to counter invading pathogens, and the memory T cells remaining dormant, awaiting stimulation to trigger activation. The crosstalk between innate immunity and adaptive immunity depends on the antigen-presenting cells, such as dendritic cells or macrophages, presenting constant signals to T cells. If the antigen-presenting cells present self-antigen to T cells, this causes the T cells to become anergic, thereby becoming regulatory T cells (Treg). Meanwhile, the cytokine profiles are more likely to be anti-inflammatory cytokines, such as IL-4, IL-10, or TGF-beta. The immune cells, either innate or adaptive immunity cells, are capable of moving outside the vessels and rolling toward the homing location according to cytokine attraction, or the adhesion molecules’ traction. In general, under the homeostasis condition, all players in the immune system, including Th1, Th2, Th17 cells, myeloid cells, and B cells, interact with normal pathogens without clinical inflammation [6].
The size, morphology, and number of mitochondria inside the cell are important clues to immune cell activities and functions. Generally speaking, the enzymes, either from the TCA cycle or from the beta-oxidation, may be a clue to mitochondrial activity elevations in cells. Different subtypes of immune cells favor different energy resources, and the resources of energy can be diverse within each cell type. Take macrophages for example, the M1 cells use not only glycolysis, but also the pentose phosphate pathway, while their TCA cycle is impaired, while M2 and mature T cells are thought to be dependent on β-oxidation associated with an increased TCA, increased respiratory capacity, and increased production of ATP by oxidative phosphorylation [7,8]. Mitochondrial dynamics, and fusion and fission events also affect immune function. Mitochondrial dynamics mostly affect the subsequent adaptive immune reaction rather than the initiation of innate immunity. In 2016, Buck et al. demonstrated that mitochondrial fusion is associated with memory T cell formation, and improves the T cell function against tumors [9]. Besides adaptive immunity, macrophages use mitofusin 2 protein against RNA virus infection [10]. In a cybrid model of diabetes, our group studied the genetic effect of mitochondria in cells with the same nuclear genomic background, and demonstrated the role of mitochondrial dynamics in regulating immunity and mitochondrial ROS [11,12]. These data exemplified that mitochondrial dynamics have a role in linking cellular immunity and energetics.

3. Immune Cells under Inflammatory and/or Injury Conditions

During tissue injury, the immediate response immune cells (innate immunity) respond to the alarms sent from injured cells in differing ways; with the adaptive immunity mechanisms subsequently joining the inflammatory reaction. The overall scope of inflammation depends on the extent of injury. The control mechanism of each immune reaction relies on the orchestration of different types of helper T cells and the activators, signal 1, 2, and 3, antigen-presenting cells, co-stimulatory molecules, and cytokines. After helper T cells (Th) are activated, either Th1, Th2, or Th17 operate and summon innate immunity cells to counteract the invasive pathogens [13,14]. The B cells are also activated, and produce relevant antibodies to combat the antigen, and through the mechanisms of isotype switch, and affinity maturation, the binding of the antibody to the antigen becomes more specific with time. To illustrate this, in lupus nephritis, the immune complex deposition on the glomerular of the kidneys can activate innate and adaptive immunity mechanisms to join and clear the battlefield. Either action could lead to organ inflammation and tissue damage, and ensuing proteinuria or hematuria, which is called glomerulonephritis. Before and during clinical evidence of glomerulonephritis, the tissue damage has already started, and necessary tissue repair is underway. The M1 macrophage is a pro-inflammatory cell that secretes pro-inflammatory cytokines, such as interferon-γ (IFN-γ), interleukins (IL) (IL-1, IL-6, IL-12, IL-23), and tumor necrosis factor-α (TNF-α) [15,16,17]; while the M2 macrophage is believed to be an anti-inflammatory phenotype [18], dependent on several other subsets of anti-inflammatory cytokines, including IL-4 and IL-13 [19]. Our heretofore unpublished data demonstrated that the macrophage phenotype had shifted from M1 type to M2 type in some lupus nephritis patients under partial treatment. Although the precise role of M1 macrophages in the inflammation, and M2 macrophages in the repair or regeneration remain unclear, the M1 to M2 switch indicates the initiation of the repair or regeneration in a dynamic condition [1].
Tissue regeneration is partially dependent on macrophages [20], regulatory T cells [21], and other cells, but primarily on stem cells [2,22,23,24]. Different organ systems have their own highly specified functions, and the necessary tissue regeneration is thus heavily regulated. Some specified molecules, such as galectins and cytokines, can either positively or negatively regulate inflammation, regeneration, and ageing in different organ systems [25] (Table 1, Figure 1). In heart failure (HF) conditions, elevation of cardiac damage markers (troponins or natriuretic peptides) have been found useful in the diagnosis of HF and in outcome prediction [26]. In the CORONA and COACH trials, repeated measurements of galectin-3 level, a protein involved in myocardial fibrosis and remodeling, provided a important and significant prognostic value for identifying HF [26]. In atherosclerosis plaque, loss of galectin-3 in macrophage has been associated with increased expression of pro-inflammatory genes, including matrix metalloproteinase 12 (MMP-12), chemokine ligand 2 (CCL2), prostaglandin-endoperoxide synthase 2 (PTGS2), and interleukin-6 (IL-6), while the expression of transforming growth factor-β1 (TGFβ1) was reduced, suggesting a prominent protective role of galectin-3 in regulating macrophage polarization and plaque progression [25]. In animal experiments of acute renal failure, galectin-3 expression was markedly up-regulated [27], and galectin-3 from macrophages is considered as a major mechanism in renal fibrosis [22]. Furthermore, galectin-3 has been studied in several models of chronic kidney diseases, including nephropathies induced by ageing, ischemia, diabetes, and chronic allograft injury [24,28]. In skin, galectin-3 facilitates intracellular epidermal growth factor receptor (EGFR) trafficking to the surface, which could promote wound re-epithelialization [23].

4. Immunity and the Ageing Condition

Ageing in the immune system occurs in a similar manner as in other organ systems, and several phenotypic and molecular hallmarks of ageing are exhibited in immune cells, including decreased MHC II expression, decreased antibody production, altered toll-like receptor expression, limited diversity in B cell receptor repertoire, and restricted diversity in the T cell receptor repertoire [41]. Recently, the Baltimore Longitudinal Study on Ageing compared CD4+ T cells obtained from younger (20–39 years-old) and older (>70 years-old) healthy participants, demonstrating that mitochondrial respiration was impaired in CD4+ T cells from older subjects. In older age groups, dysfunction of autophagy leads to defective mitochondrial turnover, triggering chronic inflammation and impairment of the immune defense system [42]. While the T cell ageing process is overwhelming, the antigen–antibody responses will be compromised, which leads to so-called immunosenescence. Immunosenescence includes increases in unnecessary terminal-differentiated memory cell populations and altered cytokine production [43]. Senescent cells are characterized by high expressions of cell-cycle inhibitors, including p16, p21, and p53, coupled with a persistent DNA damage response, which can eventually induce a pro-inflammatory state, known as senescence-associated secretory phenotype (SASP), involving interleukin-6 (IL-6), IL-8, monocyte chemoattractant protein-1 (MCP-1), IL-1β, plasminogen activator inhibitor-1 (PAI-1), and high mobility group box 1 (HMGB-1) [44]. In one large cohort study by Piber et al. of the relationship between ageing and inflammation-associated immune biomarkers [45], significant correlations between the age of the blood donor and the serological markers, IL-6 and soluble TNF receptor II were identified. In addition, intracellular signal transducer and activator of transcription (STAT) signaling, which is associated with immunity dysfunction, including STAT1, STAT3, STAT5, was also found to be significantly correlated with age.
Pinti et al. demonstrated that mitochondrial DNA (mtDNA) plasma levels increased gradually after the fifth decade of life, and there was a role for familiar/genetic background in controlling the levels of circulating mtDNA. The circulated mtDNA was associated with inflammatory cytokines, TNF-α, IL-6, RANTES, and IL-1ra [46]. Furthermore, different human populations have acquired sequence polymorphisms in mtDNA, which are maternally inherited genetic haplogroups [47]. The association of mtDNA haplogroups with altered disease susceptibility and longevity has previously been reported [47,48,49,50,51].
During the natural ageing process, mitochondrial function declines due to accumulation of damaged mitochondrial genomes resulting from localized ROS exposure, and chronic, sterile, low-grade inflammation develops, contributing to the pathogenesis of the metabolic syndromes, obesity, and diabetes mellitus (DM). Inflammageing is the result of chronic physiological stimulation of the innate immune system [52]. Recent investigation of innate immunity has revealed insights into host reactions to noninfectious diseases and the inflammatory modulation of basic cellular metabolic processes [53]. Nutrient excess triggers inflammation in metabolic-related pathways (meta-flammation) through provoked innate immunity [54,55]. Among the meta-flammation disorders, abdominal obesity and diabetes are among the main pathologies associated with the increased burden of cellular senescence [56], and are clustered with a group of risk factors related to cardiovascular diseases [57]. As toll-like receptor 4 (TLR4) plays an important role in innate immunity, we investigated the effect of TLR4 knockout (T4KO) in a high-fat high-sucrose diet-induced obesity mouse model. In a nutrient excess condition, T4KO mice, as compared with wild type mice, had increased body weight and body fat, but decreased oxidative damage, M1/M2 macrophage ratio, chronic inflammation, and insulin resistance. This protective effect is mediated through metabolic reprogramming of mitochondria in adipose tissue. The study provides critical insight into linking innate immunity–mitochondria interplay with prevention of diabetes [58].
Insulin resistance status leads to elevated blood sugar and production of advanced glycation end products [59]. Among the advanced glycation end products, the carboxymethyl lysine glycation end products may be found in obese tissue, causing dysregulation of adipokine expression [60]. Hence, ageing could lead to a vicious cycle, which starts with immunosenescence, through the status of insulin resistance, and leads to advanced glycation end products being aggregated, and a return to inflammation status. Another phenomenon in the immune system noted in ageing is increased myeloid hematopoiesis [61]. Such increased myeloid hematopoiesis and the relatively decreased responses of adaptive immunity causes the imbalanced immunity to tend toward immunosenescence status. The production of inflammatory cytokines IL-1, IL-6, and TNF-α from the myeloid lineages accumulate. The so-called myeloid-derived suppressor cells will also increase in proportion to the increased myeloid hematopoiesis, and thus further suppressing T cell production and proliferation [61].

5. Mitochondrial DNA and Double Strand RNA Release and the Innate Immune Reaction

Mitochondria, as the cellular power plant, can function as signaling platforms to regulate cellular growth and death [62]. Mitochondria assist cells to maintain homeostasis and enable their adaptation to various stressors [63]. In addition to bioenergetic and biosynthetic functions, mitochondria are increasingly recognized as a central trigger of innate immunity that influences anti-microbial responses, autoimmune diseases, metabolic disorders, and cancers [64]. Innate pattern recognition receptors, mainly toll-like receptor 2 (TLR2) and toll-like receptor 4 (TLR4), and the nucleotide oligomerization domain (NOD) receptor protein NLRP3, trigger the activation of downstream inflammatory cascades through mitochondrial ROS [65,66,67,68].
Mitochondria are able to respond to cellular damage [69], through which they may accumulate oxidative stress and become irreversibly affected. Mitochondrial DNA can be released by pro-apoptotic, apoptotic, or necrotic cells, and becoming cell-free circulating mitochondrial DNA (ccf-mtDNA) [70,71]. The mtDNA can be actively secreted as ccf-mtDNA [72,73] or transported in extracellular vesicles (EVs) in response to cell stress [74]. Recent clarification of cytosolic escape of mtDNA, and that mitochondrial double strand RNA (mtdsRNA) triggers innate immunity further underscores the pivotal role of mitochondria in inflammation-related diseases [75,76]. The accumulation of mtDNA damage, as has been seen in patients with sepsis, cancer, myocardial infarction, diabetes, post-surgical intervention, and severe trauma, activates innate immune responses via different pathways, and ccf-mtDNA levels could possibly be used as a predictor of outcome in the above-mentioned diseases [77,78,79,80].
Collins et al. [81] were the first to report the immunostimulatory potential of mtDNA in 2004, noting that mtDNA when added to mouse splenocytes elicited TNF-α, and induced arthritis when it was released in the joints of mice. A number of other studies have shown that mtDNA can directly engage pattern-recognition receptors (PRR) of the innate immune system to enhance pro-inflammatory responses [82]. MtDNA can be recognized by three important pattern recognition receptors (PRRs) of the innate immune system, including TLR9, cytosolic inflammasomes, and type I interferon response, triggering a pro-inflammatory response [83,84]. TLR9 is a DNA-sensing PRR, and it is able to detect hypomethylated CpG motifs of DNA. MtDNA, due to its similarities with bacteria DNA, can mediate a pro-inflammatory response dependent on the TLR9 pathway [85]. In a model of shock, mtDNA was released by shock, and activated neutrophils via p38 MAPK pathway and triggered the release of MMP-8 and MMP-9 via TLR9. Furthermore, C57BL/6 mice challenged intratracheally with mtDNA induced a local inflammatory response, producing proinflammatory cytokines and activating p38 MAPK via TLR9, accompanied by macrophage infiltration [86]. Moreover, mtDNA is involved in trauma-induced neutrophil extracellular trap (NETs) formation, which is an extracellular structure secreted by neutrophils, and able to bind and kill microorganisms. MtDNA is able to induce NET formation in human neutrophils via TLR9 [87]. Plasma ccf-mtDNA activates the TLR9-mediated signaling pathway in circulating neutrophils, releasing tumor necrosis factor (TNF), interleukin-6 (IL-6), and adhesion molecules.
In addition to the extracellular mtDNA, in the intracellular space, mitochondrial damage is a critical step in inflammasome activation. There is a potential role for mtDNA serving as a direct ligand for NLRP3 and AIM2 inflammasome [88]. Zhong et al. reported that newly synthesized mtDNA and its cytosolic release play a key role in triggering NLRP3 inflammasome, a central mechanism in chronic inflammatory and degenerative diseases [76]. TLR4 engagement triggers MyD88/TRIF-dependent signaling that activates IRF1 to induce the expression of mitochondrial UMP-CMP kinase 2 (CMPK2), a mitochondrial nucleotide kinase, a rate-limiting enzyme for de novo mtDNA synthesis. This event is linked to the new synthesis of mtDNA, and this freshly generated mtDNA is thought to result in the production of oxidized mtDNA (ox-mtDNA) fragments. Subsequently, ox-mtDNA exits the mitochondria via mitochondrial pore transition (MPT), which opens after exposure to NLRP3 activators. This leads to the production of inflammatory proteins, such as interleukin one beta (IL-1β). In addition, mitochondrial apoptotic signaling causes the release of mtDNA and triggers the cytosolic DNA sensor cyclic GMP-AMP synthase (cGAS)-stimulator of interferon genes (STING) to drive interferon responses [89]. Once type I IFNs are secreted, a complex web of host defenses will be induced, with the expression of hundreds of interferon-stimulated genes [90].
Dhir et al. demonstrated mtdsRNA as a source of self-nucleic acids initiating an interferon-related innate immune response via the cytoplasmic dsRNA sensor melanoma differentiation-associated gene 5 (MDA5) and mitochondrial anti-viral signaling protein (MAVS) in the context of perturbed RNA metabolism [75]. For example, upon sensing cytoplasmic viral RNA, the retinoic-acid-inducible protein I (RIG-I)-like receptor family (RLRs), RIG-I and MDA5, interact via the outer mitochondrial membrane adaptor protein MAVS, leading to activation of downstream pathways, such as interferon regulatory factor 3 (IRF3), MAP kinases, and nuclear factor-κB (NF-κB) [91]. Clinically, free circulating mtDNA has been considered as a danger-associated molecular pattern in assessing the prognosis for myocardial infarction [92]. Our previous studies demonstrated that the disease activity of systemic lupus erythematosus (SLE) is positively associated with the mitochondrial membrane APO2.7 level of CD19+ cells, but negatively associated with MAVS and caspase-9 levels, which collectively indicate a mitochondrial pathway [93].

6. The Role of Mitochondria in Regeneration

Mitochondria are intracellular energy plants and signaling organelles. Importantly, current research is further elucidating the roles of extracellular mitochondria. Forms of extracellular mitochondria can be found free, enclosed by a membrane, such as inside the platelets or vesicles, or as ccf-mtDNA. When existing outside cells, they can elicit regenerative effects, or serve as an activator of immune response [94,95]. Platelets, carrying mitochondria, perform complex activities such as tissue repair [96,97,98,99]. The major pathways reported during megakaryocytes maturation are the JAK-STAT, PI3K-AKT, MAPK, and Wnt pathways, all of which are related to mitochondrial biogenesis [100]. The Wnt signaling cascade, in particular, is a key driver of tissue morphogenesis and architecture, which has been identified as a crucial factor for pluripotent capacities [101].
Mitochondria inside platelets not only produce energy but also regulate thrombus formation via equilibrium between ROS production [102,103,104]. It has been observed that the malfunction of proteins related to mitochondrial bioenergetic function has been linked to the severity of Type 2 diabetes (T2DM), sepsis, Parkinson’s disease, thrombocytopenia, and Alzheimer’s disease [105,106,107]. Overcoming multiple immune dysfunctions is fundamental for the treatment of Type 1 diabetes (T1DM), and the neogenesis of pancreatic islet β cell is a key factor in the treatment for T1DM and long-term T2DM. Previous studies have established that platelet-derived mitochondria not only exert a function in immune tolerance, but also carry the islet-specific transcription factor (MAFA), the pancreatic progenitor-associated marker (SOX9), and embryonic stem cell-related self-renewal markers (including OCT4, SOX2, KLF4, and C-MYC) [94,108]. Zhao et al. showed that mitochondria released by platelets can be immune-regulatory and have regenerative properties in diabetic patients after receiving stem cell educator therapy [109]. Platelets and their released mitochondria, after receiving stem cell educator therapy, expressed immune tolerance-related markers, and can modulate the proliferation and function of immune cells. These mitochondria also displayed embryonic stem cell- and islet β-cell-associated markers, leading to the improvement of islet β-cell functions; more specifically, there was an expression of an immune checkpoint receptor PD-1 (programmed death receptor-1) on CD4+ and CD8+ T cells after treatment with mitochondria. Additionally, immunohistochemistry data demonstrated the migration of platelets into pancreatic islets of diabetic subjects. These platelets could heal the damaged islet β-cell [109].
In recent years, mitochondrial therapy has shown promising effects in treatments for heart ischemic events, Parkinson’s disease, and fulminating hepatitis [110,111,112]. It has been proposed that mitochondrial transplantation therapy could replace damaged mitochondria, thereby leading to the repair of heart tissue [73]. In 2006, Spees et al. first demonstrated that human mesenchymal stem cells (MSCs) could transfer mitochondria through nanotubular structures to recipient cells with defective mitochondrial DNA (mtDNA) [113]. Since then, bone marrow-derived MSCs [114,115,116] and adipose-derived MSCs [115,117] have been shown to exhibit the capability of transferring mitochondria, thereby rescuing cellular bioenergetics. Recent evidence suggests that the mitochondria isolated from muscle tissue do not exacerbate the immune response when transplanted in vitro or in vivo [118]. Once inside the cell, the exogenous mitochondria, depending on their membrane constituents, can fuse with the endogenous network, with fusion-associated proteins (e.g., MFN1, MFN2, and OPA1) [72,119]. Several pathologies affect mitochondrial function and structure, including sepsis, pancreatitis, and fulminant hepatitis, thus presenting potential candidates for mitochondrial therapy [94,110,111].
Our colleagues have conducted a series of studies demonstrating that umbilical cord Wharton’s jelly-derived MSCs are capable of conducting mitochondrial transplantation therapy by transferring healthy mitochondria to replenish lost mtDNA [120] and reducing the mutated-mtDNA burden, thus restoring mitochondrial function in cybrid cells carrying point mutation mt.8344A > G [102], and in MELAS fibroblasts carrying point mutation mt.3243A > G.

7. The Role of Mitochondria in Crosstalk between Cells

Mitochondria have been shown to be horizontally transferred between mammalian cells, with the incorporation of mitochondria into the original mitochondrial network of recipient cells, contributing to changes in the bioenergetic profile both in vitro and in vivo [121]. The transfer of mitochondria may also result in the initiation of stem cell differentiation, reprogramming cells, or activation of signaling pathways. Mitochondria exchange between cells can occur via tunneling nanotubes (TNTs), EVs, and cellular fusion [121]. TNT formation was reported in immune cells as well as in neurons, glial cells, and cardiomyocytes [122]. Recent evidence strongly supports that mitochondrial transfer does occur in vivo [74]. Astrocytes may release extracellular mitochondrial particles that enter neurons after stroke, and having a role in the neuroglial crosstalk that may contribute to endogenous neuroprotective and neural recovery mechanisms [123]. Regarding cancer, the mitochondria transfer is enhanced by some chemotherapies and confers a survival advantage to leukemic blasts and leukemia initiating cells [124]. Mitochondrial dysfunction is a key event in simulated ischemia/reperfusion (SI/R) injury. In vitro (SI/R) experiments demonstrated that bone marrow(BM)-derived mesenchymal stem cells rescue injured H9c2 cardiomyocytes, via transferring intact mitochondria through tunneling nanotubes [121]. Recent evidence has shown that in stress response, stromal cells can transfer healthy mitochondria to hematopoietic stem cells (HSC), facilitate HSC bioenergetics shift towards oxidative phosphorylation, and leukocyte expansion [125]. Mitochondria have emerged as a potential regulator of HSC fate and regenerative medicine.

8. mtDNA in Trauma and Organ Transplantation

In critically ill patients, mtDNA may be released into plasma by necrosis, apoptosis, or active ejection by leukocytes. Plasma mtDNA levels are associated with ARDS in trauma or sepsis patients [80,126]. Severe trauma results in high concentrations of mtDNA release from tissues with subsequent uncontrolled inflammation and end-organ damage, and post-injury complications [127,128,129,130,131,132,133,134]. Cells that migrate to sites of injury include leukocytes, platelets, and stem cells that all these cells can release ccf-mtDNA. The process of ccf-mtDNA in response to trauma initiates from an immediate surge of ccf-mtDNA released from injured cells into the extracellular space, which through circulation results in a secondary active release of ccf-mtDNA by immune cells and platelets, followed by release of inflammatory cytokines and end-organ damage [129,135].
The innate immune system is a critical regulator of the adaptive immune responses that lead to allograft rejection. Mitochondrial damage and mitochondrial metabolism is critical during organ implantation and reperfusion of ischemic tissue [136,137,138,139]. Mitochondria cause ischemia-reperfusion (IR) injury by generating damaging reactive oxygen species (ROS) upon reperfusion, which lead to the release of mitochondrial damage-associated molecular patterns (DAMPs), following activation of the innate immune response, contributing to organ rejection [140,141]. Increased extracellular mtDNA is associated with elevated levels of inflammatory cytokines, and early organ dysfunction in liver and kidney transplantation [142,143,144].
Lin et al. demonstrated that extracellular mitochondria are abundant in the circulation of deceased organ donors, which correlates with early allograft dysfunction. Extracellular mitochondria led to the upregulation of endothelial cell (EC) adhesion molecules, inflammatory cytokines and chemokines, and activated dendritic cells to upregulate costimulatory molecules. They concluded that ccf-mtDNA from organ donors may directly activate allograft ECs, and result in graft rejection in transplant recipients [145].

9. Targeting Mitochondria as a Novel Therapeutic Strategy

Our research indicates that a high-fat diet could lead to ROS expression, DNA and protein oxidative damage, and adipose tissue inflammation [146]; however, the inflammation could be resolved by mitochondrial manipulation, such as feeding mice with N-acetylcysteine. This phenomenon is better presented when the N-acetylcysteine is provided at an earlier stage, possibly due to the better flexibility of the immune reaction at a younger age, or related to the physiological effects of inflammageing [147].
As regulators of cellular energy homeostasis and cell death signaling, the continued integrity of mitochondria within a cell is crucial for its sustained health. With its discovery more than decades ago, metformin represented a milestone in the treatment of patients with T2DM. Mitochondria are the main subcellular targets of metformin, which accumulates selectively in mitochondria in many hundred-fold higher concentrations than those observed in the extracellular medium [148]. Recent evidence in humans indicated the novel pleiotropic actions of metformin, including antiproliferative, antifibrotic, and antioxidant effects [149], as well as potential anti-ageing [150], anticancer, and immune-modulatory features [151,152]. These effects are thought to occur through pathways that involve changes in the gut microbiota and both AMP-activated protein kinase-dependent and independent mechanisms [153].
Metformin decreased the LPS-induced production of the proinflammatory cytokine pro-IL-1β secretion in mouse bone marrow-derived macrophages, in agreement with previous findings in human monocyte-derived macrophages [154,155], and the mechanism is thought to result from the specific inhibition of mitochondrial ROS production driven by reverse electron transport at the mitochondrial respiratory chain complex 1 [154].
Wang et al. has demonstrated that metformin could control lupus activity via inhibiting the neutrophil extracellular trap formation and the production of anti-mitochondrial DNA autoantibodies [156]. In our in vitro transmitochondrial cybrid cell culture system, presence of the intracellular sensor of dsRNA, MDA5, could up-regulate MAVS and IRF3 in a high glucose condition with fatty acid, even without antigen-presenting cells. After the metformin was placed in the in vitro culture system, the levels of MDA5, MAVS, TBK1, and IRF3 were all significantly down-regulated [157]. Our study provided insight into the therapeutic activity of metformin in SLE patients. Further studies are required to elucidate the immune-mediated mode of action of metformin.
Glucagon-like peptide 1 receptor agonists (GLP-1RAs), initially designed to treat diabetes, have shown protective effects in several clinical trials in patients with Alzheimer’s or Parkinson’s disease. Post-treatment with GLP-1 analogues ameliorated mitochondrial dysfunction by increasing expression of Cox IV and SOD1, increasing HSP60 and PHB1, and increasing PDH expression [158]. A further study found that post-infarction treatment with GLP-1 receptor agonist, significantly reduced adverse LV remodeling and the decline of cardiac function. This paralleled a Parkin-mediated increase in autophagy, mitophagy, and mitochondrial biogenesis [159]. A previous study also showed that GLP-1RA activates the sirtuin pathway, and increases energy expenditure in human adipocytes, which may play a role in body weight reduction [160].
Although the benefits of mitochondrial transplantation therapy have been demonstrated, there are still questions surrounding the most effective methods to obtain and transfer/transplant the mitochondria. Optimizing techniques for detecting mtDNA, and mitochondria number or activity could have a positive impact in improving current clinical practices [94,161].

Author Contributions

P.-W.W. and S.-W.W. developed the concept of the article. Y.-J.S. and S.-W.W. wrote the manuscript with the support of P.-W.W. All authors discussed the results and contributed to the final manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Emanuele, A.S.; Giada, M.; Alessio, F.; Ciprandi, G. From Tissue Repair to Tissue Regeneration. In Wound Healing—Current Perspectives; IntechOpen: London, UK, 2019; Online published. [Google Scholar] [CrossRef] [Green Version]
  2. Galliot, B.; Crescenzi, M.; Jacinto, A.; Tajbakhsh, S. Trends in tissue repair and regeneration. Development 2017, 144, 357–364. [Google Scholar] [CrossRef] [Green Version]
  3. Breda, C.N.S.; Davanzo, G.G.; Basso, P.J.; Saraiva Camara, N.O.; Moraes-Vieira, P.M.M. Mitochondria as central hub of the immune system. Redox Biol. 2019, 26, 101255. [Google Scholar] [CrossRef]
  4. Pearce, E.L.; Pearce, E.J. Metabolic pathways in immune cell activation and quiescence. Immunity 2013, 38, 633–643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. West, A.P.; Brodsky, I.E.; Rahner, C.; Woo, D.K.; Erdjument-Bromage, H.; Tempst, P.; Walsh, M.C.; Choi, Y.; Shadel, G.S.; Ghosh, S. TLR signalling augments macrophage bactericidal activity through mitochondrial ROS. Nat. Cell Biol. 2011, 472, 476–480. [Google Scholar] [CrossRef] [Green Version]
  6. Villa, P.; Cipolla, C.; D’Ippolito, S.; Amar, I.D.; Shachor, M.; Ingravalle, F.; Scaldaferri, F.; Puca, P.; Di Simone, N.; Scambia, G. The interplay between immune system and microbiota in gynecological diseases: A narrative review. Eur. Rev. Med Pharmacol. Sci. 2020, 24, 5676–5690. [Google Scholar] [PubMed]
  7. Faas, M.M.; De Vos, P. Mitochondrial function in immune cells in health and disease. Biochim. Biophys. Acta (BBA) Mol. Basis Dis. 2020, 1866, 165845. [Google Scholar] [CrossRef]
  8. Cheng, S.-C.; Quintin, J.; Cramer, R.A.; Shepardson, K.M.; Saeed, S.; Kumar, V.; Giamarellos-Bourboulis, E.J.; Martens, J.H.A.; Rao, N.A.; Aghajanirefah, A.; et al. mTOR- and HIF-1 -mediated aerobic glycolysis as metabolic basis for trained immunity. Science 2014, 345, 1250684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Buck, M.D.; O’Sullivan, D.; Geltink, R.I.K.; Curtis, J.D.; Chang, C.-H.; Sanin, D.E.; Qiu, J.; Kretz, O.; Braas, D.; Van Der Windt, G.J.; et al. Mitochondrial Dynamics Controls T Cell Fate through Metabolic Programming. Cell 2016, 166, 63–76. [Google Scholar] [CrossRef] [Green Version]
  10. 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. USA 2013, 110, 17963–17968. [Google Scholar] [CrossRef] [Green Version]
  11. Chang, Y.-H.; Lin, H.-Y.; Shen, F.-C.; Su, Y.-J.; Chuang, J.-H.; Lin, T.-K.; Liou, C.-W.; Lin, C.-Y.; Weng, S.-W.; Wang, P.-W. The Causal Role of Mitochondrial Dynamics in Regulating Innate Immunity in Diabetes. Front. Endocrinol. 2020, 11, 445. [Google Scholar] [CrossRef] [PubMed]
  12. Lin, H.Y.; Weng, S.W.; Chang, Y.H.; Su, Y.J.; Chang, C.M.; Tsai, C.J.; Shen, F.C.; Chuang, J.H.; Lin, T.K.; Liou, C.W.; et al. The Causal Role of Mitochondrial Dynamics in Regulating In-sulin Resistance in Diabetes: Link through Mitochondrial Reactive Oxygen Species. Oxid. Med. Cell Longev. 2018, 2018, 7514383. [Google Scholar] [CrossRef] [PubMed]
  13. Yang, L.; Chen, S.; Zhao, Q.; Sun, Y.; Nie, H. The Critical Role of Bach2 in Shaping the Balance be-tween CD4(+) T Cell Subsets in Immune-Mediated Diseases. Mediators Inflamm. 2019, 2019, 2609737. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, G.; Li, X.; Zhang, L.; Abdalla, A.E.; Teng, T.; Li, Y. Crosstalk between Dendritic Cells and Immune Modulatory Agents against Sepsis. Genes 2020, 11, 323. [Google Scholar] [CrossRef] [Green Version]
  15. Chazaud, B. Macrophages: Supportive cells for tissue repair and regeneration. Immunobiology 2014, 219, 172–178. [Google Scholar] [CrossRef] [Green Version]
  16. Parisi, L.; Gini, E.; Baci, D.; Tremolati, M.; Fanuli, M.; Bassani, B.; Farronato, G.; Bruno, A.; Mortara, L. Macrophage Polarization in Chronic Inflammatory Diseases: Killers or Builders? J. Immunol. Res. 2018, 2018, 8917804. [Google Scholar] [CrossRef] [PubMed]
  17. Liu, Y.-C.; Zou, X.-B.; Chai, Y.-F.; Yao, Y.-M. Macrophage Polarization in Inflammatory Diseases. Int. J. Biol. Sci. 2014, 10, 520–529. [Google Scholar] [CrossRef] [PubMed]
  18. Aharoni, S.; Lati, Y.; Aviram, M.; Fuhrman, B. Pomegranate juice polyphenols induce a phenotypic switch in macrophage polarization favoring a M2 anti-inflammatory state. BioFactors 2015, 41, 44–51. [Google Scholar] [CrossRef]
  19. Kittan, N.A.; Allen, R.M.; Dhaliwal, A.; Cavassani, K.A.; Schaller, M.; Gallagher, K.A.; Iv, W.F.C.; Mukherjee, S.; Grembecka, J.; Cierpicki, T.; et al. Cytokine Induced Phenotypic and Epigenetic Signatures Are Key to Establishing Specific Macrophage Phenotypes. PLoS ONE 2013, 8, e78045. [Google Scholar] [CrossRef] [Green Version]
  20. Kasper, M.; Fehrenbach, H. Immunohistochemical evidence for the occurrence of similar epithelial phenotypes during lung development and radiation-induced fibrogenesis. Int. J. Radiat. Biol. 2000, 76, 493–501. [Google Scholar]
  21. Zaiss, D.M.; Minutti, C.M.; Knipper, J.A. Immune- and non-immune-mediated roles of regulatory T-cells during wound healing. Immunology 2019, 157, 190–197. [Google Scholar] [CrossRef] [Green Version]
  22. Henderson, N.C.; Mackinnon, A.C.; Farnworth, S.L.; Kipari, T.; Haslett, C.; Iredale, J.P.; Liu, F.T.; Hughes, J.; Sethi, T. Galectin-3 expression and secretion links macrophages to the promotion of re-nal fibrosis. Am. J. Pathol. 2008, 172, 288–298. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Liu, W.; Hsu, D.K.; Chen, H.Y.; Yang, R.Y.; Carraway, K.L., 3rd; Isseroff, R.R.; Liu, F.T. Galectin-3 regulates intracellular trafficking of EGFR through Alix and promotes keratinocyte migration. J. Invest. Dermatol. 2012, 132, 2828–2837. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Vansthertem, D.; Cludts, S.; Nonclercq, D.; Gossiaux, A.; Saussez, S.; Legrand, A.; Gabius, H.J.; Toubeau, G. Immunohistochemical localization of galectins-1 and -3 and monitoring of tissue ga-lectin-binding sites during tubular regeneration after renal ischemia reperfusion in the rat. Histol. Histopathol. 2010, 25, 1417–1429. [Google Scholar]
  25. Di Gregoli, K.; Somerville, M.; Bianco, R.; Thomas, A.C.; Frankow, A.; Newby, A.C.; George, S.J.; Jackson, C.L.; Johnson, J.L. Galectin-3 Identifies a Subset of Macrophages With a Potential Benefi-cial Role in Atherosclerosis. Arterioscler. Thromb. Vasc. Biol. 2020, 40, 1491–1509. [Google Scholar] [CrossRef] [Green Version]
  26. van der Velde, A.R.; Gullestad, L.; Ueland, T.; Aukrust, P.; Guo, Y.; Adourian, A.; Muntendam, P.; van Veldhuisen, D.J.; de Boer, R.A. Prognostic value of changes in galectin-3 levels over time in pa-tients with heart failure: Data from CORONA and COACH. Circ. Heart Fail 2013, 6, 219–226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Nishiyama, J.; Kobayashi, S.; Ishida, A.; Nakabayashi, I.; Tajima, O.; Miura, S.; Katayama, M.; Nogami, H. Up-Regulation of Galectin-3 in Acute Renal Failure of the Rat. Am. J. Pathol. 2000, 157, 815–823. [Google Scholar] [CrossRef] [Green Version]
  28. Saccon, F.; Gatto, M.; Ghirardello, A.; Iaccarino, L.; Punzi, L.; Doria, A. Role of galectin-3 in auto-immune and non-autoimmune nephropathies. Autoimmun. Rev. 2017, 16, 34–47. [Google Scholar] [CrossRef]
  29. Liu, S.D.; Lee, S.; La Cava, A.; Motran, C.C.; Hahn, B.H.; Miceli, M.C. Galectin-1-induced down-regulation of T lymphocyte activation protects (NZB x NZW) F1 mice from lupus-like disease. Lupus 2011, 20, 473–484. [Google Scholar] [CrossRef] [PubMed]
  30. Moritoki, M.; Kadowaki, T.; Niki, T.; Nakano, D.; Soma, G.; Mori, H.; Kobara, H.; Masaki, T.; Kohno, M.; Hirashima, M. Galectin-9 ameliorates clinical severity of MRL/lpr lupus-prone mice by inducing plasma cell apoptosis independently of Tim-3. PLoS ONE 2013, 8, e60807. [Google Scholar] [CrossRef] [Green Version]
  31. Su, Y.-J.; Cheng, T.-T.; Chen, C.-J.; Chiu, W.-C.; Hsu, C.-Y.; Chang, W.-N.; Tsai, N.-W.; Kung, C.-T.; Wang, H.-C.; Lin, W.-C.; et al. The association among leukocyte apoptosis, autoantibodies and disease severity in systemic lupus erythematosus. J. Transl. Med. 2013, 11, 261. [Google Scholar] [CrossRef] [Green Version]
  32. Su, Y.J.; Lin, I.C.; Wang, L.; Lu, C.H.; Huang, Y.L.; Kuo, H.C. Next generation sequencing identifies miRNA-based biomarker panel for lupus nephritis. Oncotarget 2018, 9, 27911–27919. [Google Scholar] [CrossRef] [PubMed]
  33. Nadella, V.; Wang, Z.; Johnson, T.S.; Griffin, M.; Devitt, A. Transglutaminase 2 interacts with syndecan-4 and CD44 at the surface of human macrophages to promote removal of apoptotic cells. Biochim. Biophys. Acta Bioenerg. 2015, 1853, 201–212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Torigoe, M.; Iwata, S.; Nakayamada, S.; Sakata, K.; Zhang, M.; Hajime, M.; Miyazaki, Y.; Narisawa, M.; Ishii, K.; Shibata, H.; et al. Metabolic Reprogramming Commits Differentiation of Human CD27(+)IgD(+) B Cells to Plasmablasts or CD27(-)IgD(-) Cells. J. Immunol. 2017, 199, 425–434. [Google Scholar] [CrossRef] [Green Version]
  35. Fessler, J.; Fasching, P.; Raicht, A.; Hammerl, S.; Weber, J.; Lackner, A.; Hermann, J.; Dejaco, C.; Graninger, W.B.; Schwinger, W.; et al. Lymphopenia in primary Sjögren’s syndrome is associated with premature aging of naïve CD4+ T cells. Rheumatology 2021, 60, 588–597. [Google Scholar] [CrossRef]
  36. Lau, E.Y.M.; Carroll, E.C.; Callender, L.A.; Hood, G.A.; Berryman, V.; Pattrick, M.; Finer, S.; Hitman, G.A.; Ackland, G.L.; Henson, S.M. Type 2 diabetes is associated with the accumulation of senes-cent T cells. Clin. Exp. Immunol. 2019, 197, 205–213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Pritz, T.; Lair, J.; Ban, M.; Keller, M.; Weinberger, B.; Krismer, M.; Grubeck-Loebenstein, B. Plasma cell numbers decrease in bone marrow of old patients. Eur. J. Immunol. 2014, 45, 738–746. [Google Scholar] [CrossRef] [Green Version]
  38. Urbanczyk, S.; Stein, M.; Schuh, W.; Jäck, H.-M.; Mougiakakos, D.; Mielenz, D. Regulation of Energy Metabolism during Early B Lymphocyte Development. Int. J. Mol. Sci. 2018, 19, 2192. [Google Scholar] [CrossRef] [Green Version]
  39. Mendelsohn, A.R.; Larrick, J.W. Interacting NAD+ and Cell Senescence Pathways Complicate Antiaging Therapies. Rejuvenation Res. 2019, 22, 261–266. [Google Scholar] [CrossRef]
  40. Ruiz Ortega, R.A.; Manzano, L.; Montero-Perez-Barquero, M. Diagnosis of acute heart failure and relevance of biomarkers in elderly patients. Med. Clin. 2014, 142 (Suppl. S1), 20–25. [Google Scholar] [CrossRef]
  41. Oh, S.-J.; Lee, J.K.; Shin, O.S. Aging and the Immune System: The Impact of Immunosenescence on Viral Infection, Immunity and Vaccine Immunogenicity. Immune Netw. 2019, 19, e37. [Google Scholar] [CrossRef]
  42. Bektas, A.; Schurman, S.H.; Gonzalez-Freire, M.; Dunn, C.A.; Singh, A.K.; Macian, F.; Cuervo, A.M.; Sen, R.; Ferrucci, L. Age-associated changes in human CD4(+) T cells point to mitochondrial dysfunction consequent to impaired autophagy. Aging 2019, 11, 9234–9263. [Google Scholar] [CrossRef] [PubMed]
  43. Crooke, S.N.; Ovsyannikova, I.G.; Poland, G.A.; Kennedy, R.B. Immunosenescence and human vaccine immune responses. Immun. Ageing 2019, 16, 25. [Google Scholar] [CrossRef] [Green Version]
  44. Stojanovic, S.D.; Fuchs, M.; Kunz, M.; Xiao, K.; Just, A.; Pich, A.; Bauersachs, J.; Fiedler, J.; Sedding, D.; Thum, T. Inflammatory Drivers of Cardiovascular Disease: Molecular Characterization of Se-nescent Coronary Vascular Smooth Muscle Cells. Front. Physiol. 2020, 11, 520. [Google Scholar] [CrossRef]
  45. Piber, D.; Olmstead, R.; Cho, J.H.-J.; Witarama, T.; Perez, C.; Dietz, N.; Seeman, T.; Breen, E.C.; Cole, S.W.; Irwin, M.R. Inflammaging: Age and Systemic, Cellular, and Nuclear Inflammatory Biology in Older Adults. J. Gerontol. Ser. A Boil. Sci. Med. Sci. 2019, 74, 1716–1724. [Google Scholar] [CrossRef] [PubMed]
  46. Pinti, M.; Cevenini, E.; Nasi, M.; De Biasi, S.; Salvioli, S.; Monti, D.; Benatti, S.; Gibellini, L.; Cotichini, R.; Stazi, M.A.; et al. Circulating mitochondrial DNA increases with age and is a familiar trait: Implications for “inflamm-aging. ” Eur. J. Immunol. 2014, 44, 1552–1562. [Google Scholar] [CrossRef] [PubMed]
  47. Soto-Hermida, A.; Fernandez-Moreno, M.; Pértega-Díaz, S.; Oreiro, N.; Fernández-López, C.; Blanco, F.J.; Rego-Pérez, I. Mitochondrial DNA haplogroups modulate the radiographic progression of Spanish patients with osteoarthritis. Rheumatol. Int. 2014, 35, 337–344. [Google Scholar] [CrossRef]
  48. Weng, S.-W.; Kuo, H.-M.; Chuang, J.-H.; Lin, T.-K.; Huang, H.-L.; Lin, H.-Y.; Liou, C.-W.; Wang, P.-W. Study of insulin resistance in cybrid cells harboring diabetes-susceptible and diabetes-protective mitochondrial haplogroups. Mitochondrion 2013, 13, 888–897. [Google Scholar] [CrossRef]
  49. Fuku, N.; Park, K.S.; Yamada, Y.; Nishigaki, Y.; Cho, Y.M.; Matsuo, H.; Segawa, T.; Watanabe, S.; Kato, K.; Yokoi, K.; et al. Mitochondrial Haplogroup N9a Confers Resistance against Type 2 Diabetes in Asians. Am. J. Hum. Genet. 2007, 80, 407–415. [Google Scholar] [CrossRef] [Green Version]
  50. Aunan, J.R.; Watson, M.M.; Hagland, H.R.; Søreide, K. Molecular and biological hallmarks of ageing. BJS 2016, 103, e29–e46. [Google Scholar] [CrossRef] [Green Version]
  51. De Benedictis, G.; Rose, G.; Carrieri, G.; De Luca, M.; Falcone, E.; Passarino, G.; Bonafe, M.; Monti, D.; Baggio, G.; Bertolini, S.; et al. Mitochondrial DNA inherited var-iants are associated with successful aging and longevity in humans. FASEB J. 1999, 13, 1532–1536. [Google Scholar] [CrossRef]
  52. Franceschi, C.; Garagnani, P.; Parini, P.; Giuliani, C.; Santoro, A. Inflammaging: A new im-mune-metabolic viewpoint for age-related diseases. Nat. Rev. Endocrinol. 2018, 14, 576–590. [Google Scholar] [CrossRef]
  53. Thomas, P.G.; Doherty, P.C. New Approaches in Immunotherapy. Science 2010, 327, 249. [Google Scholar] [CrossRef] [Green Version]
  54. Osborn, O.; Olefsky, J.M. The cellular and signaling networks linking the immune system and metabolism in disease. Nat. Med. 2012, 18, 363–374. [Google Scholar] [CrossRef]
  55. Aravindhan, V.; Madhumitha, H. Metainflammation in Diabetic Coronary Artery Disease: Emerging Role of Innate and Adaptive Immune Responses. J. Diabetes Res. 2016, 2016, 6264149. [Google Scholar] [CrossRef] [Green Version]
  56. Tchkonia, T.; Morbeck, D.E.; Von Zglinicki, T.; Van Deursen, J.; Lustgarten, J.; Scrable, H.; Khosla, S.; Jensen, M.D.; Kirkland, J.L. Fat tissue, aging, and cellular senescence. Aging Cell 2010, 9, 667–684. [Google Scholar] [CrossRef] [PubMed]
  57. Hotamisligil, G.S. Inflammation, metaflammation and immunometabolic disorders. Nat. Cell Biol. 2017, 542, 177–185. [Google Scholar] [CrossRef]
  58. Lin, H.Y.; Weng, S.W.; Shen, F.C.; Chang, Y.H.; Lian, W.S.; Hsieh, C.H.; Chuang, J.H.; Lin, T.K.; Liou, C.W.; Chang, C.S.; et al. Abrogation of Toll-Like Receptor 4 Mitigates Obe-sity-Induced Oxidative Stress, Proinflammation, and Insulin Resistance Through Metabolic Reprogramming of Mitochondria in Adipose Tissue. Antioxid. Redox Signal. 2020, 33, 66–86. [Google Scholar] [CrossRef] [PubMed]
  59. Lopez-Diez, R.; Shekhtman, A.; Ramasamy, R.; Schmidt, A.M. Cellular mechanisms and conse-quences of glycation in atherosclerosis and obesity. Biochim. Biophys. Acta 2016, 1862, 2244–2252. [Google Scholar] [CrossRef] [PubMed]
  60. Gaens, K.H.; Goossens, G.H.; Niessen, P.M.; van Greevenbroek, M.M.; van der Kallen, C.J.; Nies-sen, H.W.; Rensen, S.S.; Buurman, W.A.; Greve, J.W.; Blaak, E.E.; et al. Nepsilon-(carboxymethyl)lysine-receptor for advanced gly-cation end product axis is a key modulator of obesity-induced dysregulation of adipokine expression and insulin resistance. Arterioscler. Thromb. Vasc. Biol. 2014, 34, 1199–1208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Alves, A.S.; Bueno, V. Immunosenescence: Participation of T lymphocytes and myeloid-derived suppressor cells in aging-related immune response changes. Einstein 2019, 17, eRB4733. [Google Scholar] [CrossRef] [Green Version]
  62. Nunnari, J.; Suomalainen, A. Mitochondria: In Sickness and in Health. Cell 2012, 148, 1145–1159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Quirós, P.M.; Mottis, A.; Auwerx, J. Mitonuclear communication in homeostasis and stress. Nat. Rev. Mol. Cell Biol. 2016, 17, 213–226. [Google Scholar] [CrossRef]
  64. Weinberg, S.E.; Sena, L.A.; Chandel, N.S. Mitochondria in the Regulation of Innate and Adaptive Immunity. Immunity 2015, 42, 406–417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Alfonso-Loeches, S.; Urena-Peralta, J.R.; Morillo-Bargues, M.J.; Oliver-De La Cruz, J.; Guerri, C. Role of mitochondria ROS generation in ethanol-induced NLRP3 inflammasome activation and cell death in astroglial cells. Front. Cell Neurosci. 2014, 8, 216. [Google Scholar] [CrossRef]
  66. Wei, P.; Yang, F.; Zheng, Q.; Tang, W.; Li, J. The Potential Role of the NLRP3 Inflammasome Activa-tion as a Link between Mitochondria ROS Generation and Neuroinflammation in Postoperative Cognitive Dysfunction. Front. Cell Neurosci. 2019, 13, 73. [Google Scholar] [CrossRef] [Green Version]
  67. Zhou, R.; Yazdi, A.S.; Menu, P.; Tschopp, J. A role for mitochondria in NLRP3 inflammasome activa-tion. Nature 2011, 469, 221–225. [Google Scholar] [CrossRef]
  68. Chandel, N.S.; Schumacker, P.T.; Arch, R.H. Reactive Oxygen Species Are Downstream Products of TRAF-mediated Signal Transduction. J. Biol. Chem. 2001, 276, 42728–42736. [Google Scholar] [CrossRef] [Green Version]
  69. Pradhan, A.; Avelar, G.M.; Bain, J.M.; Childers, D.S.; Larcombe, D.E.; Netea, M.G.; Shekhova, E.; Munro, C.A.; Brown, G.D.; Erwig, L.P.; et al. Hypoxia Promotes Immune Evasion by Triggering beta-Glucan Masking on the Candida albicans Cell Surface via Mitochondrial and cAMP-Protein Kinase A Signaling. mBio 2018, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. McCully, K.S. Chemical pathology of homocysteine. IV. Excitotoxicity, oxidative stress, endothelial dysfunction, and inflammation. Ann. Clin. Lab. Sci. 2009, 39, 219–232. [Google Scholar]
  71. McCully, J.D.; Cowan, D.B.; Pacak, C.A.; Toumpoulis, I.K.; Dayalan, H.; Levitsky, S. Injection of isolated mitochondria during early reperfusion for cardioprotection. Am. J. Physiol. Heart Circ. Physiol. 2009, 296, H94–H105. [Google Scholar] [CrossRef] [Green Version]
  72. Cowan, D.B.; Yao, R.; Thedsanamoorthy, J.K.; Zurakowski, D.; Del Nido, P.J.; McCully, J.D. Transit and integration of extracellular mitochondria in human heart cells. Sci. Rep. 2017, 7, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. McCully, J.D.; Cowan, D.B.; Emani, S.M.; Del Nido, P.J. Mitochondrial transplantation: From animal models to clinical use in humans. Mitochondrion 2017, 34, 127–134. [Google Scholar] [CrossRef] [PubMed]
  74. Torralba, D.; Baixauli, F.; Sánchez-Madrid, F. Mitochondria Know No Boundaries: Mechanisms and Functions of Intercellular Mitochondrial Transfer. Front. Cell Dev. Biol. 2016, 4, 107. [Google Scholar] [CrossRef] [Green Version]
  75. Dhir, A.; Dhir, S.; Borowski, L.S.; Jimenez, L.; Teitell, M.; Rötig, A.; Crow, Y.J.; Rice, G.I.; Duffy, D.; Tamby, C.; et al. Mitochondrial double-stranded RNA triggers antiviral signalling in humans. Nat. Cell Biol. 2018, 560, 238–242. [Google Scholar] [CrossRef]
  76. Zhong, Z.; Liang, S.; Sanchez-Lopez, E.; He, F.; Shalapour, S.; Lin, X.J.; Wong, J.; Ding, S.; Seki, E.; Schnabl, B.; et al. New mitochondrial DNA syn-thesis enables NLRP3 inflammasome activation. Nature 2018, 560, 198–203. [Google Scholar] [CrossRef] [PubMed]
  77. Fang, C.; Wei, X.; Wei, Y. Mitochondrial DNA in the regulation of innate immune responses. Protein Cell 2016, 7, 11–16. [Google Scholar] [CrossRef] [Green Version]
  78. Lindqvist, D.; Wolkowitz, O.M.; Picard, M.; Ohlsson, L.; Bersani, F.S.; Fernström, J.; Westrin, Å.; Hough, C.M.; Lin, J.; Reus, V.I.; et al. Circulating cell-free mitochondrial DNA, but not leukocyte mitochondrial DNA copy number, is elevated in major depressive disorder. Neuropsychopharmacology 2018, 43, 1557–1564. [Google Scholar] [CrossRef] [Green Version]
  79. Wu, G.; Zhu, Q.; Zeng, J.; Gu, X.; Miao, Y.; Xu, W.; Lv, T.; Song, Y. Extracellular mitochondrial DNA promote NLRP3 inflammasome activation and induce acute lung injury through TLR9 and NF-kappaB. J. Thorac. Dis. 2019, 11, 4816–4828. [Google Scholar] [CrossRef]
  80. Faust, H.E.; Reilly, J.P.; Anderson, B.J.; Ittner, C.A.; Forker, C.M.; Zhang, P.; Weaver, B.A.; Holena, D.N.; Lanken, P.N.; Christie, J.D.; et al. Plasma Mitochondrial DNA Levels Are Associated with ARDS in Trauma and Sepsis Patients. Chest 2020, 157, 67–76. [Google Scholar] [CrossRef] [Green Version]
  81. Collins, L.V.; Hajizadeh, S.; Holme, E.; Jonsson, I.M.; Tarkowski, A. Endogenously oxidized mitochondrial DNA induces in vivo and in vitro inflammatory responses. J. Leukoc. Biol. 2004, 75, 995–1000. [Google Scholar] [CrossRef]
  82. West, A.P.; Shadel, G.S.; Ghosh, S. Mitochondria in innate immune responses. Nat. Rev. Immunol. 2011, 11, 389–402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Randow, F.; Youle, R.J. Self and Nonself: How Autophagy Targets Mitochondria and Bacteria. Cell Host Microbe 2014, 15, 403–411. [Google Scholar] [CrossRef] [Green Version]
  84. Grazioli, S.; Pugin, J. Mitochondrial Damage-Associated Molecular Patterns: From Inflammatory Signaling to Human Diseases. Front. Immunol. 2018, 9, 832. [Google Scholar] [CrossRef] [PubMed]
  85. Zhang, Q.; Raoof, M.; Chen, Y.; Sumi, Y.; Sursal, T.; Junger, W.; Brohi, K.; Itagaki, K.; Hauser, C.J. Circulating mitochondrial DAMPs cause inflammatory responses to injury. Nat. Cell Biol. 2010, 464, 104–107. [Google Scholar] [CrossRef] [Green Version]
  86. Gu, X.; Wu, G.; Yao, Y.; Zeng, J.; Shi, D.; Lv, T.; Luo, L.; Song, Y. Intratracheal administration of mito-chondrial DNA directly provokes lung inflammation through the TLR9-p38 MAPK pathway. Free Radic. Biol. Med. 2015, 83, 149–158. [Google Scholar] [CrossRef]
  87. Itagaki, K.; Kaczmarek, E.; Lee, Y.T.; Tang, I.T.; Isal, B.; Adibnia, Y.; Sandler, N.; Grimm, M.J.; Segal, B.H.; Otterbein, L.E.; et al. Mitochondrial DNA released by trauma induces neutrophil ex-tracellular traps. PLoS ONE 2015, 10, e0120549. [Google Scholar] [CrossRef]
  88. Yu, J.; Nagasu, H.; Murakami, T.; Hoang, H.; Broderick, L.; Hoffman, H.M.; Horng, T. Inflammasome activation leads to Caspase-1-dependent mitochondrial damage and block of mitophagy. Proc. Natl. Acad. Sci. USA 2014, 111, 15514–15519. [Google Scholar] [CrossRef] [Green Version]
  89. White, M.J.; McArthur, K.; Metcalf, D.; Lane, R.M.; Cambier, J.C.; Herold, M.J.; van Delft, M.F.; Bedoui, S.; Lessene, G.; Ritchie, M.E.; et al. Apoptotic caspases suppress mtDNA-induced STING-mediated type I IFN production. Cell 2014, 159, 1549–1562. [Google Scholar] [CrossRef] [Green Version]
  90. Schneider, W.M.; Chevillotte, M.D.; Rice, C.M. Interferon-Stimulated Genes: A Complex Web of Host Defenses. Annu. Rev. Immunol. 2014, 32, 513–545. [Google Scholar] [CrossRef] [Green Version]
  91. Seth, R.B.; Sun, L.; Ea, C.K.; Chen, Z.J. Identification and characterization of MAVS, a mitochondri-al antiviral signaling protein that activates NF-kappaB and IRF 3. Cell 2005, 122, 669–682. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Sudakov, N.P.; Apartsin, K.A.; Lepekhova, S.A.; Nikiforov, S.B.; Katyshev, A.I.; Lifshits, G.I.; Vybi-vantseva, A.V.; Konstantinov, Y.M. The level of free circulating mitochondrial DNA in blood as predictor of death in case of acute coronary syndrome. Eur. J. Med. Res. 2017, 22, 1. [Google Scholar] [CrossRef] [Green Version]
  93. Su, Y.-J.; Cheng, T.-T.; Chen, C.-J.; Chang, W.-N.; Tien-Tsai, C.; Kung, C.-T.; Wang, H.-C.; Lin, W.-C.; Huang, C.-C.; Chang, Y.-T.; et al. Investigation of the caspase-dependent mitochondrial apoptotic pathway in mononuclear cells of patients with systemic lupus erythematosus. J. Transl. Med. 2014, 12, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Miliotis, S.; Nicolalde, B.; Ortega, M.; Yepez, J.; Caicedo, A. Forms of extracellular mitochondria and their impact in health. Mitochondrion 2019, 48, 16–30. [Google Scholar] [CrossRef] [PubMed]
  95. Boudreau, L.H.; Duchez, A.-C.; Cloutier, N.; Soulet, D.; Martin, N.; Bollinger, J.; Paré, A.; Rousseau, M.; Naika, G.S.; Lévesque, T.; et al. Platelets release mitochondria serving as substrate for bactericidal group IIA-secreted phospholipase A2 to promote inflammation. Blood 2014, 124, 2173–2183. [Google Scholar] [CrossRef] [Green Version]
  96. Koupenova, M.; Clancy, L.; Corkrey, H.A.; Freedman, J.E. Circulating Platelets as Mediators of Immunity, Inflammation, and Thrombosis. Circ. Res. 2018, 122, 337–351. [Google Scholar] [CrossRef] [PubMed]
  97. Sunderland, N.; Skroblin, P.; Barwari, T.; Huntley, R.P.; Lu, R.; Joshi, A.; Lovering, R.C.; Mayr, M. MicroRNA Biomarkers and Platelet Reactivity: The Clot Thickens. Circ. Res. 2017, 120, 418–435. [Google Scholar] [CrossRef] [Green Version]
  98. Fejes, Z.; Póliska, S.; Czimmerer, Z.; Káplár, M.; Penyige, A.; Szabó, G.G.; Debreceni, I.B.; Kunapuli, S.P.; Kappelmayer, J.; Nagy, B. Hyperglycaemia suppresses microRNA expression in platelets to increase P2RY12 and SELP levels in type 2 diabetes mellitus. Thromb. Haemost. 2017, 117, 529–542. [Google Scholar] [CrossRef]
  99. Yun, S.H.; Sim, E.H.; Goh, R.Y.; Park, J.I.; Han, J.Y. Platelet Activation: The Mechanisms and Potential Biomarkers. BioMed Res. Int. 2016, 2016, 9060143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Undi, R.B.; Gutti, U.; Gutti, R.K. LiCl regulates mitochondrial biogenesis during megakaryocyte development. J. Trace Elements Med. Biol. 2017, 39, 193–201. [Google Scholar] [CrossRef]
  101. Rasmussen, M.L.; Ortolano, N.A.; Romero-Morales, A.I.; Gama, V. Wnt Signaling and Its Impact on Mitochondrial and Cell Cycle Dynamics in Pluripotent Stem Cells. Genes 2018, 9, 109. [Google Scholar] [CrossRef] [Green Version]
  102. Chuang, Y.-C.; Liou, C.-W.; Chen, S.-D.; Wang, P.-W.; Chuang, J.-H.; Tiao, M.-M.; Hsu, T.-Y.; Lin, H.-Y.; Lin, T.-K. Mitochondrial Transfer from Wharton’s Jelly Mesenchymal Stem Cell to MERRF Cybrid Reduces Oxidative Stress and Improves Mitochondrial Bioenergetics. Oxidative Med. Cell. Longev. 2017, 2017, 1–22. [Google Scholar] [CrossRef] [PubMed]
  103. Kramer, P.A.; Ravi, S.; Chacko, B.; Johnson, M.S.; Darley-Usmar, V.M. A review of the mitochondri-al and glycolytic metabolism in human platelets and leukocytes: Implications for their use as bioenergetic biomarkers. Redox Biol. 2014, 2, 206–210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Sjövall, F.; Ehinger, J.K.; Marelsson, S.E.; Morota, S.; Frostner, E. Åsander; Uchino, H.; Lundgren, J.; Arnbjörnsson, E.; Hansson, M.J.; Fellman, V.; et al. Mitochondrial respiration in human viable platelets—Methodology and influence of gender, age and storage. Mitochondrion 2013, 13, 7–14. [Google Scholar] [CrossRef] [PubMed]
  105. Kholmukhamedov, A.; Jobe, S. Mitochondria and Platelet Cell Death. Thromb. Haemost. 2017, 117, 2207–2208. [Google Scholar] [CrossRef]
  106. Fisar, Z.; Hroudova, J.; Hansikova, H.; Spacilova, J.; Lelkova, P.; Wenchich, L.; Jirak, R.; Zverova, M.; Zeman, J.; Martasek, P.; et al. Mitochondrial Respiration in the Platelets of Patients with Alzheimer’s Disease. Curr. Alzheimer Res. 2016, 13, 930–941. [Google Scholar] [CrossRef]
  107. Cardenes, N.; Corey, C.; Geary, L.; Jain, S.; Zharikov, S.; Barge, S.; Novelli, E.M.; Shiva, S. Platelet bioenergetic screen in sickle cell patients reveals mitochondrial complex V inhibition, which contributes to platelet activation. Blood 2014, 123, 2864–2872. [Google Scholar] [CrossRef] [Green Version]
  108. Takahashi, K.; Tanabe, K.; Ohnuki, M.; Narita, M.; Ichisaka, T.; Tomoda, K.; Yamanaka, S. Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell 2007, 131, 861–872. [Google Scholar] [CrossRef] [Green Version]
  109. Zhao, Y.; Jiang, Z.; Delgado, E.; Li, H.; Zhou, H.; Hu, W.; Perez-Basterrechea, M.; Janostakova, A.; Tan, Q.; Wang, J.; et al. Platelet-Derived Mitochondria Display Embryonic Stem Cell Markers and Improve Pan-creatic Islet beta-cell Function in Humans. Stem Cells Transl. Med. 2017, 6, 1684–1697. [Google Scholar] [CrossRef]
  110. Shi, X.; Zhao, M.; Fu, C.; Fu, A. Intravenous administration of mitochondria for treating experi-mental Parkinson’s disease. Mitochondrion 2017, 34, 91–100. [Google Scholar] [CrossRef]
  111. Shi, X.; Bai, H.; Zhao, M.; Li, X.; Sun, X.; Jiang, H.; Fu, A. Treatment of acetaminophen-induced liver injury with exogenous mitochondria in mice. Transl. Res. 2018, 196, 31–41. [Google Scholar] [CrossRef]
  112. McCully, J.D.; Levitsky, S.; Del Nido, P.J.; Cowan, D.B. Mitochondrial transplantation for therapeutic use. Clin. Transl. Med. 2016, 5, 16. [Google Scholar] [CrossRef] [Green Version]
  113. Spees, J.L.; Olson, S.D.; Whitney, M.J.; Prockop, D.J. Mitochondrial transfer between cells can rescue aerobic respiration. Proc. Natl. Acad. Sci. USA 2006, 103, 1283–1288. [Google Scholar] [CrossRef] [Green Version]
  114. Cselenyak, A.; Pankotai, E.; Horvath, E.M.; Kiss, L.; Lacza, Z. Mesenchymal stem cells rescue car-diomyoblasts from cell death in an in vitro ischemia model via direct cell-to-cell connections. BMC Cell Biol. 2010, 11, 29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Acquistapace, A.; Bru, T.; Lesault, P.F.; Figeac, F.; Coudert, A.E.; le Coz, O.; Christov, C.; Baudin, X.; Auber, F.; Yiou, R.; et al. Human mesenchymal stem cells repro-gram adult cardiomyocytes toward a progenitor-like state through partial cell fusion and mito-chondria transfer. Stem Cells 2011, 29, 812–824. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Islam, M.N.; Das, S.R.; Emin, M.T.; Wei, M.; Sun, L.; Westphalen, K.; Rowlands, D.J.; Quadri, S.K.; Bhattacharya, S.; Bhattacharya, J. Mitochondrial transfer from bone-marrow–derived stromal cells to pulmonary alveoli protects against acute lung injury. Nat. Med. 2012, 18, 759–765. [Google Scholar] [CrossRef] [Green Version]
  117. Figeac, F.; Lesault, P.F.; Le Coz, O.; Damy, T.; Souktani, R.; Trebeau, C.; Schmitt, A.; Ribot, J.; Mounier, R.; Guguin, A.; et al. Nano-tubular crosstalk with distressed cardiomyocytes stimulates the paracrine repair function of mesenchymal stem cells. Stem Cells 2014, 32, 216–230. [Google Scholar] [CrossRef] [PubMed]
  118. Jackson, M.V.; Krasnodembskaya, A.D. Analysis of Mitochondrial Transfer in Direct Co-cultures of Human Monocyte-derived Macrophages (MDM) and Mesenchymal Stem Cells (MSC). Bio-Protocol 2017, 7, e2255. [Google Scholar] [CrossRef] [Green Version]
  119. Cowan, D.B.; Yao, R.; Akurathi, V.; Snay, E.R.; Thedsanamoorthy, J.K.; Zurakowski, D.; Ericsson, M.; Friehs, I.; Wu, Y.; Levitsky, S.; et al. Intracoronary Delivery of Mitochondria to the Ischemic Heart for Cardioprotection. PLoS ONE 2016, 11, e0160889. [Google Scholar] [CrossRef] [Green Version]
  120. Lin, H.Y.; Liou, C.W.; Chen, S.D.; Hsu, T.Y.; Chuang, J.H.; Wang, P.W.; Huang, S.T.; Tiao, M.M.; Chen, J.B.; Lin, T.K.; et al. Mitochondrial transfer from Wharton’s jelly-derived mesen-chymal stem cells to mitochondria-defective cells recaptures impaired mitochondrial function. Mitochondrion 2015, 22, 31–44. [Google Scholar] [CrossRef] [PubMed]
  121. Bagheri, H.S.; Bani, F.; Tasoglu, S.; Zarebkohan, A.; Rahbarghazi, R.; Sokullu, E. Mitochondrial donation in translational medicine; from imagination to reality. J. Transl. Med. 2020, 18, 367. [Google Scholar] [CrossRef]
  122. Wang, J.; Li, H.; Yao, Y.; Zhao, T.; Chen, Y.Y.; Shen, Y.L.; Wang, L.L.; Zhu, Y. Stem cell-derived mitochondria transplantation: A novel strategy and the challenges for the treatment of tissue injury. Stem Cell Res. Ther. 2018, 9, 106. [Google Scholar] [CrossRef] [Green Version]
  123. Hayakawa, K.; Esposito, E.; Wang, X.; Terasaki, Y.; Liu, Y.; Xing, C.; Ji, X.; Lo, E.H. Transfer of mitochondria from astrocytes to neurons after stroke. Nature 2016, 535, 551–555. [Google Scholar] [CrossRef] [Green Version]
  124. Moschoi, R.; Imbert, V.; Nebout, M.; Chiche, J.; Mary, D.; Prebet, T.; Saland, E.; Castellano, R.; Pouyet, L.; Collette, Y.; et al. Protective mitochondrial transfer from bone marrow stromal cells to acute myeloid leukemic cells during chemotherapy. Blood 2016, 128, 253–264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Singh, A.K.; Cancelas, J.A. Mitochondria Transfer in Bone Marrow Hematopoietic Activity. Curr. Stem Cell Rep. 2021, 1–12. [Google Scholar] [CrossRef]
  126. Zhang, J.Z.; Wang, J.; Qu, W.C.; Wang, X.W.; Liu, Z.; Ren, J.X.; Han, L.; Sun, T.S. Plasma mitochondrial DNA levels were independently associated with lung injury in elderly hip fracture patients. Injury 2017, 48, 454–459. [Google Scholar] [CrossRef] [PubMed]
  127. Hu, Q.; Ren, J.; Wu, J.; Li, G.; Wu, X.; Liu, S.; Wang, G.; Gu, G.; Li, J. Elevated Levels of Plasma Mito-chondrial DNA Are Associated with Clinical Outcome in Intra-Abdominal Infections Caused by Se-vere Trauma. Surg. Infect. 2017, 18, 610–618. [Google Scholar] [CrossRef]
  128. Tuboly, E.; McLlroy, D.; Briggs, G.; Lott, N.; Balogh, Z.J. Clinical implications and pathological associations of circulating mitochondrial DNA. Front. Biosci. 2017, 22, 1011–1022. [Google Scholar]
  129. Yamanouchi, S.; Kudo, D.; Yamada, M.; Miyagawa, N.; Furukawa, H.; Kushimoto, S. Plasma mitochondrial DNA levels in patients with trauma and severe sepsis: Time course and the association with clinical status. J. Crit. Care 2013, 28, 1027–1031. [Google Scholar] [CrossRef]
  130. Xie, L.; Liu, S.; Cheng, J.; Wang, L.; Liu, J.; Gong, J. Exogenous administration of mitochondrial DNA promotes ischemia reperfusion injury via TLR9-p38 MAPK pathway. Regul. Toxicol. Pharmacol. 2017, 89, 148–154. [Google Scholar] [CrossRef]
  131. Zhang, Q.; Itagaki, K.; Hauser, C.J. Mitochondrial DNA is released by shock and activates neutrophils via p38 map kinase. Shock 2010, 34, 55–59. [Google Scholar] [CrossRef]
  132. Martinez-Quinones, P.A.; McCarthy, C.G.; Mentzer, C.J.; Wenceslau, C.F.; Holsten, S.B.; Webb, R.C.; O’Malley, K. Peritoneal cavity lavage reduces the presence of mitochondrial damage associated molecular patterns in open abdomen patients. J. Trauma Acute Care Surg. 2017, 83, 1062–1065. [Google Scholar] [CrossRef]
  133. Sandler, N.; Kaczmarek, E.; Itagaki, K.; Zheng, Y.; Otterbein, L.; Khabbaz, K.; Liu, D.; Senthilnathan, V.; Gruen, R.L.; Hauser, C.J. Mitochondrial DAMPs Are Released During Cardiopulmonary Bypass Surgery and Are Associated With Postoperative Atrial Fibrillation. Hear. Lung Circ. 2018, 27, 122–129. [Google Scholar] [CrossRef] [Green Version]
  134. Casa, L.D.; Deaton, D.H.; Ku, D.N. Role of high shear rate in thrombosis. J. Vasc. Surg. 2015, 61, 1068–1080. [Google Scholar] [CrossRef] [Green Version]
  135. McIlroy, D.J.; Bigland, M.; White, A.E.; Hardy, B.M.; Lott, N.; Smith, D.W.; Balogh, Z.J. Cell necrosis-independent sustained mitochondrial and nuclear DNA release following trauma surgery. J. Trauma Acute Care Surg. 2015, 78, 282–288. [Google Scholar] [CrossRef] [PubMed]
  136. Chouchani, E.T.; Pell, V.R.; James, A.M.; Work, L.M.; Saeb-Parsy, K.; Frezza, C.; Krieg, T.; Murphy, M.P. A Unifying Mechanism for Mitochondrial Superoxide Production during Ischemia-Reperfusion Injury. Cell Metab. 2016, 23, 254–263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Hausenloy, D.J.; Botker, H.E.; Engstrom, T.; Erlinge, D.; Heusch, G.; Ibanez, B.; Kloner, R.A.; Ovize, M.; Yellon, D.M.; Garcia-Dorado, D. Targeting reperfusion injury in patients with ST-segment ele-vation myocardial infarction: Trials and tribulations. Eur. Heart J. 2017, 38, 935–941. [Google Scholar] [PubMed] [Green Version]
  138. Heusch, G.; Gersh, B.J. The pathophysiology of acute myocardial infarction and strategies of protection beyond reperfusion: A continual challenge. Eur. Heart J. 2017, 38, 774–784. [Google Scholar] [CrossRef]
  139. Martin, J.L.; Costa, A.S.H.; Gruszczyk, A.V.; Beach, T.E.; Allen, F.M.; Prag, H.A.; Hinchy, E.C.; Mahbubani, K.; Hamed, M.; Tronci, L.; et al. Succinate accumulation drives ischaemia-reperfusion injury during organ transplantation. Nat. Metab. 2019, 1, 966–974. [Google Scholar] [CrossRef] [Green Version]
  140. Nagakawa, K.; Soyama, A.; Hidaka, M.; Adachi, T.; Ono, S.; Hara, T.; Takatsuki, M.; Eguchi, S. Elevated Plasma Levels of Mitochondria-Derived Damage-Associated Molecular Patterns during Liver Transplantation: Predictors for Postoperative Multi-Organ Dysfunction Syndrome. Tohoku J. Exp. Med. 2020, 250, 87–93. [Google Scholar] [CrossRef] [Green Version]
  141. Martin, J.L.; Gruszczyk, A.V.; Beach, T.E.; Murphy, M.P.; Saeb-Parsy, K. Mitochondrial mechanisms and therapeutics in ischaemia reperfusion injury. Pediatr. Nephrol. 2019, 34, 1167–1174. [Google Scholar] [CrossRef] [Green Version]
  142. Pollara, J.; Edwards, R.W.; Lin, L.; Bendersky, V.A.; Brennan, T.V. Circulating mitochondria in deceased organ donors are associated with immune activation and early allograft dysfunction. JCI Insight 2018, 3. [Google Scholar] [CrossRef]
  143. Jansen, M.P.B.; Pulskens, W.P.C.; Uil, M.; Claessen, N.; Nieuwenhuizen, G.; Standaar, D.; Hau, C.M.; Nieuwland, R.; Florquin, S.; Bemelman, F.J.; et al. Urinary mitochondrial DNA associates with delayed graft function following renal transplantation. Nephrol. Dial. Transplant. 2018, 35, 1320–1327. [Google Scholar] [CrossRef]
  144. Kim, K.; Moon, H.; Lee, Y.H.; Seo, J.-W.; Kim, Y.G.; Moon, J.-Y.; Kim, J.S.; Jeong, K.-H.; Lee, T.W.; Ihm, C.-G.; et al. Clinical relevance of cell-free mitochondrial DNA during the early postoperative period in kidney transplant recipients. Sci. Rep. 2019, 9, 1–9. [Google Scholar] [CrossRef] [PubMed]
  145. Lin, L.; Xu, H.; Bishawi, M.; Feng, F.; Samy, K.; Truskey, G.; Barbas, A.S.; Kirk, A.D.; Brennan, T.V. Circulating mitochondria in organ donors promote allograft rejection. Arab. Archaeol. Epigr. 2019, 19, 1917–1929. [Google Scholar] [CrossRef] [PubMed]
  146. Wang, P.-W.; Kuo, H.-M.; Huang, H.-T.; Chang, A.Y.W.; Weng, S.-W.; Tai, M.-H.; Chuang, J.-H.; Chen, I.-Y.; Huang, S.-C.; Lin, T.-K.; et al. Biphasic Response of Mitochondrial Biogenesis to Oxidative Stress in Visceral Fat of Diet-Induced Obesity Mice. Antioxid. Redox Signal. 2014, 20, 2572–2588. [Google Scholar] [CrossRef]
  147. Shen, F.-C.; Weng, S.-W.; Tsao, C.-F.; Lin, H.-Y.; Chang, C.-S.; Lin, C.-Y.; Lian, W.-S.; Chuang, J.-H.; Lin, T.-K.; Liou, C.-W.; et al. Early intervention of N-acetylcysteine better improves insulin resistance in diet-induced obesity mice. Free. Radic. Res. 2018, 52, 1296–1310. [Google Scholar] [CrossRef] [PubMed]
  148. Owen, M.R.; Doran, E.; Halestrap, A.P. Evidence that metformin exerts its anti-diabetic effects through inhibition of complex 1 of the mitochondrial respiratory chain. Biochem. J. 2000, 348 Pt 3, 607–614. [Google Scholar] [CrossRef]
  149. Anabtawi, A.; Miles, J.M. Metformin: Nonglycemic Effects and Potential Novel Indications. Endocr. Pract. 2016, 22, 999–1007. [Google Scholar] [CrossRef]
  150. Barzilai, N.; Crandall, J.P.; Kritchevsky, S.B.; Espeland, M.A. Metformin as a Tool to Target Aging. Cell Metab. 2016, 23, 1060–1065. [Google Scholar] [CrossRef] [Green Version]
  151. Pollak, M. The effects of metformin on gut microbiota and the immune system as research frontiers. Diabetologia 2017, 60, 1662–1667. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Tomczynska, M.; Bijak, M.; Saluk, J. Metformin—The Drug for the Treatment of Autoimmune Diseases; A New Use of a Known Anti-Diabetic Drug. Curr. Top Med. Chem. 2016, 16, 2223–2230. [Google Scholar] [CrossRef]
  153. Foretz, M.; Guigas, B.; Viollet, B. Understanding the glucoregulatory mechanisms of metformin in type 2 diabetes mellitus. Nat. Rev. Endocrinol. 2019, 15, 569–589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Kelly, B.; Tannahill, G.M.; Murphy, M.P.; O’Neill, L.A. Metformin Inhibits the Production of Reactive Oxygen Species from NADH:Ubiquinone Oxidoreductase to Limit Induction of Interleukin-1beta (IL-1beta) and Boosts Interleukin-10 (IL-10) in Lipopolysaccharide (LPS)-activated Macrophages. J. Biol. Chem. 2015, 290, 20348–20359. [Google Scholar] [CrossRef] [Green Version]
  155. Lee, H.-M.; Kim, J.-J.; Kim, H.J.; Shong, M.; Ku, B.J.; Jo, E.-K. Upregulated NLRP3 Inflammasome Activation in Patients with Type 2 Diabetes. Diabetes 2013, 62, 194–204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Wang, H.; Li, T.; Chen, S.; Gu, Y.; Ye, S. Neutrophil Extracellular Trap Mitochondrial DNA and Its Autoantibody in Systemic Lupus Erythematosus and a Proof-of-Concept Trial of Metformin. Arthritis Rheumatol. 2015, 67, 3190–3200. [Google Scholar] [CrossRef] [Green Version]
  157. Su, Y.J.; Lin, H.Y.; Weng, S.W.; Lu, C.H.; Lin, C.Y.; Chiu, W.C.; Wang, P.W. Metformin Represses Interferonopathy Through Suppression of Melanoma Differentiation-Associated Protein 5 and Mitochondrial Antiviral Signaling Protein Activation: Comment on the Article by Wang et al. Arthritis Rheumatol. 2016, 68, 3042–3043. [Google Scholar] [CrossRef] [PubMed]
  158. Athauda, D.; Foltynie, T. The glucagon-like peptide 1 (GLP) receptor as a therapeutic target in Parkinson’s disease: Mechanisms of action. Drug Discov. Today 2016, 21, 802–818. [Google Scholar] [CrossRef] [Green Version]
  159. Germano, J.D.F.; Huang, C.; Sin, J.; Song, Y.; Tucker, K.C.; Taylor, D.J.R.; Saadaeijahromi, H.; Stotland, A.; Piplani, H.; Gottlieb, R.A.; et al. Intermittent Use of a Short-Course Glucagon-like Peptide-1 Receptor Agonist Therapy Limits Adverse Cardiac Remodeling via Parkin-dependent Mitochondrial Turnover. Sci. Rep. 2020, 10, 1–13. [Google Scholar] [CrossRef] [PubMed]
  160. Goralska, J.; Sliwa, A.; Gruca, A.; Razny, U.; Chojnacka, M.; Polus, A.; Solnica, B.; Malczewska-Malec, M. Glucagon-like peptide-1 receptor agonist stimulates mitochondrial bioenergetics in human adipocytes. Acta Biochim. Pol. 2017, 64, 423–429. [Google Scholar] [CrossRef]
  161. Blanco, F.J.; Valdes, A.M.; Rego-Perez, I. Mitochondrial DNA variation and the pathogenesis of osteoarthritis phenotypes. Nat. Rev. Rheumatol. 2018, 14, 327–340. [Google Scholar] [CrossRef]
Figure 1. Summary of the cooperation of mitochondria (central circle) in controlling the balance between immune cells in homeostasis (green box), inflammageing (red box), and regeneration (blue box) status. Figure legends: immune cells or tissue cells in either homeostasis (green box), ageing (red box), and regeneration (blue box) status have different phenotypes and mitochondrial markers, which represent the interactions between the cell itself and its environment. The mitochondrial markers represent different mitochondrial DNA activation in each status, homeostasis (green), ageing (red), and regeneration (blue). The proteins shown inside the hexagonal boxes could be either found in regenerative status or in the inflammatory status, respectively. Abbreviations: TGF-beta, transforming growth factor beta; MFN1, mitofusin-1; MFN2, mitofusin-2; OPA1, optic atrophy-1; MAVS, mitochondrial antiviral-signaling protein; CMPK2, cytidine/uridine monophosphate kinase 2; OCT4, octamer-binding transcription factor 4; SOX2, SRY-Box transcription factor 2; KLF4, Kruppel like factor 4; C-MYC, cellular-Myelocytomatosis; MCP-1, monocyte chemoattractant protein-1; MMP, matrix metallopeptidases; STAT, signal transducer and activator of transcription; CCL-2, chemokine (C-C motif) ligand 2; PTGS2, prostaglandin-endoperoxide synthase 2.
Figure 1. Summary of the cooperation of mitochondria (central circle) in controlling the balance between immune cells in homeostasis (green box), inflammageing (red box), and regeneration (blue box) status. Figure legends: immune cells or tissue cells in either homeostasis (green box), ageing (red box), and regeneration (blue box) status have different phenotypes and mitochondrial markers, which represent the interactions between the cell itself and its environment. The mitochondrial markers represent different mitochondrial DNA activation in each status, homeostasis (green), ageing (red), and regeneration (blue). The proteins shown inside the hexagonal boxes could be either found in regenerative status or in the inflammatory status, respectively. Abbreviations: TGF-beta, transforming growth factor beta; MFN1, mitofusin-1; MFN2, mitofusin-2; OPA1, optic atrophy-1; MAVS, mitochondrial antiviral-signaling protein; CMPK2, cytidine/uridine monophosphate kinase 2; OCT4, octamer-binding transcription factor 4; SOX2, SRY-Box transcription factor 2; KLF4, Kruppel like factor 4; C-MYC, cellular-Myelocytomatosis; MCP-1, monocyte chemoattractant protein-1; MMP, matrix metallopeptidases; STAT, signal transducer and activator of transcription; CCL-2, chemokine (C-C motif) ligand 2; PTGS2, prostaglandin-endoperoxide synthase 2.
Ijms 22 02668 g001
Table 1. Summary of the immune cells or other cells/molecules in different states of mitochondrial condition in this review article.
Table 1. Summary of the immune cells or other cells/molecules in different states of mitochondrial condition in this review article.
MacrophageT CellB CellCombined Immune Cells or OthersMitochondrial activity
Different phases of tissue conditionsInflammationMacrophage polarization in inflammatory diseases [16,17].Down-regulation of T lymphocyte activation protects (NZB x NZW) F1 mice from lupus-like disease [29].Plasma cell inducing clinical severity of MRL/lpr lupus-prone mice [30].Leukocyte apoptosis, autoantibodies and disease severity in systemic lupus erythematosus [25,31,32].
Glycolysis [3]Glycolysis [3]Glycolysis [3]Increased reactive oxygen species [3]
Repair/RegenerationMacrophages: supportive cells for tissue repair and regeneration [15,20,33].Immune mediated roles of regulatory T-cells during wound healing [21].Metabolic reprogramming commits differentiation of human B cells [34]Trends in tissue repair and regeneration, phagocytosis of apoptotic cells [2,22,23,24]
Fatty acid oxidation [3]Fatty acid oxidation [3]Decreased reactive oxygen species [3]Mitochondrial fission [3]
AgeingSimilar epithelial phenotypes during lung development and radiation-induced fibrogenesis [20].Type 2 diabetes is associated with the accumulation of senescent T cells [35,36].Regulation of energy metabolism during early B lymphocyte development [37,38]Macrophages with a potential beneficial role in atherosclerosis [26,39,40].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Su, Y.-J.; Wang, P.-W.; Weng, S.-W. The Role of Mitochondria in Immune-Cell-Mediated Tissue Regeneration and Ageing. Int. J. Mol. Sci. 2021, 22, 2668. https://doi.org/10.3390/ijms22052668

AMA Style

Su Y-J, Wang P-W, Weng S-W. The Role of Mitochondria in Immune-Cell-Mediated Tissue Regeneration and Ageing. International Journal of Molecular Sciences. 2021; 22(5):2668. https://doi.org/10.3390/ijms22052668

Chicago/Turabian Style

Su, Yu-Jih, Pei-Wen Wang, and Shao-Wen Weng. 2021. "The Role of Mitochondria in Immune-Cell-Mediated Tissue Regeneration and Ageing" International Journal of Molecular Sciences 22, no. 5: 2668. https://doi.org/10.3390/ijms22052668

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop